Green Hydrogen in Power Systems (Green Energy and Technology) [2024 ed.] 3031524284, 9783031524288

​Green Hydrogen in Power Systems examines state-of-the-art applications and the latest developments in technology, proto

123 36 13MB

English Pages 367 [363] Year 2024

Report DMCA / Copyright

DOWNLOAD PDF FILE

Table of contents :
Preface
Contents
Chapter 1: An Overview of Energy and Exergy Analysis for Green Hydrogen Power Systems
1.1 Introduction
1.1.1 Green Hydrogen as a Potential Source of Clean Energy
1.2 Hydrogen Economy
1.3 Economic and Environmental Effects of GH2 Production
1.3.1 Economic Issues
1.3.1.1 Expensive Production
1.3.1.2 Investment Needed
1.3.1.3 Infrastructure Limitations
1.3.2 Environmental Issues
1.3.2.1 Carbon Emissions
1.3.2.2 Land Use
1.3.2.3 Water Use
1.4 GH2 Production Methods and Explanation of How Electrolysis Works
1.4.1 Electrolysis
1.4.2 Fuel Cells
1.5 Energy Crisis
1.6 Integration of GH2 Systems with Renewable Energy Sources and Energy Hub
1.7 Future of Energy Supply Systems and Related Works
1.8 Energy and Exergy Analysis
1.9 Related Works
1.10 Conclusion
References
Chapter 2: Hydrogen-Incorporated Sector-Coupled Smart Grids: A Systematic Review and Future Concepts
2.1 Introduction
2.2 Fundamentals of Hydrogen Integration in Smart Grids
2.2.1 Introduction to Hydrogen as an Energy Carrier
2.2.2 Benefits and Challenges of Hydrogen Integration
2.2.3 Benefits of Hydrogen Integration
2.2.3.1 Decarbonization
2.2.3.2 Energy Storage
2.2.3.3 Versatility and Flexibility
2.2.3.4 Energy Independence and Security
2.2.3.5 Air Quality Improvement
2.2.4 Challenges of Hydrogen Integration
2.2.4.1 Cost and Infrastructure
2.2.4.2 Hydrogen Production
2.2.4.3 Storage and Transportation
2.2.4.4 System Integration
2.2.4.5 Market Development and Regulations
2.3 Sector Coupling in Smart Grids
2.3.1 Definition and Principles of Sector Coupling
2.3.1.1 Digitalization and Automation
2.3.1.2 Decentralization and Local Energy Systems
2.3.2 Power-to-X in Sector-Coupled Smart Grids
2.3.2.1 Power-to-Hydrogen (PtH2)
2.3.2.2 Power-to-Gas (PtG)
2.3.2.3 Power-to-Liquid (PtL)
2.3.2.4 Power-to-Heat (PtH)
2.3.2.5 Power-to-Mobility
2.3.2.6 Hydrogen-to-Power (H2-to-Power)
2.3.2.7 Hydrogen-to-Gas (H2-to-Gas)
2.3.2.8 Hydrogen-to-Industry
2.3.2.9 Hydrogen-to-Chemical
2.3.2.10 Hydrogen-to-Fuel
2.3.3 Energy Management
2.3.3.1 Demand-Side Management (DSM)
2.3.3.2 Demand Response
2.3.3.3 System Optimization
2.3.3.4 Energy Market Integration
2.3.4 Power Market
2.3.5 Hydrogen Economy
2.4 Hydrogen-Incorporated Smart Grid Projects
2.4.1 Haeolus Project (Denmark)
2.4.2 SmartPowerFlow Project (Germany)
2.4.3 HyEnergy Project (Netherlands)
2.4.4 H2Future Project (Austria)
2.4.5 Hydrogen Link Project (Japan)
2.5 Technological Advancements and Innovations
2.5.1 Emerging Technologies for Hydrogen Production, Storage, and Distribution
2.5.1.1 Advanced Electrolysis Technologies
2.5.1.2 High-Temperature Electrolysis (HTE)
2.5.1.3 Low-Temperature Electrolysis (LTE)
2.5.1.4 Bipolar Membrane Electrolysis
2.5.2 Renewable Hydrogen
2.5.2.1 Biomass Gasification
2.5.2.2 Photoelectrochemical (PEC) Water Splitting
2.5.3 Hydrogen Storage
2.5.3.1 Liquid Organic Hydrogen Carriers (LOHCs)
2.5.3.2 Metal Hydrides
2.5.3.3 Nanostructured Materials
2.5.4 Hydrogen Distribution
2.5.4.1 Hydrogen Pipelines
2.5.4.2 Liquid Hydrogen Carriers
2.5.5 Hydrogen Infrastructure
2.5.5.1 Hydrogen Refueling Stations
2.5.5.2 Hydrogen Hubs and Networks
2.5.6 Smart Grid Control and Management Systems for Efficient Hydrogen Utilization
2.5.6.1 Demand Response and Load Management
2.5.6.2 Energy Management Systems
2.5.6.3 Grid Integration and Power Balancing
2.5.6.4 Intelligent Monitoring and Predictive Maintenance
2.5.6.5 Decentralized Control and Peer-to-Peer (P2P) Energy Trading
2.6 Economic and Environmental Considerations
2.6.1 Cost Analysis of Hydrogen-Integrated Smart Grid Systems
2.6.2 Evaluation of the Environmental Impacts and Sustainability Aspects
2.6.3 Economic and Financial Incentives for Promoting Hydrogen Integration
2.6.3.1 Carbon Pricing and Emissions Trading
2.6.3.2 Research and Development Funding
2.6.3.3 Financing Mechanisms
2.7 Future Prospects and Research Directions
2.7.1 Identification of Key Challenges and Gaps in Current Hydrogen-Integrated Smart Grid Systems
2.7.1.1 Cost-Effectiveness
2.7.1.2 Scalability and Infrastructure Development
2.7.1.3 Technology Efficiency and Reliability
2.7.1.4 Safety Considerations
2.7.2 Exploration of Future Prospects and Potential Advancements
2.7.2.1 Advanced Electrolysis Technologies
2.7.2.2 Renewable Hydrogen Sources
2.7.2.3 Energy Storage and Conversion
2.7.2.4 Hydrogen Grid Integration
2.7.3 Research Directions and Areas for Further Investigation
2.7.3.1 Techno-Economic Analysis
2.7.3.2 Lifecycle Assessments and Sustainability
2.7.3.3 System Modeling and Optimization
2.7.3.4 Demonstration Projects and Real-World Applications
2.8 Conclusion
References
Chapter 3: Techno-Economic Analysis for Centralized GH2 Power Systems
3.1 Introduction
3.2 Energy Democracy in Energy Communities
3.3 Market Design in Peer-to-Peer Energy Trading
3.3.1 Centralized Peer-to-Peer Energy Market
3.3.2 Role of the Aggregators in Energy Communities
3.3.3 Prosumer Preferences in Energy Communities
3.4 Feed-in-Tariffs (FiT) Mechanism in Energy Trading Market
3.5 GH2 System Modeling
3.6 Problem Formulation
3.7 Simulation and Numerical Results
3.7.1 Input Data
3.7.2 Simulations Results
3.7.2.1 Optimal Operation of Centralized P2P Trading
3.7.2.2 Optimal Coalition Operation of Centralized P2P Transactions
3.8 Conclusion
References
Chapter 4: Techno-Economic Analysis for Decentralized GH2 Power Systems
4.1 Introduction
4.2 Technological and Policy Approaches for the Integration of GH2 Resources in Renewable Energy Systems
4.3 Peer-to-Peer (P2P) Energy Trading Concept
4.3.1 Decentralized P2P Trading Mechanism
4.4 Hydrogen Energy Storage System (HESS)
4.5 Problem Formulation
4.5.1 Cost Modeling
4.5.1.1 Power Balance Constraint
4.5.1.2 Plug-in Electric Vehicle Modeling
4.5.1.3 Hydrogen Storage Unit Modeling
4.5.1.4 ADMM Algorithm Implementation
4.6 Result and Analysis
4.6.1 Model Implementation and Data
4.7 Conclusion
References
Chapter 5: Hydrogenation from Renewable Energy Sources for Developing a Carbon-Free Society: Methods, Real Cases, and Standards
5.1 Status Quo, Challenges, and Outlook
5.2 Hydrogenation Using Renewable Energy
5.2.1 Water Splitting
5.2.1.1 Hydrogen Production Using Electrolysis Method
5.2.1.1.1 Advantages and Disadvantages of Electrolyzer Methods
5.2.1.2 Hydrogen Production Using Photocatalyst Method
5.2.1.3 Hydrogen Production Using Thermolysis of Water
5.2.2 Biohydrogen (Production of Hydrogen from Biomass)
5.2.2.1 Biological Methods
5.2.2.1.1 Water Photolysis
5.2.2.1.2 Photo-Fermentation
5.2.2.1.3 Dark-Fermentation
5.2.2.1.4 Dark-Photo Co-fermentation
5.2.2.2 Chemical Methods
5.2.2.2.1 Pyrolysis Reforming
5.2.2.2.2 Supercritical Water Conversion
5.2.2.2.3 Biomass Gasification
5.2.2.2.4 Catalytic Reforming of Small Organic Molecules
5.3 Applications of Hydrogen in Power and Energy Systems
5.3.1 Facilitating the Operation of Renewable Resources and Their Integration
5.3.2 Reducing the Emission of Polluting Gases
5.3.3 Fuel Cell
5.3.4 Use of Green Hydrogen in the Transportation System
5.4 Hydrogen Storage Technologies and Issues and Problems of Hydrogen Storage and Transfer
5.4.1 Storage Systems
5.4.1.1 Hydrogen Storage by Gas or Compression
5.4.1.2 Hydrogen Storage by Liquefaction
5.4.1.3 Hydrogen Storage by Physical Absorption
5.4.1.3.1 Carbon Materials
5.4.1.4 Hydrogen Storage with Chemical Absorption
5.4.2 The Best Hydrogen Storage Technology
5.4.3 Hydrogen Transmission and Distribution Technology
5.4.3.1 Transmission by Pipelines
5.4.3.2 Transportation by Road and Rail or by Sea
5.5 Green Hydrogen Standard
5.5.1 Definition of Green Hydrogen
5.5.2 System Boundary
5.5.3 Greenhouse Gases Emissions Threshold
5.5.4 Qualification Level and Qualifying Technical Route
5.6 Conclusion
References
Chapter 6: The Role of Green Hydrogen in Achieving Low and Net-Zero Carbon Emissions: Climate Change and Global Warming
6.1 Overview and General Background Outlook
6.2 Low and Net-Zero Emission Clarification
6.2.1 Barriers and Required Infrastructure for Reaching Low-Net-Zero Emission
6.2.2 How Penalties and Taxes on Available Technology Can Lead to Low-Net-Zero Emission
6.3 Road to Carbon Neutrality
6.3.1 Carbon Capture, Storage, and Utilization Versus Pollution Emitting
6.3.2 Economic and Environmental Aspects of Carbon Neutrality
6.4 Green Hydrogen Explanation and Definition
6.4.1 How Can We Pass Through Different Types of H2 to GH2
6.4.2 Status Quo, Challenges, and Outlook to Achieving and Developing GH2: Infrastructure, Limits, and Policies
6.4.3 The Intermittency Effect of Renewable Units for Transition to GH2
6.5 How Can GH2 Help to Achieve Low and Net-Zero Emissions?
6.5.1 Utilizing GH2 in the Residential and Mobility Sectors: GH2 Route in Energy Sectors
6.5.2 Future Scopes and Effects of GH2 on Pollution Emission
6.6 Renewable-Based Units´ Role in Smoothing the Way to Reach Low-Net-Zero Emission: Clean and Adequate GH2 Generation
6.7 Numerical Analysis of GH2
6.8 Summary and Conclusion
References
Chapter 7: Bioreactor Design Selection for Biohydrogen Production Using Immobilized Cell Culture System
7.1 Introduction
7.2 Immobilization of Microbial Culture in Biohydrogen Production
7.2.1 Cell Entrapment
7.2.2 Adsorption
7.2.3 Encapsulation
7.2.4 Containment Within Synthetic Polymers
7.3 Types of Bioreactors
7.3.1 Bioreactor for Batch System
7.3.2 Bioreactor for Continuous System
7.3.3 Bioreactor Performance
7.4 Status Quo, Challenges, and Outlook
References
Chapter 8: Biomass-Based Polygeneration Systems with Hydrogen Production: A Concise Review and Case Study
8.1 Introduction
8.1.1 Biomass-Driven Polygeneration Systems
8.1.2 Hydrogen
8.1.3 Thermodynamic Evaluation of Polygeneration Systems
8.1.4 The Objective of This Work
8.2 Hydrogen Production in Biomass-Based Polygeneration Systems
8.2.1 Biomass as a Fuel for Driving Hydrogen Production Unit
8.2.2 Biomass as a Feedstock for Hydrogen Production
8.3 Case Study
8.3.1 System Description
8.3.2 Materials and Methods
8.3.2.1 Energy Analysis
8.3.2.2 Exergy Analysis
8.3.2.3 Exergoeconomic Analysis
8.3.2.4 Exergoenvironmental Analysis
8.3.3 Results and Discussion
8.4 Status Quo, Challenges, and Outlook
8.5 Conclusions
References
Chapter 9: Integration of Solar PV and GH2 in the Future Power Systems
9.1 Introduction
9.2 Status Quo, Challenges, and Outlook
9.3 Related Works
9.4 Operation of Solar and GH2 Integrated Energy System
9.4.1 Solar Panels
9.4.2 Inverter
9.4.3 Electrolyzers
9.4.4 Fuel Cells
9.4.5 Hydrogen Storage Tanks
9.4.6 Objective Function
9.4.7 Constraints
9.4.8 Case Study
9.5 Planning of Solar and Gh2 Integrated Energy System
9.5.1 Reliability Indexes
9.5.2 Objective Function
9.5.3 Constraints
9.5.4 Solving Method
9.5.5 Case Study
9.6 Conclusion
References
Chapter 10: GH2 Networks: Production, Supply Chain, and Storage
10.1 Definition of Green Hydrogen
10.2 Characteristics and Initiatives of Green Hydrogen
10.3 Hydrogen Supply Chain Network (HSCN)
10.4 Green Hydrogen Production Energy Sources
10.4.1 Wind Power
10.4.2 Solar Energy
10.5 Green Hydrogen Production Methods
10.5.1 Water Electrolysis
10.5.2 Photocatalysis
10.5.3 Biomass
10.6 Green Hydrogen Transportation
10.7 Green Hydrogen Safety
10.8 Green Hydrogen Technologies in Region
10.8.1 Patterns
10.8.2 Other Policies Related to GH2 in Different Continents
10.9 Standards of Green Hydrogen Refueling Places
10.10 Challenges and Outlook
10.10.1 Greenhouse Gases (GHG)
10.10.2 Continuous and Economical GOs Scheme
10.10.3 Emission Intensity Threshold
10.11 Conclusion
References
Chapter 11: Supply Chains of Green Hydrogen Based on Liquid Organic Carriers Inside China: Economic Assessment and Greenhouse ...
11.1 Introduction
11.1.1 Status Quo, Challenges, and Outlook
11.1.2 Problem Definition and Research Aim
11.2 Hydrogen Storage and Transport: State of the Art
11.2.1 Road Transportation
11.2.2 Rail Transportation
11.2.3 Pipeline Transportation
11.3 Case Study
11.3.1 China´s Main Consumption Centres: Jing-Jin-Ji, Yangtze River Delta, Pearl River Delta
11.3.2 China´s Main Renewable Power Producer Provinces: Xinjiang, Tibet, Inner Mongolia, Gansu
11.4 Methodology
11.4.1 Supply Chain Steps
11.4.1.1 Potential for Green H2 Production
11.4.1.2 Projected Demand for Green H2
11.4.1.3 Transportation
11.4.2 Green H2 Cost-Price
11.4.3 CO2-eq Emission Savings
11.5 Results and Discussion
11.5.1 Final Cost-Price of H2 Per Region
11.5.2 Total Primary Energy Supply Per Region
11.5.3 Total Final Energy Consumption Per Region
11.5.4 CO2-eq Emission Savings Per Region
11.6 Conclusion
References
Chapter 12: Green Hydrogen Research and Development Projects in the European Union
12.1 Introduction
12.2 The Hydrogen Technology Ecosystem
12.2.1 Hydrogen as a Sustainable Energy Source
12.2.2 Fuel Cells
12.2.2.1 Working Principles of Fuel Cells
12.2.2.2 Advantages of Hydrogen and Fuel Cells
12.2.3 Challenges and Considerations
12.3 Hydrogen Policy of the European Union (EU)
12.3.1 European Hydrogen Bank
12.4 Advantages of Hydrogen in Connection with Renewable Energy Sources
12.5 GH2 Projects in the World
12.6 Hydrogen Roadmap in Europe
12.6.1 Horizon 2019
12.6.2 Horizon 2020
12.6.2.1 REFHYNE Project
12.6.2.2 GENCOMM Project
12.6.3 Horizon Europe Projects
12.6.4 Clean Hydrogen Joint Undertaking
12.6.4.1 Green Hysland Project
12.6.4.2 H2FUTURE Project
12.6.4.3 Hybalance Project
12.7 Implications and Conclusion
References
Chapter 13: Hydrogen-Combined Smart Electrical Power Systems: An Overview of United States Projects
13.1 Introduction
13.2 National Renewable Energy Laboratory (NREL) Hydrogen Technologies
13.2.1 Description of NREL´s Involvement in Advancing Green Hydrogen Technologies
13.2.2 Overview of Projects Related to Hydrogen Production, Storage, Fuel Cells, and Integrated Energy Systems in NREL
13.3 Western Green Hydrogen Initiative
13.3.1 Explanation of the Initiative´s Goals to Develop a Regional GH2 Economy in the Western United States
13.3.2 Collaborative Efforts Among Stakeholders to Accelerate GH2 Deployment
13.3.3 Gulf Coast Clean Energy Application Center
13.4 Hawaii Hydrogen Initiative
13.5 California Hydrogen Infrastructure Project
13.5.1 California Energy Commission´s Support for Hydrogen Infrastructure Development
13.6 Los Angeles Green Hydrogen Power-to-Gas Project
13.7 Advanced Clean Energy Storage (ACES)
13.8 Utah Advanced Clean Energy Storage (UACES)
13.9 New York Offshore Wind-to-Hydrogen Project
13.10 Texas Gulf Coast Hydrogen Hub
13.11 Conclusion
References
Chapter 14: An Overview of the Pilot Hydrogen Projects
14.1 Status Quo, Challenges, and Outlook
14.2 Introduction
14.3 Definitions and Classification of Hydrogen
14.3.1 Production
14.3.2 Purification
14.3.3 Compression
14.4 Background and Motivations
14.5 Conclusion
References
Index
Recommend Papers

Green Hydrogen in Power Systems (Green Energy and Technology) [2024 ed.]
 3031524284, 9783031524288

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

Green Energy and Technology

Vahid Vahidinasab Behnam Mohammadi-Ivatloo Jeng Shiun Lim   Editors

Green Hydrogen in Power Systems

Green Energy and Technology

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

Vahid Vahidinasab • Behnam Mohammadi-Ivatloo Jeng Shiun Lim Editors

Green Hydrogen in Power Systems

Editors Vahid Vahidinasab Department of Engineering, School of Science and Technology Nottingham Trent University Nottingham, UK

Behnam Mohammadi-Ivatloo Electrical Engineering LUT University Lappeenranta, Finland

Jeng Shiun Lim Faculty of Chemical and Energy Engineering Universiti Teknologi Malaysia Johor, Malaysia

ISSN 1865-3529 ISSN 1865-3537 (electronic) Green Energy and Technology ISBN 978-3-031-52428-8 ISBN 978-3-031-52429-5 (eBook) https://doi.org/10.1007/978-3-031-52429-5 © The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 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 Paper in this product is recyclable.

Preface

Green Hydrogen (GH2) as a new solution for emission reduction activities worldwide attracted a lot of attention recently. Considering the significant role of hydrogen and particularly GH2 in decarbonization and very limited green hydrogen production and usage in the energy sector, there is a need for rapid scale-up in GH2 integration from where we are at present. GH2 can support deep decarbonization, especially in energy sectors that are challenging to decarbonize. GH2 is able to give green and flexible energy in a range of energy sectors from power systems to heat and transport. There are some strategic challenges in developing GH2 to make it available at scale, including production costs, technological uncertainties, policy and regulation, infrastructure development, and demand that call for more studies. This book brought together experts from the different disciplines related to hydrogen energy as we strongly believe that there is a wealth of knowledge available in each discipline that is not widely known in the other disciplines and could be usefully employed to face the challenges we are facing at this time and provide a comprehensive and in-depth reference. Chapter 1, entitled “An Overview of Energy and Exergy Analysis for Green Hydrogen Power Systems,” provides a comprehensive overview of energy and exergy analysis methods applied to green hydrogen power systems. It explains the fundamental principles of energy and exergy analyses, emphasizing their role in quantifying energy flows and assessing system thermodynamic efficiency. Key components of green hydrogen power systems, including renewable energy sources, electrolyzers, hydrogen storage, and fuel cells, are discussed in the context of energy and exergy analysis. The chapter also examines various efficiency metrics, technoeconomic analysis, and opportunities for enhancing system performance in the pursuit of sustainable low-carbon energy systems. Chapter 2, entitled “Hydrogen-Incorporated Sector Coupled Smart Grids: A Systematic Review and Future Concepts,” introduces that the surge in solar system adoption, driven by renewable energy awareness and reduced technology costs, faces challenges like intermittency and limited storage. Incorporating hydrogen into smart grids can mitigate these challenges by storing excess solar energy as v

vi

Preface

hydrogen for later use. This chapter provides a thorough review of green hydrogenintegrated smart grids, covering their significance, fundamentals, sector coupling, existing projects, technological advancements, economic and environmental aspects, and future prospects. This analysis enhances our understanding of hydrogen’s role in advancing sustainable energy systems. Chapter 3, entitled “Techno-economic Analysis for Centralized GH2 Power Systems,” delves into a comprehensive economic analysis of a centralized GH2 power system. Emphasizing the escalating demand for clean and renewable energy solutions, the study outlines the constraints of conventional power setups and their environmental implications. Through the integration of hydrogen storage, renewable sources gain potency. The exploration extends to optimum coalition strategies and peer-to-peer energy trading, fostering cost-efficient and eco-friendly energy transition. Chapter 4, entitled “Techno-economic Analysis for Decentralized GH2 Power Systems Summary,” conducts a comprehensive techno-economic analysis of decentralized GH2 power systems, integrated with transactive energy and peer-topeer (P2P) energy trading. The study emphasizes the significance of integrating renewable energy sources and decentralized systems in achieving a sustainable low-carbon future. By optimizing energy generation, consumption, and distribution, these systems offer resilience, cost reductions, and enhanced grid stability. The findings underscore the potential of P2P trading, hydrogen storage, and efficient resource utilization for advancing toward a sustainable energy landscape. Chapter 5, entitled “Hydrogenation from Renewable Energy Sources for Developing a Carbon-Free Society: Methods, Real Cases, and Standards,” assesses status quo, challenges, and outlook of hydrogen production. Then, a comprehensive review on hydrogenation methods from renewable energy sources, requirements, advantages, and limitations of each process is provided. Applications of hydrogen, hydrogen storage technologies, transportation issues, and standards associated with green hydrogen are discussed. Finally, conclusion will be presented. Chapter 6, entitled “The Role of Green Hydrogen in Achieving Low and Net-Zero Carbon Emissions: Climate Change and Global Warming,” delves into the definition and significance of green hydrogen in achieving climate-neutral economies. It outlines challenges in attaining low or net-zero emissions through GH2, emphasizing economic viability. The impact of taxes and penalties on technologies and Carbon Capture Utilization and Storage (CCUS) integration is explored. The roadmap for carbon neutrality, enhanced GH2 production, availability, and pricing are discussed. The conversion from various H2 types to GH2 and related concerns are addressed. Chapter 7, entitled “Bioreactor Design Selection for Biohydrogen Production Using Immobilized Cell Culture System,” discusses the design selection for bioreactors using immobilized cell culture systems in the fermentative biohydrogen production. This chapter discusses the advantages of immobilized cell culture over free cell cultures. The advantages include the repeatability, space efficiency, and reduction of the lag phase. Different immobilization methods are explained, including entrapment, adsorption, encapsulation, and containment within synthetic

Preface

vii

polymers. The chapter also covers various types of bioreactors suitable for immobilized culture, including continuous stirred tank reactors (CSTR), up-flow anaerobic sludge bioreactors (UASB), fluidized bed reactors (FBR), and fixed/ packed bed reactors (PBR). The chapter emphasizes the importance of optimal conditions for immobilized microbial activity and provides examples of different bioreactors used with various substrates and its biohydrogen production performance. Chapter 8, entitled “Biomass-Based Polygeneration Systems with Hydrogen Production: A Concise Review and Case Study,” discusses the importance of biomass-based polygeneration systems in producing clean and safe hydrogen as an energy carrier. The study reviews previous research and introduces a new multigeneration system with hydrogen production, which is thermodynamically evaluated. Overall, the benefits of biomass-based polygeneration systems, which can produce multiple products and minimize wastes, along with their potential for green hydrogen production are highlighted in this chapter. Chapter 9, entitled “Integration of Solar PV and GH2 in the Future Power Systems,” explores the integration of GH2 with solar energy in future power systems, emphasizing its decarbonization and energy storage potential, and it addresses challenges, reliability indicators, planning, and case studies using optimization techniques. The chapter highlights how GH2 can enhance energy system stability, reduce costs, and contribute to a more sustainable and reliable energy future. Chapter 10, entitled “GH2 Networks: Production, Supply Chain and Storage,” examines the ways to produce GH2 with today’s standards and technologies in the GH2 network. It also discusses the principles, development rate, significant research points and technologies and challenges of GH2 production around the world and discusses the general state of global hydrogen energy. Chapter 11, entitled “Supply Chains of Green Hydrogen Based on Liquid Organic Carriers Inside China: Economic Assessment and Greenhouse Gases Footprint,” investigates the potential of electrolysis to provide hydrogen, sourced from remote Chinese provinces. It analyzes the economic and environmental impacts of transporting green hydrogen to major industrial centers. Liquid Organic Hydrogen Carriers facilitate storage and transportation. Data is sourced from literature, expert interviews, and databases. Results suggest feasible green hydrogen contribution toward China’s carbon neutrality targets. Chapter 12, entitled “Green Hydrogen Research and Development Projects in the European Union,” examines EU’s strategic research projects, particularly under Horizon 2020 and Horizon Europe, aimed at advancing green hydrogen technology. This promising technology involves using renewable sources like wind and solar to electrolyze water, producing clean hydrogen. By scrutinizing projects from 2010 to 2023, the analysis assesses progress, challenges, and implications for EU’s ambitious carbon neutrality goal by 2050.

viii

Preface

Chapter 13, entitled “Hydrogen-Combined Smart Electrical Power Systems: An Overview of United States Projects,” underscores hydrogen’s integration into intelligent grids, yielding energy storage, sector integration, and decentralized generation benefits. The endeavors in California, Hawaii, and other areas in the USA demonstrate hydrogen’s capacity for grid flexibility, renewable assimilation, and emissions reduction, thereby cultivating a robust and sustainable energy landscape. Chapter 14, entitled “An Overview of the Pilot Hydrogen Projects,” reviews the developments and prospects of hydrogenated technologies in power systems including their application in power systems for hydrogen production. The increase in demand for electricity consumption besides the importance of concentrating on environmental pollutants and greenhouse gas emissions reduction has caused traditional power systems to be pushed toward the use of clean energy in electrical energy production and make a decision to deal with climate changes. As a multidisciplinary reference, the book is appropriate for both specific and general audiences, encompassing researchers and industry stakeholders who have been involved in the integration of hydrogen into power and energy systems, as well as researchers and developers from various fields, including engineering, energy, computer sciences, data, economics, and operation research. Advanced undergraduate or graduate modules and courses on energy systems would benefit from the material in this book. The book provides an adequate mixture of technology and engineering background and modeling approaches that makes it a suitable reference for students as well as researchers and engineers in academia and industry who are active in the field. In conclusion, we wish to express our appreciation for all of the contributions from the authors who have contributed to the book as well as for the insightful observations and helpful comments of all of the reviewers. In the hopes that this book will be helpful to researchers, graduate students, and practitioners in this field, the editors and authors have dedicated their time and enthusiasm to creating it. Nottingham Trent University, Nottingham, UK LUT University, Lappeenranta, Finland Universiti Teknologi Malaysia, Johor Bahru, Malaysia

Vahid Vahidinasab Behnam Mohammadi-Ivatloo Jeng Shiun Lim

Contents

1

2

3

4

An Overview of Energy and Exergy Analysis for Green Hydrogen Power Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Mohammad Mohsen Hayati, Hassan Majidi-Gharehnaz, Hossein Biabani, Ali Aminlou, and Mehdi Abapour Hydrogen-Incorporated Sector-Coupled Smart Grids: A Systematic Review and Future Concepts . . . . . . . . . . . . . . . . . . . Mohammad Mohsen Hayati, Ashkan Safari, Morteza Nazari-Heris, and Arman Oshnoei Techno-Economic Analysis for Centralized GH2 Power Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Mohammad Mohsen Hayati, Behzad Motallebi Azar, Ali Aminlou, Mehdi Abapour, and Kazem Zare Techno-Economic Analysis for Decentralized GH2 Power Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Ali Aminlou, Mohammad Mohsen Hayati, Hassan Majidi-Garehnaz, Hossein Biabani, Kazem Zare, and Mehdi Abapour

1

25

59

85

5

Hydrogenation from Renewable Energy Sources for Developing a Carbon-Free Society: Methods, Real Cases, and Standards . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105 Mehdi Talaie, Farkhondeh Jabari, and Asghar Akbari Foroud

6

The Role of Green Hydrogen in Achieving Low and Net-Zero Carbon Emissions: Climate Change and Global Warming . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141 Mohammad Shaterabadi, Saeid Sadeghi, and Mehdi Ahmadi Jirdehi

ix

x

Contents

7

Bioreactor Design Selection for Biohydrogen Production Using Immobilized Cell Culture System . . . . . . . . . . . . . . . . . . . . . 155 Nur Kamilah Abd Jalil, Umi Aisah Asli, Haslenda Hashim, Mimi Haryani Hassim, Norafneza Norazahar, and Aziatulniza Sadikin

8

Biomass-Based Polygeneration Systems with Hydrogen Production: A Concise Review and Case Study . . . . . . . . . . . . . . . . 173 Zahra Hajimohammadi Tabriz, Mousa Mohammadpourfard, Gülden Gökçen Akkurt, and Başar Çağlar

9

Integration of Solar PV and GH2 in the Future Power Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203 Hassan Majidi-Gharehnaz, Hossein Biabani, Ali Aminlou, Mohammad Mohsen Hayati, and Mehdi Abapour

10

GH2 Networks: Production, Supply Chain, and Storage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 225 Mahsa Sedaghat, Amir Amini, and Adel Akbarimajd

11

Supply Chains of Green Hydrogen Based on Liquid Organic Carriers Inside China: Economic Assessment and Greenhouse Gases Footprint . . . . . . . . . . . . . . . . . 245 João Godinho, João Graça Gomes, Juan Jiang, Ana Sousa, Ana Gomes, Bruno Henrique Santos, Henrique A. Matos, José Granjo, Pedro Frade, Shuyang Wang, Xu Zhang, Xinyi Li, and Yu Lin

12

Green Hydrogen Research and Development Projects in the European Union . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 301 Hossein Biabani, Ali Aminlou, Mohammad Mohsen Hayati, Hassan Majidi-Gharehnaz, and Mehdi Abapour

13

Hydrogen-Combined Smart Electrical Power Systems: An Overview of United States Projects . . . . . . . . . . . . . . . . . . . . . . 321 Ashkan Safari, Mohammad Mohsen Hayati, and Morteza Nazari-Heris

14

An Overview of the Pilot Hydrogen Projects . . . . . . . . . . . . . . . . . . 341 Maryam Shahbazitabar and Hamdi Abdi

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 353

Chapter 1

An Overview of Energy and Exergy Analysis for Green Hydrogen Power Systems Mohammad Mohsen Hayati, Hassan Majidi-Gharehnaz, Hossein Biabani, Ali Aminlou, and Mehdi Abapour

1.1 1.1.1

Introduction Green Hydrogen as a Potential Source of Clean Energy

Green hydrogen (GH2) is a highly efficient and desirable energy carrier that has the potential to address present and future energy demands while circumventing the limitations of traditional energy sources [1]. Microgrids (MGs) can play a crucial role in the integration of green hydrogen systems into the power system [2, 3]. MGs play a significant role in utilizing energy storage systems (ESSs) and distributed energy resources (DER) to fulfill the energy requirements of both manageable and unmanageable loads [4, 5]. As a fuel that is not metallic, hydrogen is free of carbon, safe to use, and boasts a greater specific energy (by mass) than gasoline. Hydrogenbased energy systems must take into account four key areas: usage of hydrogen, production, storage, and safety [6, 7]. The main sources of hydrogen production in the world are mainly crude oil, natural gas industries, electrolysis processes, and coal, of which the natural gas sector has a volume of 49% and is least related to the electrolysis sector at 4% [8]. The two sectors primarily responsible for the highest levels of hydrogen production involve the oxidation of fossil fuels and the modification of alcohol and hydrocarbons. However, these methods pose significant challenges due to their carbon emissions, associated environmental problems, and extensive energy consumption [9]. Also, a large part of the demand for the use of hydrogen is related to the production of chemical derivatives and oil refineries [10]. On the other hand, to achieve environmentally friendly technologies, it is

M. M. Hayati (✉) · H. Majidi-Gharehnaz · H. Biabani · A. Aminlou · M. Abapour Faculty of Electrical and Computer Engineering, Energy Systems Research Institute (ESRI), Smart Energy Systems Lab, University of Tabriz, Tabriz, Iran e-mail: [email protected]; [email protected]; [email protected] © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 V. Vahidinasab et al. (eds.), Green Hydrogen in Power Systems, Green Energy and Technology, https://doi.org/10.1007/978-3-031-52429-5_1

1

2

M. M. Hayati et al.

Fuel

Chemical

Hydrogen

Fuel Cells, Energy

Petroleum Refining,

Clean Energy Carrier

Storage, Turbines, or

Methanol Production,

Engines

Ammonia Production

Renewable Sources wind, solar, biomass, geothermal, hydro

Fig. 1.1 Hydrogen – a flexible, reliable, and environmentally friendly energy carrier in future energy networks

possible to use the combination of water electrolysis technologies, which are being developed more recently, with renewable energy sources. It should be noted that recently, in detailed research in North America, the target challenges related to GH2 production, storage capacity and its effects in the transportation sector, the state of available resources, and government regulations and policies in the field of green hydrogen have been presented [11]. In the field of hydrogen production and the development of GH2 technology in modern power systems, extensive research and scientific investigations have been carried out, mainly resulting in the creation of economical methods and approaches. This issue has paved the way for creating a stable and clean energy source. Figure 1.1 illustrates hydrogen as a flexible, reliable, and environmentally friendly energy carrier in future energy networks [12]. To achieve the decarbonization of energy industries and bring about crucial changes in meeting the requirements of modern energy systems, renewable energy sources such as wind and solar energy are presently utilized in the production of hydrogen [13, 14]. This plays a vital part in meeting the rising demand for renewable, clean energies. Due to the lightweight of hydrogen gas, the process of transporting and storing it is relatively difficult, although recently, with the emergence of hydrogen carrier technologies that are able to transport and transport solid and liquid hydrogen, the hydrogen storage process has become somewhat easier [15]. Such developments in hydrogen production provide the basis for a more sustainable future in the field of clean energy and reducing dependence on fossil fuels [16]. In 2022, electricity demand increased in India and the United States, while the COVID-19 restrictions affected China’s electricity demand growth [17]. China’s zero-covid policy has severely impacted the country’s economic activity, and there is still a degree of uncertainty regarding the rate of growth in the country’s electricity demand [18]. The International Energy Agency (IEA) estimates that this figure will reach 2.6% in 2022, which is significantly lower than the pre-COVID pandemic average of more than 5% in the period between 2015 and 2019. In 2022, India is projected to experience an 8.4% surge in electricity demand, primarily driven by a robust economic rebound following the COVID-19 pandemic and unusually high summer temperatures. Similarly, the United States witnessed a notable 2.6% rise in electricity demand during the same year, primarily fueled by economic activity and increased residential electricity consumption for heating and cooling purposes due to hotter summers and colder winters than usual [10, 19].

1

An Overview of Energy and Exergy Analysis for Green Hydrogen Power Systems

3

Fig. 1.2 A brief comparison of different types of hydrogen technology and CO2 emissions

In the years 2022–2025, global electricity production using natural gas and coal burning is expected to remain substantially constant [20]. During 2022, the European Union saw an increase in the production of electricity from burning gas. While the European Union anticipates a decline in electricity generation from natural gas, the substantial growth of this energy source in the Middle East is expected to partially offset the decrease. Likewise, as coal-fired power generation diminishes in Europe and America, there is a corresponding increase in Asia and the Pacific region [19, 20]. According to reports published by the IEA, between 2017 and 2040, the increase in global energy demand is estimated to be about 25%. It is also estimated that in the next 3 years, the share of the developing countries and the emerging economies of the Middle East, such as China and India, and the countries of Southeast Asia in energy consumption and demand will continue to be high, reaching more than 70%. It can be said that China and India are the centers of gravity of energy demand in the world, which will drive energy consumption in the world from 25% in 2015 to about 33% of the total by 2025. In addition, by 2025, sources that emit less greenhouse gases will cover almost all of the growth in global electricity demand. Globally, the higher costs of electricity production were mainly due to the increase in the price of energy carriers. In other words, affordability is still a challenge for emerging and developing economies [21, 22]. As the proportion of renewable resources incorporated into power systems continues to rise worldwide, the significance of energy storage systems is expected to grow. These systems will contribute by offering frequency control capabilities, operational storage, and facilitating wholesale arbitrage. Furthermore, their implementation will lead to a decrease in network integration costs [23]. The deployment of energy storage systems is increasing [24]. The United States, Europe, and China are leading the way in the latest annual capacity additions in this area [25]. However, according to 2022 estimates, emerging and developing economies are catching up to these leading countries. In 2022, it was approximated that the global emissions from electricity generation reached a peak limit of approximately 13.2 gigatons of carbon dioxide. This represents a growth of around 1.3% compared to the emissions recorded in 2021 [26, 27]. The record emissions in 2022 were mainly due to the growth of electricity generation using fossil fuels in Asia and the Pacific. Europe and Eurasia also contributed to this increase [28, 29]. Figure 1.2 illustrates a comparison of different types of hydrogen technology and CO2 emissions according to Ref. [30] (Fig. 1.3).

M. M. Hayati et al.

4

Fossil Fuel with Solar Photo Voltaic CHP Carbon Capture Unit Statuns

Geothermal

Electrical Network

Gas Infrastructure/Gas Network

Other Renewable Energies

Concentrated Solar Power Hydrogen/ Electrical Vehicles Heating

Wind Turbine

Battery or Conventional Storage

Nuclear Power Plant

Hydrogen Generation Network Power Generation Water Waste Home

Transportation Clean H2 GH2 Storage Distribution

Fuel Cells

Office N2

NH3 Metal Refining

Other End Use

Petrochemical Process

Co2 Synthetic Fuels

Biomass

Fig. 1.3 Hydrogen generation, utilization, storage and environmental impacts in future energy systems

1.2

Hydrogen Economy

Energy prices have reached their highest levels since 2008, affecting all energyconsuming sectors. Subsequently, this price increase has had severe inflationary effects on all energy-consuming sectors. In this regard, the IEA has estimated that the effect of the increase in the price of fossil fuels has contributed 90% to the increase in the price of electricity in 2022, and this year the price of fossil fuels has increased by more than 50% [31].

1

An Overview of Energy and Exergy Analysis for Green Hydrogen Power Systems

5

The concept of the hydrogen economy refers to an energy infrastructure that relies on hydrogen (H2) as a substitute for conventional fossil fuels in order to fulfill energy needs [32, 33]. By transitioning to hydrogen as a primary energy carrier, we can reduce our dependence on fossil fuels and mitigate the associated environmental challenges [34, 35]. This paradigm shift involves the production, distribution, and utilization of hydrogen across various sectors, including transportation, power generation, and industrial applications [36]. Implementing a robust hydrogen economy requires the establishment of a comprehensive infrastructure that encompasses hydrogen production methods, storage and transportation systems, as well as efficient conversion technologies such as fuel cells [37, 38]. Embracing the hydrogen economy holds great potential in promoting sustainability, reducing carbon emissions, and fostering a cleaner and more diversified energy landscape [39, 40].

1.3

Economic and Environmental Effects of GH2 Production

The production of GH2 can have several economic and environmental effects on power systems. These effects depend on various factors, including the availability of renewable energy sources, infrastructure development, and the overall demand for hydrogen as a fuel. As technology continues to advance and renewable energy costs decrease, green hydrogen may become a more viable and sustainable alternative to traditional fossil fuels [41]. In the following subsections, some important economic and environmental implications are described.

1.3.1

Economic Issues

1.3.1.1

Expensive Production

The cost of producing GH2 is currently higher compared to conventional methods due to the expensive renewable energy sources required, such as wind and solar [42].

1.3.1.2

Investment Needed

To produce GH2, there is a need for substantial investment in renewable energy infrastructure and production facilities [43]. This may pose a challenge, especially in areas where renewable energy is not yet well-established.

6

M. M. Hayati et al.

1.3.1.3

Infrastructure Limitations

The distribution and storage infrastructure for hydrogen are not as developed as traditional fuels, resulting in added expenses and logistical difficulties for the transportation and storage of GH2 [44, 45].

1.3.2

Environmental Issues

1.3.2.1

Carbon Emissions

Although GH2 production does not produce carbon emissions during the production process [46], renewable energy generation may still emit carbon in regions where fossil fuels are the primary source of energy [47, 48].

1.3.2.2

Land Use

The production of GH2 requires significant land use for renewable energy infrastructure installation, such as wind turbines and solar panels, which can have adverse environmental effects on local ecosystems and habitats.

1.3.2.3

Water Use

Large amounts of water are necessary for GH2 production, particularly in the electrolysis process, which can strain water resources in areas experiencing water scarcity [49, 50].

1.4

GH2 Production Methods and Explanation of How Electrolysis Works

There are various methods available for the production of green hydrogen, with electrolysis being the most prevalent and widely adopted approach [51]. It has emerged as the preferred method for green hydrogen production due to its compatibility with renewable energy sources and its ability to generate hydrogen with high purity. When an electric current is passed through the water, it causes the water molecules to break apart into their constituent atoms: two hydrogen atoms and one oxygen atom. The oxygen atoms then react with the electrode to form oxygen gas, which bubbles out of the solution [52]. Meanwhile, the hydrogen atoms are attracted to the other electrode and combine to form hydrogen gas. The resulting hydrogen gas can be collected and used as a fuel for various applications, including powering

1

An Overview of Energy and Exergy Analysis for Green Hydrogen Power Systems

7

vehicles, generating electricity, and heating homes. The process of electrolysis can be powered using renewable energy sources, such as solar or wind power, to produce green hydrogen [53]. Other methods of GH2 production are biomass gasification, photoelectrochemical (PEC) water splitting, and thermolysis [54, 55]. In improving the flexibility of energy systems as well as paying attention to the issue of energy security and sustainability, hydrogen acts as an efficient secondary energy source with integrated smart networks [56, 57]. In general, a transition in the future power and energy systems is unavoidable because of issues such as the depletion of fossil fuels and environmental damage [58, 59].

1.4.1

Electrolysis

Electrolysis is a procedure that employs an electrical current for the purpose of separating water (H2O) into its constituent gases, hydrogen (H2), and oxygen (O2) [60, 61]. It is widely regarded as the preferred method of hydrogen production due to its reliance on electricity rather than fossil fuels [62]. Additionally, electrolysis can operate effectively across a broad range of electrical energy capacities, making it adaptable to leverage surplus electricity available during nighttime hours. Central to the electrolysis process is the electrolyzer, which comprises multiple cells, each containing a positive and negative electrode [63, 64]. The electrodes are immersed in a conductive aqueous solution, which is created by introducing hydrogen or hydroxyl ions, typically achieved through the addition of alkaline potassium hydroxide [65]. Electrolysis holds significant potential for enabling large-scale hydrogen production, fostering the transition to a low-carbon economy, and supporting the integration of renewable energy sources into the energy system [66, 67]. The anode, commonly made from a combination of nickel and copper, is coated with metal oxides such as manganese, tungsten, and ruthenium. These metal oxides on the anode promote effective binding of atomic oxygen to oxygen pairs present on the electrode’s surface, ensuring efficient operation. This catalytic process promotes the desired reactions and enhances the overall efficiency of electrolysis. The negative electrode, commonly known as the cathode, is typically constructed using nickel material that is coated with small quantities of platinum, serving as a catalyst. This catalyst plays a crucial role in facilitating the swift combination of atomic hydrogen, forming hydrogen pairs on the electrode surface and effectively enhancing the rate of hydrogen production [68]. The presence of this catalyst prevents the accumulation of atomic hydrogen on the electrode, which could impede the flow of electric current if left unaddressed. In optimal circumstances, the electrolysis process necessitates 39.4 kWh of energy and 8.9 L of water to generate 1 kg of hydrogen [69]. This aspect highlights the notable calorific value of hydrogen, encompassing the entirety of energy needed to dissociate water under normal conditions. On certain occasions, the lower heating value of hydrogen is employed for efficiency comparisons, which amounts to 33.3 kWh/kg of hydrogen. To assess the effectiveness of the system, the heating value is divided by the actual input energy in kilowatt-hours per kilogram (kWh/kg), as specified in the provided reference [70].

8

1.4.2

M. M. Hayati et al.

Fuel Cells

As the demand for electricity continues to rise, renewable energy sources have garnered widespread support across various sectors. However, a challenge associated with relying on these renewable sources is their output being disconnected from the actual demand [71, 72]. The utilization of electricity storage offers a valuable opportunity to effectively manage and balance the supply and demand of electricity [73, 74]. Fuel cells have emerged as a rapidly advancing technology that is increasingly dominating energy markets, providing an efficient solution for storing electricity in the form of hydrogen [75, 76]. This enables the storage of electrical energy for later use, ensuring a reliable and flexible energy system [77]. Fuel cells operate by combining fuel and oxidant gases at the anode and cathode, respectively. This process necessitates a well-designed physical structure that facilitates the controlled flow of gases to both sides of the electrolyte [78]. The fuel cell’s key characteristic lies in its unique electrolyte, which varies among different types of fuel cells and determines the specific ions it conducts [79, 80]. The ability to convert approximately 60% of the chemical energy stored in hydrogen into electricity is a notable advantage of hydrogen fuel cells, contributing to their high efficiency [81–83]. With the emergence and advancement of fuel cells and electrolytic hydrogen technology, the interaction between hydrogen and electricity is growing steadily in this emerging field [84]. The expansion of renewable energy in power systems, as well as its variability and unpredictability, poses challenges to the performance of energy systems [85]. On the other hand, fuel cells and the process of water electrolysis can improve operational flexibility by linking energy and hydrogen systems [86]. Fuel cells act as a source of energy in the conditions of extreme events, with real-time reactions to restore the load of power systems [87, 88]. Therefore, hydrogen systems have the ability to quickly support smart grids due to their long-term storage properties, and this greatly contributes to energy security and stability [89].

1.5

Energy Crisis

The energy crisis has escalated parallel to the advancement and growth of societies. However, the inherent characteristics of renewable energies, such as their intermittent and unpredictable nature, pose significant challenges for these emerging technologies [48, 90, 91]. This dilemma becomes particularly apparent in wind and solar energy, where issues such as power intermittency and variability of sunlight and wind speed hinder economic efficiency for energy production companies, leading to substantial energy losses. Consequently, this presents a clear contradiction in the pursuit of developing clean and sustainable renewable energy sources [92, 93]. To address these challenges, various strategies are being implemented to enhance the integration and reliability of renewable energy systems [94]. One approach involves the development of advanced energy storage technologies, allowing excess energy

1

An Overview of Energy and Exergy Analysis for Green Hydrogen Power Systems

9

generated during peak periods to be stored for use during low-demand periods [95, 96]. Also, sophisticated forecasting and monitoring systems are being employed to improve the accuracy of predicting renewable energy availability, enabling better management and optimization of power generation and consumption [97, 98]. Furthermore, efforts are underway to diversify the renewable energy mix by exploring complementary sources, such as combining wind and solar energy with other forms of renewable generation, such as hydropower or geothermal energy [99]. This diversification aims to create a more balanced and reliable renewable energy portfolio that can mitigate the inherent challenges associated with intermittency and uncertainty. Addressing the contradiction between renewable energy’s clean attributes and the challenges it poses requires a comprehensive approach involving technological advancements, supportive policies, and investment in research and development. By overcoming these hurdles, the potential of clean renewable energies can be fully harnessed, contributing to a sustainable and resilient energy future [100].

1.6

Integration of GH2 Systems with Renewable Energy Sources and Energy Hub

The combination of green hydrogen systems with renewable energy presents a broad outlook. As previously mentioned, wind and solar power are clean and have significant storage capacity [101]. By integrating hydrogen and the power network with wind and solar energy networks, the global issues of environmental pollution and greenhouse gas emissions can be effectively and efficiently reduced [102, 103]. This multiple energy supply system, referred to as an energy hub (EH), offers systematic flexibility for energy and load management [57, 104]. Integrating different types of energy into multiple energy infrastructures, including electricity, natural gas, heat, hydrogen, and other renewable sources in an integrated way, provides cost-effective demand [105]. EHs offer a great opportunity for energy system operators to create a more efficient and higher performing system. The investigation of the economic viability of EHs in the face of uncertain conditions, further compounded by the existence of diverse energy sources, is a focal point of interest within this field. The overall performance of the power system relies on the optimal performance of each component of the EH [106]. By integrating solar heat into the wind–solar–hydrogen multiple energy supply system, the overall energy efficiency of power systems is enhanced, resulting in an increased utilization of renewable energy at a macro level [107]. Notably, green hydrogen holds immense promise as a clean energy carrier that can be stored over extended periods without experiencing degradation [108]. Its deployment within an EH framework enables the effective storage and utilization of renewable energy resources, contributing to a sustainable and resilient energy ecosystem. By leveraging the capabilities of an EH and harnessing the benefits of green hydrogen, we can propel the transition toward a greener and more secure energy future.

10

1.7

M. M. Hayati et al.

Future of Energy Supply Systems and Related Works

Given the dynamic nature of future energy supply systems, it is vital to embrace the advancement of multi-energy complementary distributed energy systems and harness the full potential of the synergistic advantages provided by a wide range of energy sources [109]. These approaches serve as vital strategies to tackle the prevailing challenges associated with high costs and low efficiency in traditional distributed systems [110]. At present, extensive research efforts are focused on distributed energy systems that capitalize on the complementary time scales of different energy sources. These systems have made significant strides and have even achieved successful large-scale implementations in certain domains. By adopting a multi-energy approach, these distributed systems intelligently integrate and optimize the utilization of various energy sources. The time sequence of solar and wind energy sources exhibits a certain level of complementarity [111]. Researchers have extensively investigated the integration of wind power (WP), concentrated solar power (CSP), and photovoltaic power generation (PV) to create a synergistic system for solar and wind energy generation [112, 113]. Given the inherent variability of wind and solar energy sources [114], an effective approach to mitigate their fluctuations and ensure a stable and reliable power output is to integrate solar thermal power generation with wind power (WP) and photovoltaic (PV) power generation. By combining these technologies, the stability, continuity, and dispatchability of solar thermal power can be leveraged to counteract the fluctuations observed in wind and PV generation. This integration enables the production of high-quality output power [115]. In situations where there is surplus energy, it is more advantageous to generate clean fuels such as hydrogen [116], which serves as an appropriate energy carrier and effective storage medium [117, 118]. The energy storage technique, which involves electrolyzing water using wind energy or PV power to generate hydrogen and subsequently utilizing hydrogen fuel cells for electricity generation, has been extensively developed [119]. This approach has been proven to offer several benefits, including (1) decreased consumption of fossil fuels and the release of pollutants [120]; and (2) enhanced energy utilization and reduced waste of wind and solar resources [121]. Furthermore, Sezer et al. [122] introduce a multi-energy system that integrates solar, hydro, and wind energy storage. The study evaluates the overall energy efficiency and fuel consumption efficiency of the system, yielding values of 3.61% and 8.47%, respectively. The utilization of wind energy for hydrogen and electricity production dates back to 1981 when Denmark pioneered this conversion [123]. Subsequently, in 1983, solar energy was harnessed for similar purposes at the Florida Solar Energy Center [124]. A significant milestone in renewable energy storage occurred in 1991 with the construction of the first gas-fired power plant that utilized hydrogen as a renewable energy storage medium [125]. Established in California in 1995, the initial facility comprising a photovoltaic (PV) system and an electrolyzer was capable of generating approximately 50–70 N m3/day of hydrogen [126]. Subsequently, numerous hybrid renewable energy–hydrogen systems have been

1

An Overview of Energy and Exergy Analysis for Green Hydrogen Power Systems

11

constructed worldwide [76]. Combined heat and power (CHP) power plants generally have a higher efficiency than when working in separate processes. In an assessment made in Ref. [127], the efficiency of power plants in mechanical or electrical production ranges is estimated from about 40% to about 85% in simultaneous production. A study was conducted [128] to examine the expenses associated with reducing CO2 emissions through the use of extensive biomass-fired cogeneration technologies combined with CO2 storage. The results revealed these plants, which utilize integrated gasification combined cycle technology, are highly efficient in terms of energy utilization and emissions [129]. However, when compared to technologies such as batteries, hydrogen systems still exhibit lower efficiency, presenting a significant barrier to the widespread adoption of hydrogen technology in practical applications [130, 131]. Recently, cogeneration in hydrogen energy systems has attracted more attention [132, 133]. Numerous studies and evaluations have examined the production of renewable energies, particularly focusing on modeling these sources with and without hydrogenation systems [134]. Additionally, investigations have been conducted on the utilization of hybrid renewable systems for diverse purposes, including electricity generation, space heating, and cooling [135]. In recent times, cogeneration within hydrogen energy systems has garnered increased attention [130, 136, 137]. Furthermore, various processes in hydrogen storage systems, such as electrolysis and fuel cells, generate heat that can be recovered and employed for various applications. Cogeneration plays a significant role in boosting the overall efficiency of specific power plants, leading to potential energy savings of around 40% [138]. The reference [134] explores different methods of hydrogen production, specifically focusing on photoelectrolysis and solar thermal hydrogen production. It provides a comprehensive assessment of fuel cell, hydrogen, and solar systems, along with their respective applications. Since the early 1980s, nearly 99 hydrogen projects have been implemented worldwide, utilizing renewable resources across a variety of applications and scales, including industrial and experimental contexts. This research specifically examines a subset of these projects that involve simultaneous production.

1.8

Energy and Exergy Analysis

Some studies have focused on analyzing the energy or exergy of multi-energy systems that utilize biogas to produce hydrogen, electricity, and heat [139, 140]. However, these studies have not considered the evaluation of the purification, compression, and storage stages required for utilizing hydrogen in the transportation sector. Other studies have explored this application, but they did not employ a multi-energy systems approach, instead, they relied on the grid to meet the electricity demands of the plant. The novelty of this study lies insights into the most effective energy and exergy analysis methods in green hydrogen power systems and highlights challenges and opportunities for future research in this area. The exergy analysis of a GH2 power system can identify the sources of exergy losses due to

12

M. M. Hayati et al.

irreversibilities, such as heat transfer, fluid flow, and chemical reactions, as well as opportunities for exergy recovery and optimization. One of the key benefits of an energy and exergy analysis of a GH2 power system is that it can identify the most significant sources of energy and exergy losses in the system, which can then be targeted for improvement. For example, the energy analysis of a GH2 power system may reveal that the hydrogen production unit consumes a significant amount of energy due to the inefficiencies of the electrolysis process, and the exergy analysis may identify the irreversibilities in the electrode polarization and heat transfer. By minimizing the energy and exergy losses, the environmental impact of the system can be reduced, which can be an important consideration for applications such as transportation and stationary power generation. Also, by identifying the sources of energy and exergy losses and opportunities for improvement, the energy and exergy efficiency of the system can be improved, which can lead to significant economic, environmental, and social benefits. Exergy encompasses various concepts such as effective energy and energy availability [141], providing a comprehensive understanding of energy that goes beyond mere quantity. By integrating the first and second laws of thermodynamics, exergy analysis evaluates energy not only in terms of its quantity but also its quality, offering deeper insights into energy degradation during its utilization. The standard measure of system efficiency is the ratio of output energy to input energy [142]. In the present study, the input energy sources of the system consist of wind power, concentrated solar power, and photovoltaic systems, while the output energy sources include both electric energy and hydrogen energy [143, 144]. By employing exergy analysis, the evaluation of system efficiency extends beyond a simple calculation of energy conversion ratios. It takes into account the quality and availability of energy, shedding light on the overall effectiveness of the energy conversion processes within the system. This comprehensive assessment allows for a more accurate understanding of energy utilization and degradation, facilitating the optimization of system performance and the identification of potential areas for improvement.

1.9

Related Works

Colakoglu and Durmayaz [145] propose a novel solar tower-based system designed to produce green hydrogen. The system consists of multiple power cycles, namely a solar-driven open Brayton cycle incorporating intercooling, regeneration, and reheat, as well as a regenerative Rankine cycle and a Kalina cycle-11. A significant portion of the electricity generated is dedicated to the electrolysis process for producing green hydrogen. Additional components of the system include thermal energy storage, a single-effect absorption refrigeration cycle, and two domestic hot water heaters. To evaluate the system’s performance, various analyses such as energy, exergy, economic, and detailed parametric analyses are conducted. The researchers employ multiobjective optimization techniques to determine the optimal performance parameters. The resulting optimum values obtained from the study

1

An Overview of Energy and Exergy Analysis for Green Hydrogen Power Systems

13

include energy and exergy efficiencies of 39.81 and 34.44%, respectively, a unit exergy product cost of 0.0798/kWh, and a total cost rate of 182.16/h. Incer-Valverde et al. [146] examine a future green hydrogen hub in Hamburg, Germany, where a large-scale power-to-liquid hydrogen system is evaluated. This system utilizes renewable electricity and employs a polymer electrolyte membrane electrolyzer to generate hydrogen. The produced hydrogen is then liquefied and stored at cryogenic temperatures under ambient pressure. The evaluation of this system involves various exergy-based methods, including exergetic, exergoeconomic, and exergoenvironmental analyses. The liquefaction process demonstrates an exergetic efficiency of 42%, while the electrolyzer achieves an efficiency of 47%. The overall exergetic efficiency of the power-to-liquid hydrogen system is calculated to be 44%. Through the analysis, the researchers identify the electrolyzer and hydrogen compressors as the components with the highest exergy destruction values and investment costs. Furthermore, the compressors and recuperators are found to have the most significant exergoenvironmental impact within the system. Minutillo et al. [147] compare two biogas-based hydrogen production plants designed as polygeneration systems. These plants generate high-pressure hydrogen, heat, and electricity to meet their own energy needs for purification, compression, and storage. One plant utilizes steam reforming, while the other employs autothermal reforming. The study finds that the steam reforming-based configuration achieves superior energy-based efficiency (59.8%) and exergy-based efficiency (59.4%) for hydrogen production. It also performs better in terms of coproducing heat and hydrogen (energy-based efficiency: 73.5%, exergy-based efficiency: 64.4%). The ATR-based layout, on the other hand, is more exothermic and suitable for larger local heat demands (energy-based efficiency: 73.9%, exergy-based efficiency: 54.8%). Nalbant Atak et al. [148] explore the development of an integrated membrane reactor and CO2 capture system for decarbonized hydrogen production. The article presents the results of energy and exergy analyses conducted on the integrated system. A one-dimensional model of the membrane reactor was created, validated, and used to assess the effects of various operating parameters. The membrane reactor model was then incorporated into a system-level model, considering a CO2 capture unit and other plant components. This allowed for a theoretical analysis of the system’s potential to generate decarbonized hydrogen. The study’s novelty lies in the application of system-level modeling based on electrochemistry and thermodynamics, enabling a detailed energy and exergy analysis. The study also calculates the rate of exergy destruction for each component of the integrated membrane reactor system. Under baseline simulation conditions, the thermal efficiency (based on lower heating value), methane conversion, hydrogen yield, and CO2 yield are determined to be 51, 67, 22, and 66%, respectively. Nasser et al. [149] evaluate a renewable energy-based hydrogen production system using solar and wind. The hybrid system is analyzed for energy, exergy, economics, and environmental aspects. It incorporates PV panels, wind turbines, and a water electrolyzer. The system achieves an overall energy efficiency of 16.42% and an exergy efficiency of 12.76%. Economic analysis considers various degradation rates and scenarios for electricity production, revealing ranges for LCOE, LCOH, and LCOCH. The payback period varies from

14

M. M. Hayati et al.

7 to 13.85 years, and the system reduces CO2 emissions by 689.4 tons over its lifetime. Nikitin et al. [150] discuss a dynamic multigeneration system that utilizes solar and wind energy to provide cooling, heating, electricity, and freshwater to a residential building. The system undergoes comprehensive analysis, considering energy, exergy, economic, and environmental aspects, across different weather conditions in Khabarovsk, Yakutsk, Yekaterinburg, and St. Petersburg over a year. It consists of key components such as flat plate collectors, wind turbines, thermal energy storage, absorption chillers, reverse osmosis systems, internal combustion engines, and hydrogen energy storage systems. The simulation employs TRNSYS, Energy Plus, and Engineering Equation Solver packages. The economic analysis reveals the best payback period under a 0.03 interest rate: 8, 21, 11.2, and 9.8 years for Khabarovsk, Yakutsk, Yekaterinburg, and St. Petersburg, respectively. ÖZdemİR and GenÇ [151] present an energy and exergy analysis of a thermochemical hydrogen production facility powered by solar energy. The study compares 3 cycles: low-temperature MgeCl, H2SO4, and UT-3 cycles. Additionally, it investigates the integration of reheat–regenerative Rankine and recompression SeCO2 Brayton power cycles to provide the necessary electricity for the MgeCl and H2SO4 cycles. The integration of the SeCO2 Brayton power cycle improves the system performance. The system achieves a maximum exergy efficiency of 27% when combining the MgeCl thermochemical cycle with the recompression SeCO2 Brayton power cycle. The energy and exergy efficiencies decrease with increasing solar radiation but increase with higher concentration ratios. The solar energy unit exhibits the highest exergy destruction. Abuşoğlu et al. [152] conducted a study to determine the most suitable model for a sewage treatment plant, focusing on exergy efficiency. The study considered five models that utilize biogas-based electricity and sewage sludge from a municipal sewage treatment plant to produce green hydrogen. These models include alkaline processes, PEM, high-temperature water electrolysis, alkaline hydrogen sulfide electrolysis, and dark fermentation biohydrogen production processes. Thermodynamic methods were applied to conduct energy and exergy analysis on these models, and the results were compared. The calculated exergetic efficiencies for the models were found to be 19.81, 20.66, 25.83, 24.86, and 60.5%, respectively. Based on the findings, it was concluded that the dark fermentation biohydrogen production process exhibited the highest exergetic efficiency among the models, followed by the high temperature steam electrolysis process. Qi and Huang [153] conducted an extensive exploration into supercritical carbon dioxide cycles applied to water-injected hydrogen gas turbines. The investigation highlighted the numerous advantages of this approach, including zero carbon emissions, low pollution, high efficiency, and affordability. Several representative combined cycles were carefully selected from a pool of more than 12 designs, and a thermodynamic and exergy energy analysis model was developed and validated using experimental models. By conducting parameter sensitivity analysis, water mixing research, and exergy analysis, the researchers were able to achieve maximum energy yield and exergy efficiency. The findings indicate that increasing the ratio of water to hydrogen results in a decrease in the energy efficiency of the combined cycle. Combustion was identified as the component with the highest exergy loss,

1

An Overview of Energy and Exergy Analysis for Green Hydrogen Power Systems

15

accounting for 23.58% of the total. Among the studied designs, the transcritical CO2 double recovery combined cycle emerged as the most favorable, boasting a combined cycle energy efficiency of 64.39% and a combined cycle exergy efficiency of 62.96%. The insights and research presented in this article provide a solid foundation for the design of future-generation gas turbines. Arslan and Yılmaz [154] conducted an assessment of biogas energy production and explored the potential of green hydrogen as an energy carrier derived from biomass. To make use of waste gases, the researchers investigated the integration of an organic Rankine cycle (ORC). The power generated by the ORC system was utilized for electrolysis of water to produce hydrogen, eliminate H2S generated during biogas production, and store excess electricity. A comprehensive analysis involving thermodynamic, thermoeconomic, and optimization aspects was conducted for the combined heat and power (CHP) system designed for this purpose. The system design and analysis were performed using Engineering Equation Solver (EES) and Aspen Plus software. The thermodynamic analysis revealed that the energy and exergy efficiency of the existing power plant were 28.69 and 25.15%, respectively. In contrast, the new integrated system demonstrated improved performance, with energy and exergy efficiencies of 41.55 and 36.42%, as well as a power capacity of 5792 kW. Yang et al. [155] introduce a novel approach to the utilization of renewable energy, specifically focusing on the synergistic integration of hydrogen liquefaction and liquid air energy storage. The study presents a comprehensive evaluation of the techno-economic performance of this energy process, considering energy, exergy, and economic factors. The primary objective is to achieve load balancing in the grid and explore the potential for commercializing a combined system comprising liquid air energy storage and hydrogen liquefaction power plants while assessing the efficiency of renewable energy sources. The investigation conducted in this study reveals promising results. The proposed process exhibits a return efficiency of 58.9%, with a specific energy consumption of 7.25 kWh/kg for liquid hydrogen production, and an overall exergy efficiency of 53.2%. Liu et al. [156] conducted a comprehensive study that evaluated a wind–solar–hydrogen multi-energy supply system, considering energy, exergy, economic, and environmental aspects. The assessment was carried out using MATLAB/Simulink software. In this system, a fuel cell was employed as a peak energy source, operating in coordination with other renewable energy sources to mitigate fluctuations in wind and photovoltaic power generation. Controlled solar thermal power generation and hydrogen production were utilized to achieve this objective. The evaluation of the system yielded notable results. The energy efficiency was determined to be 16.03%, while the exergy efficiency reached 17.94%..

1.10

Conclusion

The study assesses the efficiency of renewable energy sources and their integration into multi-energy supply systems. Additionally, the research highlights the utilization of fuel cells as peak energy sources, working in coordination with other

16

M. M. Hayati et al.

renewable energy sources to mitigate fluctuations in power generation. Also, this research contributes to the understanding of green hydrogen power systems by providing an overview of energy and exergy analysis methodologies. It sheds light on the potential for renewable energy integration, load balancing in the grid, and the commercialization of green hydrogen technologies. The insights gained from this study can inform future developments and advancements in the field, ultimately contributing to a more sustainable and efficient energy landscape. In this research, an overview of energy and exergy analysis is carried out for a green hydrogen (GH2) power system, which produces electrical power through a fuel cell system using hydrogen produced from renewable sources.

References 1. Schrotenboer, A. H., Veenstra, A. A., uit het Broek, M. A., & Ursavas, E. (2022). A green hydrogen energy system: Optimal control strategies for integrated hydrogen storage and power generation with wind energy. Renewable and Sustainable Energy Reviews, 168, 112744. 2. Mansour-Saatloo, A., et al. (2021). Multi-objective IGDT-based scheduling of low-carbon multi-energy microgrids integrated with hydrogen refueling stations and electric vehicle parking lots. Sustainable Cities and Society, 74, 103197. 3. Mansour-Saatloo, A., Mirzaei, M. A., Mohammadi-Ivatloo, B., & Zare, K. (2020). A riskaverse hybrid approach for optimal participation of power-to-hydrogen technology-based multi-energy microgrid in multi-energy markets. Sustainable Cities and Society, 63, 102421. 4. Alilou, M., Azami, H., Oshnoei, A., Mohammadi-Ivatloo, B., & Teodorescu, R. (2023). Fractional-order control techniques for renewable energy and energy-storage-integrated power systems: A review. Fractal and Fractional, 7(5), 391. 5. Nazari-Heris, M., Abapour, S., & Mohammadi-Ivatloo, B. (2017). Optimal economic dispatch of FC-CHP based heat and power micro-grids. Applied Thermal Engineering, 114, 756–769. 6. Daneshvar, M., Mohammadi-Ivatloo, B., Zare, K., & Asadi, S. (2021). Transactive energy management for optimal scheduling of interconnected microgrids with hydrogen energy storage. International Journal of Hydrogen Energy, 46(30), 16267–16278. 7. Aminlou, A., Hayati, M. M., & Zare, K. (2023). Local peer-to-peer energy trading evaluation in micro-grids with centralized approach. In 2023 8th international conference on technology and energy management (ICTEM) (pp. 1–6). 8. Avargani, V. M., Zendehboudi, S., Saady, N. M. C., & Dusseault, M. B. (2022). A comprehensive review on hydrogen production and utilization in North America: Prospects and challenges. Energy Conversion and Management, 269, 115927. 9. Miland, H., Glöckner, R., Taylor, P., Jarle Aaberg, R., & Hagen, G. (2006). Load control of a wind-hydrogen stand-alone power system. International Journal of Hydrogen Energy, 31(9), 1215–1235. 10. IEA. (2019). The future of hydrogen. IEA, Paris https://www.iea.org/reports/the-future-ofhydrogen, License: CC BY 4.02019. Accessed 21 May 2023. 11. Madadi Avargani, V., Zendehboudi, S., Cata Saady, N. M., & Dusseault, M. B. (2022). A comprehensive review on hydrogen production and utilization in North America: Prospects and challenges. Energy Conversion and Management, 269, 115927. 12. Satyapal, S. (2016). US Department of Energy hydrogen and fuel cells program. In 2017 Annu. Merit review and peer evaluation meeting. 13. Novotny, V. (2023). Blue hydrogen can be a source of green energy in the period of decarbonization. International Journal of Hydrogen Energy, 48(20), 7202–7218.

1

An Overview of Energy and Exergy Analysis for Green Hydrogen Power Systems

17

14. Wang, J., et al. (2023). Role of electrolytic hydrogen in smart city decarbonization in China. Applied Energy, 336, 120699. 15. Pleshivtseva, Y., Derevyanov, M., Pimenov, A., & Rapoport, A. (2023). Comparative analysis of global trends in low carbon hydrogen production towards the decarbonization pathway. International Journal of Hydrogen Energy. https://doi.org/10.1016/j.ijhydene.2023.04.264 16. Daneshvar, M., Mohammadi-Ivatloo, B., & Zare, K. (2023). An innovative transactive energy architecture for community microgrids in modern multi-carrier energy networks: A Chicago case study. Scientific Reports, 13(1), 1529. 17. Li, H., Jiang, T., Wu, T., Skitmore, M., & Talebian, N. (2023). Exploring the impact of the COVID-19 pandemic on residential energy consumption: A global literature review. Journal of Environmental Planning and Management, 1–22. 18. Aghahosseini, A., et al. (2023). Energy system transition pathways to meet the global electricity demand for ambitious climate targets and cost competitiveness. Applied Energy, 331, 120401. 19. IEA. (2022). World Energy Outlook 2022. IEA, Paris. https://www.iea.org/reports/worldenergy-outlook-2022, License: CC BY 4.0 (report); CC BY NC SA 4.0 (Annex A)2022. Accessed 21 May 2023. 20. Cronshaw, I. (2015). World energy outlook 2014 projections to 2040: Natural gas and coal trade, and the role of China. Australian Journal of Agricultural and Resource Economics, 59(4), 571–585. 21. Conti, J., Holtberg, P., Diefenderfer, J., LaRose, A., Turnure, J. T., & Westfall, L. (2016). International energy outlook 2016 with projections to 2040. USDOE Energy Information Administration (EIA). 22. Newell, R. G., & Raimi, D. (2020). Global energy outlook comparison methods: 2020 update. Resources for the future. 23. Hayati, M. M., Aminlou, A., Zare, K., & Abapour, M. (2023). A two-stage stochastic optimization scheduling approach for integrating renewable energy sources and deferrable demand in the spinning reserve market. In 2023 8th international conference on technology and energy management (ICTEM) (pp. 1–7). 24. Aminlou, A., Hayati, M., & Zare, K. (2023). ADMM-based fully decentralized peer to peer energy trading considering a shared CAES in a local community. Journal of Energy Management and Technology. https://doi.org/10.22109/JEMT.2023.394182.1444 25. Habibi, M., Vahidinasab, V., Aghaei, J., & Mohammadi-Ivatloo, B. (2021). Assessment of energy storage systems as a reserve provider in stochastic network constrained unit commitment. IET Smart Grid, 4(2), 139–150. 26. Robbins, P., Hintz, J. G., & Moore, S. A. (2022). Environment and society: A critical introduction. Wiley. 27. Olabi, A., et al. (2022). Assessment of the pre-combustion carbon capture contribution into sustainable development goals SDGs using novel indicators. Renewable and Sustainable Energy Reviews, 153, 111710. 28. Gielen, D. et al. (2019). Global energy transformation: A roadmap to 2050. 29. IEA. (2023). CO2 Emissions in 2022. IEA, Paris https://www.iea.org/reports/co2-emissionsin-2022, License: CC BY 4.02023. Accessed 27 May 2023. 30. Newsroom, P. (2020). POSCO to establish hydrogen production capacity of 5 million tons. Available: https://newsroom.posco.com/en/posco-to-establish-hydrogen-production-capacityof-5-million-tons/. Accessed 2 Jun 2023. 31. Available: https://www.iea.org/topics/global-energy-crisis. 32. Siang, T. J., Singh, S., Omoregbe, O., Bach, L. G., Phuc, N. H. H., & Vo, D.-V. N. (2018). Hydrogen production from CH4 dry reforming over bimetallic Ni–Co/Al2O3 catalyst. Journal of the Energy Institute, 91(5), 683–694. 33. Abe, J. O., Popoola, A., Ajenifuja, E., & Popoola, O. M. (2019). Hydrogen energy, economy and storage: Review and recommendation. International Journal of Hydrogen Energy, 44(29), 15072–15086.

18

M. M. Hayati et al.

34. Genovese, M., Schlüter, A., Scionti, E., Piraino, F., Corigliano, O., & Fragiacomo, P. (2023). Power-to-hydrogen and hydrogen-to-X energy systems for the industry of the future in Europe. International Journal of Hydrogen Energy. 35. Zhiznin, S., Shvets, N., Timokhov, V., & Gusev, A. (2023). Economics of hydrogen energy of green transition in the world and Russia. Part I. International Journal of Hydrogen Energy. 36. Dincer, I., & Aydin, M. I. (2023). New paradigms in sustainable energy systems with hydrogen. Energy Conversion and Management, 283, 116950. 37. Tang, D., et al. (2023). State-of-the-art hydrogen generation techniques and storage methods: A critical review. Journal of Energy Storage, 64, 107196. 38. Aghdam, F. H., Mudiyanselage, M. W., Mohammadi-Ivatloo, B., & Marzband, M. (2023). Optimal scheduling of multi-energy type virtual energy storage system in reconfigurable distribution networks for congestion management. Applied Energy, 333, 120569. 39. Arsad, A., et al. (2023). Hydrogen electrolyser technologies and their modelling for sustainable energy production: A comprehensive review and suggestions. International Journal of Hydrogen Energy. https://doi.org/10.1016/j.ijhydene.2023.04.014 40. Vahidinasab, V., & Mohammadi-Ivatloo, B. (2023). Energy systems transition: Digitalization, decarbonization, decentralization and democratization. Springer Nature. 41. Kåberger, T. (2018). Progress of renewable electricity replacing fossil fuels. Global Energy Interconnection, 1(1), 48–52. 42. Reuß, M., Grube, T., Robinius, M., Preuster, P., Wasserscheid, P., & Stolten, D. (2017). Seasonal storage and alternative carriers: A flexible hydrogen supply chain model. Applied Energy, 200, 290–302. 43. Lebrouhi, B., Djoupo, J., Lamrani, B., Benabdelaziz, K., & Kousksou, T. (2022). Global hydrogen development-A technological and geopolitical overview. International Journal of Hydrogen Energy, 47, 7016. 44. Reuß, M., Grube, T., Robinius, M., & Stolten, D. (2019). A hydrogen supply chain with spatial resolution: Comparative analysis of infrastructure technologies in Germany. Applied Energy, 247, 438–453. 45. Wickham, D., Hawkes, A., & Jalil-Vega, F. (2022). Hydrogen supply chain optimisation for the transport sector–focus on hydrogen purity and purification requirements. Applied Energy, 305, 117740. 46. Kibria, M., McManus, D., & Bhattacharya, S. (2023). Options for net zero emissions hydrogen from Victorian lignite. Part 1: Gaseous and liquefied hydrogen. International Journal of Hydrogen Energy. 47. Sapountzaki, E., Rova, U., Christakopoulos, P., & Antonopoulou, I. (2023). Renewable hydrogen production and storage via enzymatic interconversion of CO2 and formate with electrochemical cofactor regeneration. ChemSusChem, e202202312. 48. Hassan, Q., Al-Hitmi, M., Tabar, V. S., Sameen, A. Z., Salman, H. M., & Jaszczur, M. (2023). Middle East energy consumption and potential renewable sources: An overview. Cleaner Engineering and Technology, 100599. 49. Mansour-Saatloo, A., et al. (2021). Robust decentralized optimization of multi-microgrids integrated with power-to-X technologies. Applied Energy, 304, 117635. 50. Nazari-heris, M., Jabari, F., Mohammadi-ivatloo, B., Asadi, S., & Habibnezhad, M. (2020). An updated review on multi-carrier energy systems with electricity, gas, and water energy sources. Journal of Cleaner Production, 275, 123136. 51. Moradzadeh, A., Nazari-Heris, M., Mohammadi-Ivatloo, B., & Asadi, S. (2021). Energy storage fundamentals and components. In Energy storage in energy markets (pp. 23–39). Elsevier. 52. Xu, Y., Wang, C., Huang, Y., & Fu, J. (2021). Recent advances in electrocatalysts for neutral and large-current-density water electrolysis. Nano Energy, 80, 105545. 53. Kakoulaki, G., Kougias, I., Taylor, N., Dolci, F., Moya, J., & Jäger-Waldau, A. (2021). Green hydrogen in Europe–A regional assessment: Substituting existing production with electrolysis powered by renewables. Energy Conversion and Management, 228, 113649.

1

An Overview of Energy and Exergy Analysis for Green Hydrogen Power Systems

19

54. Pukazhselvan, D., Saravanan, R., & Yadav, T. (2012). Towards sustainable green energy development and insights on few scientific problems leading to less carbon economy. Reviews in Advanced Sciences and Engineering, 1(4), 302–318. 55. Myers, D. J. et al. (2021). Roundtable on foundational science for carbon-neutral hydrogen technologies (Technology Status Document). 56. Hoang, A. T., & Nguyen, X. P. (2021). Integrating renewable sources into energy system for smart city as a sagacious strategy towards clean and sustainable process. Journal of Cleaner Production, 305, 127161. 57. Lasemi, M. A., Arabkoohsar, A., Hajizadeh, A., & Mohammadi-Ivatloo, B. (2022). A comprehensive review on optimization challenges of smart energy hubs under uncertainty factors. Renewable and Sustainable Energy Reviews, 160, 112320. 58. Lopes, J. A. P., et al. (2020). The future of power systems: Challenges, trends, and upcoming paradigms. Wiley Interdisciplinary Reviews: Energy and Environment, 9(3), e368. 59. Muttaqi, K. M., Islam, M. R., & Sutanto, D. (2019). Future power distribution grids: Integration of renewable energy, energy storage, electric vehicles, superconductor, and magnetic bus. IEEE Transactions on Applied Superconductivity, 29(2), 1–5. 60. Wang, Q., & Han, L. (2022). Hydrogen production. In Handbook of climate change mitigation and adaptation (pp. 1855–1900). Springer. 61. Zoulias, E., Varkaraki, E., Lymberopoulos, N., Christodoulou, C. N., & Karagiorgis, G. N. (2004). A review on water electrolysis. Tcjst, 4(2), 41–71. 62. Steinberg, M., & Cheng, H. C. (1989). Modern and prospective technologies for hydrogen production from fossil fuels. International Journal of Hydrogen Energy, 14(11), 797–820. 63. Colbertaldo, P., Agustin, S. B., Campanari, S., & Brouwer, J. (2019). Impact of hydrogen energy storage on California electric power system: Towards 100% renewable electricity. International Journal of Hydrogen Energy, 44(19), 9558–9576. 64. Smolinka, T., Ojong, E. T., & Garche, J. (2015). Hydrogen production from renewable energies – Electrolyzer technologies. In Electrochemical energy storage for renewable sources and grid balancing (pp. 103–128). Elsevier. 65. He, W., et al. (2020). Integration of renewable hydrogen in light-duty vehicle: Nexus between energy security and low carbon emission resources. International Journal of Hydrogen Energy, 45(51), 27958–27968. 66. Acar, C., Bicer, Y., Demir, M. E., & Dincer, I. (2019). Transition to a new era with light-based hydrogen production for a carbon-free society: An overview. International Journal of Hydrogen Energy, 44(47), 25347–25364. 67. Amores, E., Sánchez, M., Rojas, N., & Sánchez-Molina, M. (2021). Renewable hydrogen production by water electrolysis. In Sustainable fuel technologies handbook (pp. 271–313). Elsevier. 68. Yasri, N. G., & Gunasekaran, S. (2017). Electrochemical technologies for environmental remediation. In Enhancing cleanup of environmental pollutants: Volume 2: Non-biological approaches (pp. 5–73). 69. Sanghai, Y. (2013). Techno-economic analysis of hydrogen fuel cell systems used as an electricity storage technology in a wind farm with large amounts of intermittent energy. Renewable Energy, 32(4), 680–696. 70. Saur, G. (2008). Wind-to-hydrogen project: Electrolyzer capital cost study. National Renewable Energy Lab. (NREL). 71. Nojavan, S., Zare, K., & Mohammadi-Ivatloo, B. (2017). Application of fuel cell and electrolyzer as hydrogen energy storage system in energy management of electricity energy retailer in the presence of the renewable energy sources and plug-in electric vehicles. Energy Conversion and Management, 136, 404–417. 72. Kalavani, F., Mohammadi-Ivatloo, B., & Zare, K. (2019). Optimal stochastic scheduling of cryogenic energy storage with wind power in the presence of a demand response program. Renewable Energy, 130, 268–280.

20

M. M. Hayati et al.

73. Kabiri-Renani, Y., Daneshvar, M., & Mohammadi-Ivatloo, B. (2022). Transactive energy revolution: Innovative leverage for reliable operation of modern energy networks – A critical review. IET Renewable Power Generation, 16(15), 3368–3383. 74. Mirzaei, M. A., et al. (2020). Integrated energy hub system based on power-to-gas and compressed air energy storage technologies in the presence of multiple shiftable loads. IET Generation, Transmission & Distribution, 14(13), 2510–2519. 75. Nazerifard, R., Khani, L., Mohammadpourfard, M., Mohammadi-Ivatloo, B., & Akkurt, G. G. (2021). Design and thermodynamic analysis of a novel methanol, hydrogen, and power trigeneration system based on renewable energy and flue gas carbon dioxide. Energy Conversion and Management, 233, 113922. 76. Kovač, A., Paranos, M., & Marciuš, D. (2021). Hydrogen in energy transition: A review. International Journal of Hydrogen Energy, 46(16), 10016–10035. 77. Mohammadi, M., Noorollahi, Y., & Mohammadi-ivatloo, B. (2020). Fuzzy-based scheduling of wind integrated multi-energy systems under multiple uncertainties. Sustainable Energy Technologies and Assessments, 37, 100602. 78. Jiang, C., Ma, J., Corre, G., Jain, S. L., & Irvine, J. T. (2017). Challenges in developing direct carbon fuel cells. Chemical Society Reviews, 46(10), 2889–2912. 79. Figueiredo, F., & Marques, F. (2013). Electrolytes for solid oxide fuel cells. Wiley Interdisciplinary Reviews: Energy and Environment, 2(1), 52–72. 80. Zhu, B. (2006). Next generation fuel cell R&D. International Journal of Energy Research, 30(11), 895–903. 81. Miller, E. L., et al. (2016). US Department of energy hydrogen and fuel cells program: Progress, challenges and future directions. MRS Advances, 1(42), 2839–2855. 82. Miller, E. L., Thompson, S. T., Randolph, K., Hulvey, Z., Rustagi, N., & Satyapal, S. (2020). US Department of Energy hydrogen and fuel cell technologies perspectives. MRS Bulletin, 45(1), 57–64. 83. Ahluwalia, R. K., Wang, X., & Kumar, R. (2014). Fuel cell systems analysis. In US Department of Energy hydrogen and fuel cells program 2011 annual merit review and peer evaluation meeting. US Department of Energy. 84. Aminudin, M., Kamarudin, S., Lim, B., Majilan, E., Masdar, M., & Shaari, N. (2023). An overview: Current progress on hydrogen fuel cell vehicles. International Journal of Hydrogen Energy, 48(11), 4371–4388. 85. Oskouei, M. Z., Mohammadi-Ivatloo, B., Abapour, M., Shafiee, M., & Anvari-Moghaddam, A. (2021). Privacy-preserving mechanism for collaborative operation of high-renewable power systems and industrial energy hubs. Applied Energy, 283, 116338. 86. Körner, A., Tam, C., Bennett, S., & Gagné, J. (2015). Technology roadmap-hydrogen and fuel cells. International Energy Agency (IEA). 87. Han, J. et al. (2023). Hydrogen-powered smart grid resilience. In Energy conversion and economics. 88. Arias, A., Suarez, W., Rivas, E., & Mosquera, D. (2023). Virtual power plants as a management strategy to improve resilience in electrical distribution networks. In Smart technologies, systems and applications: 3rd international conference, SmartTech-IC 2022 (pp. 385–399). Springer. 89. Agabalaye-Rahvar, M., Mansour-Saatloo, A., Mirzaei, M. A., Mohammadi-Ivatloo, B., & Zare, K. (2021). Economic-environmental stochastic scheduling for hydrogen storage-based smart energy hub coordinated with integrated demand response program. International Journal of Energy Research, 45(14), 20232–20257. 90. Liu, T., et al. (2023). Key technologies and developments of multi-energy system: Three-layer framework, modelling and optimisation. Energy, 127697. 91. Ghasemi, H., Farahani, E. S., Fotuhi-Firuzabad, M., Dehghanian, P., Ghasemi, A., & Wang, F. (2023). Equipment failure rate in electric power distribution networks: An overview of concepts, estimation, and modeling methods. Engineering Failure Analysis, 107034.

1

An Overview of Energy and Exergy Analysis for Green Hydrogen Power Systems

21

92. Nasiri, T., Moeini-Aghtaie, M., Foroughi, M., & Azimi, M. (2023). Energy optimization of multi-carrier energy systems to achieve a low carbon community. Journal of Cleaner Production, 390, 136154. 93. Moradi-Sarvestani, S., Jooshaki, M., Fotuhi-Firuzabad, M., & Lehtonen, M. (2023). Incorporating direct load control demand response into active distribution system planning. Applied Energy, 339, 120897. 94. Heidari, S., & Fotuhi-Firuzabad, M. (2019). Integrated planning for distribution automation and network capacity expansion. IEEE Transactions on Smart Grid, 10(4), 4279–4288. 95. Basu, J. B. et al. (2023). Risk mitigation & profit improvement of a wind-fuel cell hybrid System with TCSC placement. In IEEE access. 96. Nasiri, N., et al. (2023). A robust bi-level optimization framework for participation of multienergy service providers in integrated power and natural gas markets. Applied Energy, 340, 121047. 97. Pandey, A. K., Singh, P. K., Nawaz, M., & Kushwaha, A. K. (2023). Forecasting of non-renewable and renewable energy production in India using optimized discrete grey model. Environmental Science and Pollution Research, 30(3), 8188–8206. 98. Zheng, J., Du, J., Wang, B., Klemeš, J. J., Liao, Q., & Liang, Y. (2023). A hybrid framework for forecasting power generation of multiple renewable energy sources. Renewable and Sustainable Energy Reviews, 172, 113046. 99. Manimegalai, V., Rukkumani, V., Gayathri, A., Pandiyan, P., & Mohanapriya, V. (2023). An overview of global renewable energy resources. Renewable Energy and AI for Sustainable Development, 2(4), 2–5. 100. Salyani, P., Nourollahi, R., Zare, K., & Razzaghi, R. (2023). A cooperative game approach for optimal resiliency-oriented scheduling of transactive multiple microgrids. Sustainable Cities and Society, 89, 104358. 101. Hassan, Q., et al. (2023). Renewable energy-to-green hydrogen: A review of main resources routes, processes and evaluation. International Journal of Hydrogen Energy. 102. Liu, Z., & Nazari-Heris, M. (2023). Robust bidding strategy of interconnected multi-carrier systems in the electricity markets under the uncertainty in electricity load. Sustainable Energy Technologies and Assessments, 57, 103245. 103. Nazari-Heris, M., & Asadi, S. (2023). Reliable energy management of residential buildings with hybrid energy systems. Journal of Building Engineering, 71, 106531. 104. Nozarian, M., Fereidunian, A., & Barati, M. (2023). Reliability-oriented planning framework for smart cities: From interconnected micro energy hubs to macro energy hub scale. IEEE Systems Journal. 105. Carvallo, C., Jalil-Vega, F., & Moreno, R. (2023). A multi-energy multi-microgrid system planning model for decarbonisation and decontamination of isolated systems. Applied Energy, 343, 121143. 106. Khouzestani, L. B., Sheikh-El-Eslami, M. K., Salemi, A. H., & Moghaddam, I. G. (2023). Virtual smart energy hub: A powerful tool for integrated multi energy systems operation. Energy, 265, 126361. 107. Li, R., et al. (2023). Techno-economic analysis of a wind-photovoltaic-electrolysis-battery hybrid energy system for power and hydrogen generation. Energy Conversion and Management, 281, 116854. 108. Azadnia, A. H., McDaid, C., Andwari, A. M., & Hosseini, S. E. (2023). Green hydrogen supply chain risk analysis: A European hard-to-abate sectors perspective. Renewable and Sustainable Energy Reviews, 182, 113371. 109. Li, C., et al. (2023). A multi-objective planning method for multi-energy complementary distributed energy system: Tackling thermal integration and process synergy. Journal of Cleaner Production, 135905. 110. Liu, C., Wang, H., Wang, Z., Liu, Z., Tang, Y., & Yang, S. (2022). Research on life cycle low carbon optimization method of multi-energy complementary distributed energy system: A review. Journal of Cleaner Production, 130380.

22

M. M. Hayati et al.

111. Wu, X., Liao, B., Su, Y., & Li, S. (2023). Multi-objective and multi-algorithm operation optimization of integrated energy system considering ground source energy and solar energy. International Journal of Electrical Power & Energy Systems, 144, 108529. 112. Yang, Y., Guo, S., Liu, Q., & Li, R. (2019). Day-ahead scheduling for a new wind-CSP hybrid system. Energy Procedia, 158, 6254–6259. 113. Chennaif, M., Zahboune, H., Elhafyani, M., & Zouggar, S. (2021). Electric system cascade extended analysis for optimal sizing of an autonomous hybrid CSP/PV/wind system with battery energy storage system and thermal energy storage. Energy, 227, 120444. 114. Wahbah, M., Mohandes, B., El-Fouly, T. H. M., & El Moursi, M. S. (2022). Unbiased crossvalidation kernel density estimation for wind and PV probabilistic modelling. Energy Conversion and Management, 266, 115811. 115. Yang, J., Yang, Z., & Duan, Y. (2021). Load matching and techno-economic analysis of CSP plant with S–CO2 Brayton cycle in CSP-PV-wind hybrid system. Energy, 223, 120016. 116. Al-Ghussain, L., Ahmad, A. D., Abubaker, A. M., & Hassan, M. A. (2022). Exploring the feasibility of green hydrogen production using excess energy from a country-scale 100% solarwind renewable energy system. International Journal of Hydrogen Energy, 47(51), 21613–21633. 117. Najjar, Y. S. H. (2013). Hydrogen safety: The road toward green technology. International Journal of Hydrogen Energy, 38(25), 10716–10728. 118. Acar, C., & Dincer, I. (2019). Review and evaluation of hydrogen production options for better environment. Journal of Cleaner Production, 218, 835–849. 119. Ruiming, F. (2019). Multi-objective optimized operation of integrated energy system with hydrogen storage. International Journal of Hydrogen Energy, 44(56), 29409–29417. 120. Wang, Z., Zhang, X., & Rezazadeh, A. (2021). Hydrogen fuel and electricity generation from a new hybrid energy system based on wind and solar energies and alkaline fuel cell. Energy Reports, 7, 2594–2604. 121. Li, Z., et al. (2021). Modeling and analysis of hydrogen storage wind and gas complementary power generation system. Energy Exploration & Exploitation, 39(4), 1306–1323. 122. Sezer, N., Biçer, Y., & Koç, M. (2019). Design and analysis of an integrated concentrated solar and wind energy system with storage. International Journal of Energy Research, 43(8), 3263–3283. 123. Carton, J. G., & Olabi, A. G. (2010). Wind/hydrogen hybrid systems: Opportunity for Ireland’s wind resource to provide consistent sustainable energy supply. Energy, 35(12), 4536–4544. 124. Luo, L., Cristofari, C., & Levrey, S. (2023). Cogeneration: Another way to increase energy efficiency of hybrid renewable energy hydrogen chain–a review of systems operating in cogeneration and of the energy efficiency assessment through exergy analysis. Journal of Energy Storage, 66, 107433. 125. Gahleitner, G. (2013). Hydrogen from renewable electricity: An international review of power-to-gas pilot plants for stationary applications. International Journal of Hydrogen Energy, 38(5), 2039–2061. 126. Abbasi, T., & Abbasi, S. A. (2011). ‘Renewable’ hydrogen: Prospects and challenges. Renewable and Sustainable Energy Reviews, 15(6), 3034–3040. 127. DeLovato, N., Sundarnath, K., Cvijovic, L., Kota, K., & Kuravi, S. (2019). A review of heat recovery applications for solar and geothermal power plants. Renewable and Sustainable Energy Reviews, 114, 109329. 128. Uddin, S. N., & Barreto, L. (2007). Biomass-fired cogeneration systems with CO2 capture and storage. Renewable Energy, 32(6), 1006–1019. 129. Ren, S., Feng, X., & Wang, Y. (2021). Emergy evaluation of the integrated gasification combined cycle power generation systems with a carbon capture system. Renewable and Sustainable Energy Reviews, 147, 111208. 130. Jovan, D. J., Dolanc, G., & Pregelj, B. (2021). Cogeneration of green hydrogen in a cascade hydropower plant. Energy Conversion and Management, 10, 100081.

1

An Overview of Energy and Exergy Analysis for Green Hydrogen Power Systems

23

131. Assareh, E., et al. (2022). Thermodynamic assessment of a cogeneration system with CSP driven-Brayton and Rankine cycles for electric power and hydrogen production in the framework of the energy and water nexus. Energy Nexus, 5, 100031. 132. Asgari, A., Yari, M., & Mahmoudi, S. M. S. (2022). Exergy and exergoeconomic analyses and multi-objective optimization of a novel cogeneration system for hydrogen and cooling production. International Journal of Hydrogen Energy, 47(62), 26114–26134. 133. Altayib, K., & Dincer, I. (2022). Development of an integrated hydropower system with hydrogen and methanol production. Energy, 240, 122780. 134. Yilanci, A., Dincer, I., & Ozturk, H. K. (2009). A review on solar-hydrogen/fuel cell hybrid energy systems for stationary applications. Progress in Energy and Combustion Science, 35(3), 231–244. 135. Guo, S., Liu, Q., Sun, J., & Jin, H. (2018). A review on the utilization of hybrid renewable energy. Renewable and Sustainable Energy Reviews, 91, 1121–1147. 136. Soleymani, E., Ghavami Gargari, S., & Ghaebi, H. (2021). Thermodynamic and thermoeconomic analysis of a novel power and hydrogen cogeneration cycle based on solid SOFC. Renewable Energy, 177, 495–518. 137. Boulmrharj, S., Khaidar, M., Bakhouya, M., Ouladsine, R., Siniti, M., & Zine-dine, K. (2020). Performance assessment of a hybrid system with hydrogen storage and fuel cell for cogeneration in buildings. Sustainability, 12(12), 4832. 138. Özgür, T., & Yakaryılmaz, A. C. (2018). A review: Exergy analysis of PEM and PEM fuel cell based CHP systems. International Journal of Hydrogen Energy, 43(38), 17993–18000. 139. Bamisile, O., et al. (2023). Thermo-enviro-exergoeconomic analysis and multi-objective optimization of a novel geothermal-solar-wind micro-multi-energy system for cleaner energy production. Process Safety and Environmental Protection, 170, 157–175. 140. Di Micco, S., Romano, F., Jannelli, E., Perna, A., & Minutillo, M. (2023). Techno-economic analysis of a multi-energy system for the co-production of green hydrogen, renewable electricity and heat. International Journal of Hydrogen Energy. 141. Østergaard, P. A., & Lund, H. (2011). A renewable energy system in Frederikshavn using low-temperature geothermal energy for district heating. Applied Energy, 88(2), 479–487. 142. Hassan, H., Yousef, M. S., & Abo-Elfadl, S. (2021). Energy, exergy, economic and environmental assessment of double pass V-corrugated-perforated finned solar air heater at different air mass ratios. Sustainable Energy Technologies and Assessments, 43, 100936. 143. Shaygan, M., Ehyaei, M. A., Ahmadi, A., Assad, M. E. H., & Silveira, J. L. (2019). Energy, exergy, advanced exergy and economic analyses of hybrid polymer electrolyte membrane (PEM) fuel cell and photovoltaic cells to produce hydrogen and electricity. Journal of Cleaner Production, 234, 1082–1093. 144. Nosratabadi, S. M., Hemmati, R., Bornapour, M., & Abdollahpour, M. (2021). Economic evaluation and energy/exergy analysis of PV/wind/PEMFC energy resources employment based on capacity, type of source and government incentive policies: Case study in Iran. Sustainable Energy Technologies and Assessments, 43, 100963. 145. Colakoglu, M., & Durmayaz, A. (2022). Energy, exergy and economic analyses and multiobjective optimization of a novel solar multi-generation system for production of green hydrogen and other utilities. International Journal of Hydrogen Energy, 47(45), 19446–19462. 146. Incer-Valverde, J., Mörsdorf, J., Morosuk, T., & Tsatsaronis, G. (2023). Power-to-liquid hydrogen: Exergy-based evaluation of a large-scale system. International Journal of Hydrogen Energy, 48(31), 11612–11627. 147. Minutillo, M., Perna, A., & Sorce, A. (2020). Green hydrogen production plants via biogas steam and autothermal reforming processes: Energy and exergy analyses. Applied Energy, 277, 115452. 148. Nalbant Atak, Y., Ozgur Colpan, C., & Iulianelli, A. (2022). Energy and exergy analyses of an integrated membrane reactor and CO2 capture system to generate decarbonized hydrogen. Energy Conversion and Management, 272, 116367.

24

M. M. Hayati et al.

149. Nasser, M., Megahed, T. F., Ookawara, S., & Hassan, H. (2022). Performance evaluation of PV panels/wind turbines hybrid system for green hydrogen generation and storage: Energy, exergy, economic, and enviroeconomic. Energy Conversion and Management, 267, 115870. 150. Nikitin, A., et al. (2023). Energy, exergy, economic and environmental (4E) analysis using a renewable multi-generation system in a near-zero energy building with hot water and hydrogen storage systems. Journal of Energy Storage, 62, 106794. 151. ÖZdemİR, A., & GenÇ, G. (2022). A comprehensive comparative energy and exergy analysis in solar based hydrogen production systems. International Journal of Hydrogen Energy, 47(24), 12189–12203. 152. Abuşoğlu, A., Özahi, E., Kutlar, A. İ., & Demir, S. (2017). Exergy analyses of green hydrogen production methods from biogas-based electricity and sewage sludge. International Journal of Hydrogen Energy, 42(16), 10986–10996. 153. Qi, Y., & Huang, D. (2022). Energy and exergy analysis of supercritical/transcritical CO2 cycles for water injected hydrogen gas turbine. Energy, 260, 124931. 154. Arslan, M., & Yılmaz, C. (2023). Investigation of green hydrogen production and development of waste heat recovery system in biogas power plant for sustainable energy applications. International Journal of Hydrogen Energy. 155. Yang, Y., et al. (2023). A novel integrated system of hydrogen liquefaction process and liquid air energy storage (LAES): Energy, exergy, and economic analysis. Energy Conversion and Management, 280, 116799. 156. Liu, L., Zhai, R., & Hu, Y. (2023). Performance evaluation of wind-solar-hydrogen system for renewable energy generation and green hydrogen generation and storage: Energy, exergy, economic, and enviroeconomic. Energy, 276, 127386.

Chapter 2

Hydrogen-Incorporated Sector-Coupled Smart Grids: A Systematic Review and Future Concepts Mohammad Mohsen Hayati and Arman Oshnoei

2.1

, Ashkan Safari

, Morteza Nazari-Heris

,

Introduction

In the current era, renewable energy sources (RESs) have become seamlessly integrated within smart grids on a widespread scale [1]. These sources encompass various forms of energy storage, including batteries, solar photovoltaics, wind, thermal, and hydrogen, and they hold a significant position within the framework of smart grids [2, 3]. Presently, the majority of research efforts are focused on exploring the potential of renewable energy within smart grid systems. Hydrogen presents substantial opportunities as a promising fuel for the future, carrying a multitude of social, economic, and environmental implications. Hydrogen plays a crucial and innovative role within the context of a smart grid, serving as a versatile solution that operates as an energy carrier, storage medium, and a clean fuel cell. Its integration with the smart grid allows for the effective mitigation of environmental impacts and the achievement of optimal sustainability [4]. This signifies a M. M. Hayati (✉) Faculty of Electrical and Computer Engineering, Energy Systems Research Institute (ESRI), Smart Energy Systems Lab, University of Tabriz, Tabriz, Iran e-mail: [email protected] A. Safari Faculty of Electrical and Computer Engineering, University of Tabriz, Tabriz, Iran M. Nazari-Heris Lawrence Technological University, Southfield, Michigan, USA A. Oshnoei (✉) Faculty of Electrical and Computer Engineering, Energy Systems Research Institute (ESRI), Smart Energy Systems Lab, University of Tabriz, Tabriz, Iran Department of Energy, Aalborg University, Aalborg, Denmark e-mail: [email protected] © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 V. Vahidinasab et al. (eds.), Green Hydrogen in Power Systems, Green Energy and Technology, https://doi.org/10.1007/978-3-031-52429-5_2

25

26

M. M. Hayati et al.

progressive shift toward a society centered around hydrogen utilization, paving the way for the development of HISCSG. Hydrogen’s versatility as an energy carrier makes it a valuable asset within smart grids. It can be produced through various methods, including electrolysis, steam methane reforming, and biomass gasification [5]. This flexibility enables the integration of hydrogen production with intermittent renewable energies (REs), helping to balance the supply–demand dynamics of the grid. Furthermore, hydrogen can be stored in large quantities, providing long-term energy storage capabilities that complement the inherent intermittency of renewable sources. The integration of hydrogen as a storage medium within smart grids enhances grid stability and resilience. By converting excess renewable energy into hydrogen during periods of low demand, surplus energy can be efficiently stored and later converted back into electricity or utilized in other energy-intensive sectors such as transportation, industry, and heating. This feature ensures the maximization of renewable energy utilization, minimizing curtailment and grid congestion, and facilitating the integration of larger shares of intermittent renewable energy. In addition to its role as an energy carrier and storage medium, hydrogen’s use as a clean fuel cell further expands its significance within smart grids. Hydrogen fuel cells offer high energy efficiency and emit only water vapor as a by-product, making them an environmentally friendly alternative to conventional combustion-based technologies [6]. By integrating hydrogen fuel cells into various applications within the smart grid, such as distributed power generation, transportation, and heating systems, the overall carbon footprint of the energy sector can be significantly reduced [7]. The adoption of HISCSG represents a paradigm shift toward a more sustainable and resilient energy system. This concept encompasses the integration of multiple sectors, including power, transportation, industry, and buildings, into a cohesive and interconnected network. By incorporating hydrogen as a versatile solution across these sectors, the smart grid becomes a nexus for optimizing energy production, consumption, and storage, thereby contributing to the decarbonization of the overall energy system [8]. In this chapter, an overview of the important issues implementing HISCSG has been done. First, an introduction to Hydrogen Integration in smart grids has been made, and then the concepts of sector coupling in smart grids, energy management, electricity market, and hydrogen economics have been reviewed. Also, the focus has been placed on technological advancements and innovations concerning hydrogen integration in smart grid systems. The discussion has encompassed emerging technologies for hydrogen production, storage, and distribution, with particular attention given to advancements in electrolysis, renewable hydrogen sources, and storage methods. Furthermore, the integration of renewable energy sources with hydrogen in smart grids has been examined, emphasizing the potential benefits derived from the utilization of hydrogen as a means to store and balance intermittent renewable energy generation. Similarly, the significance of smart grid control and management systems for efficient hydrogen utilization has been emphasized, highlighting the importance of real-time monitoring, demand-response strategies, and optimization algorithms. Economic and environmental considerations have been addressed, including the cost analysis of hydrogen-integrated smart grid systems, evaluation

2

Hydrogen-Incorporated Sector-Coupled Smart Grids: A Systematic Review. . .

27

of environmental impacts and sustainability aspects, and the necessity for economic and financial incentives to promote hydrogen integration. Additionally, key challenges and gaps in current hydrogen-integrated smart grid systems have been identified, future prospects and potential advancements have been explored, and research directions for further investigation have been outlined. Ultimately, the findings of this review underscore the significance of hydrogen integration in sector-coupled smart grids, offering opportunities for decarbonization, grid flexibility, and energy security. To fully leverage these opportunities, future endeavors should focus on continued research, supportive policies, collaboration, and demonstration projects.

2.2

Fundamentals of Hydrogen Integration in Smart Grids

The integration of hydrogen into smart grids encompasses various fundamental aspects. This integration involves the utilization of hydrogen as an energy carrier within the smart grid infrastructure. One fundamental aspect is hydrogen production through electrolysis, where electricity derived from REs is used to split water molecules into hydrogen and oxygen. This green hydrogen (GH2) production process ensures the use of clean and sustainable energy sources. Another key aspect is the storage and distribution of hydrogen within the smart grid. Hydrogen can be stored in various forms, such as compressed hydrogen gas, liquid hydrogen, or chemical compounds. Efficient storage systems are crucial to ensure the availability of hydrogen during periods of high demand or when renewable energy generation is low. The integration of hydrogen into the smart grid also requires the development of appropriate infrastructure [9]. This includes hydrogen pipelines, storage facilities, and hydrogen refueling stations for fuel cell vehicles. The existing electricity grid may need upgrades or modifications to accommodate the transportation and distribution of hydrogen. In addition to storage and distribution, the utilization of hydrogen within the smart grid is another fundamental aspect. Hydrogen can be used in fuel cells to generate electricity on-site or be converted back into electricity during peak demand periods. It can also be utilized for heating, industrial processes, and as a feedstock for chemical production, contributing to sector coupling and energy system integration. To ensure effective integration, advanced control and communication systems are required to manage and optimize the flow of hydrogen within the smart grid [10, 11]. These systems enable real-time monitoring, demandresponse capabilities, and efficient utilization of hydrogen resources. Furthermore, policy and regulatory frameworks play a vital role in facilitating the integration of hydrogen into smart grids. Supportive policies, incentives, and market mechanisms can encourage investment, research, and development of hydrogen technologies and infrastructure.

28

2.2.1

M. M. Hayati et al.

Introduction to Hydrogen as an Energy Carrier

Hydrogen is gaining attention as a promising energy carrier for a sustainable, lowcarbon future. It offers advantages such as abundance, efficient storage and transportation, and minimal carbon footprint when produced using renewable energy sources. Its versatility allows for various applications, including fuel cells for electricity generation. With its high energy density, hydrogen is suitable for longterm storage, enabling the integration of renewable energy into the grid. Furthermore, hydrogen has the potential to facilitate decarbonization in transportation and industrial sectors. However, challenges exist, including the need for significant investments in infrastructure and technology advancements for large-scale production using renewable energy sources. Transportation and storage considerations also require specialized infrastructure [12, 13].

2.2.2

Benefits and Challenges of Hydrogen Integration

Hydrogen integration offers numerous benefits and opportunities for the energy sector, but it also presents various challenges that need to be addressed, as in Fig. 2.1.

2.2.3

Benefits of Hydrogen Integration

2.2.3.1

Decarbonization

Hydrogen is a clean energy carrier that, when produced using renewable sources, has minimal or zero greenhouse gas emissions. Its integration can significantly contribute to decarbonizing sectors such as transportation, industry, and heating, reducing reliance on fossil fuels and mitigating climate change.

Fig. 2.1 Benefits of hydrogen integration in smart grids

2

Hydrogen-Incorporated Sector-Coupled Smart Grids: A Systematic Review. . .

2.2.3.2

29

Energy Storage

Hydrogen can be an efficient and scalable energy storage solution, addressing the intermittent nature of REs. Excess renewable energy can be used to produce hydrogen through electrolysis and stored for later use, providing a means for balancing energy supply and demand [14].

2.2.3.3

Versatility and Flexibility

Hydrogen can be utilized across various sectors and applications. It can be used in fuel cells to generate electricity, powering vehicles, homes, and businesses. Additionally, hydrogen can serve as a feedstock for industrial processes, including chemical production and refining [15].

2.2.3.4

Energy Independence and Security

Hydrogen offers an opportunity to diversify energy sources and reduce dependence on imported fossil fuels. Countries can leverage domestic resources and develop hydrogen production capabilities, enhancing energy independence and security [16].

2.2.3.5

Air Quality Improvement

Hydrogen fuel cell vehicles produce zero tailpipe emissions, contributing to improved air quality and reduced pollution in urban areas, particularly in densely populated regions [17].

2.2.4

Challenges of Hydrogen Integration

2.2.4.1

Cost and Infrastructure

The cost of producing, storing, and distributing hydrogen is currently higher compared to traditional fossil fuels. Scaling up hydrogen infrastructure and technology advancements are needed to reduce costs and establish an efficient and reliable hydrogen supply chain [18].

2.2.4.2

Hydrogen Production

While hydrogen can be produced through various methods, most commercially available hydrogen is derived from natural gas, resulting in carbon emissions.

30

M. M. Hayati et al.

Expanding the production of GH2 through electrolysis, using renewable energy, requires significant investments and supportive policies [19].

2.2.4.3

Storage and Transportation

Hydrogen has a low energy density, making storage and transportation challenging. It requires specialized infrastructure, such as pipelines or cryogenic tanks, and safety measures to handle its characteristics, including high flammability and leakage risks [20].

2.2.4.4

System Integration

Integrating hydrogen into existing energy systems and infrastructure poses technical challenges. Adapting grids, retrofitting vehicles, and ensuring compatibility with current technologies require careful planning, standardization, and coordination among stakeholders [21].

2.2.4.5

Market Development and Regulations

Establishing a robust market for hydrogen requires supportive incentives and regulations. Ensuring fair competition, incentivizing investments, and establishing safety and environmental standards are essential for market development [22].

2.3 2.3.1

Sector Coupling in Smart Grids Definition and Principles of Sector Coupling

Sector coupling in smart grids refers to the integration and coordination of different energy sectors, such as electricity, heating and cooling, transportation, and industry, to optimize overall system efficiency and reliability [23, 24]. It involves connecting these traditionally separate sectors through advanced technologies, information systems, and market mechanism. The concept of sector coupling recognizes that various energy sectors are interconnected and that leveraging synergies between them can lead to more efficient energy systems [25]. Sector coupling and GH2 are interconnected concepts that play a significant role in advancing the transition to a sustainable and decarbonized energy system [26]. In addition, this concept is implemented in Fig. 2.2.

2

Hydrogen-Incorporated Sector-Coupled Smart Grids: A Systematic Review. . .

31

Fig. 2.2 Sector coupling in smart grids – an integrated energy system based on hydrogen and renewable energy resources

2.3.1.1

Digitalization and Automation

Smart grid technologies play a crucial role in sector coupling. Digitalization and automation enable the efficient monitoring, control, and coordination of energy flows across different sectors. Advanced sensors, data analytics, and communication systems facilitate real-time decision-making and optimize energy management [27].

2.3.1.2

Decentralization and Local Energy Systems

Sector coupling encourages the development of decentralized and local energy systems. By utilizing distributed energy resources (DERs) such as rooftop solar

32

M. M. Hayati et al.

panels, battery storage, and electric vehicles, local communities can actively participate in sector coupling, promoting energy self-sufficiency and resilience [28, 29].

2.3.2

Power-to-X in Sector-Coupled Smart Grids

Electricity has the capacity to be utilized directly in various sectors such as through the use of battery electric vehicles [30]. Alternatively, it can be converted into alternative energy carriers that offer greater versatility in their applications and improved storage capabilities. These concepts, collectively known as power-to-X (PtX), involve the transformation of electrical energy into different products [31, 32]. The transition toward a more sustainable energy system necessitates a comprehensive approach that integrates low-carbon energy sources, energy efficiency measures, and the interconnection of various energy sectors [21]. Within this framework, the implementation of “power-to-hydrogen” concepts has garnered considerable attention in recent years. These concepts serve to address demand management, facilitate seasonal storage, and establish connections between different sectors, contributing to the overall sustainability objectives. PtX refers to a set of technologies that convert electrical power into different forms of energy or energy carriers such as hydrogen, synthetic fuels, or chemicals [33]. In the context of sectorcoupled smart grids, PtX technologies play a crucial role in integrating various sectors, including electricity, heat, transportation, and industry, into a unified and efficient energy system [34]. By 2020, substantial advancements had been achieved in PtX research initiatives across the Europe Union, with a notable proportion having reached completion or being actively planned [35]. Furthermore, the smart grid infrastructure and advanced energy management systems facilitate the optimal coordination and utilization of PtX technologies based on real-time electricity generation, demand, and pricing conditions. It is important to note that the deployment and scaling of PtX technologies in sector-coupled smart grids require supportive policies, appropriate infrastructure, and market mechanisms to incentivize their adoption and ensure their economic viability. Additionally, the environmental impacts, energy efficiency, and overall lifecycle emissions associated with PtX technologies should be carefully evaluated to ensure their contribution to sustainable and low-carbon energy systems [36]. The main concept of using PtX in SCSGs are as follows:

2.3.2.1

Power-to-Hydrogen (PtH2)

PtH2 involves using excess electricity from renewable sources to electrolyze water and produce hydrogen gas (H2). The produced hydrogen can be stored, transported, or utilized as a feedstock for various industrial processes [37, 38]. In sector-coupled smart grids, PtH2 enables the storage of surplus electricity during periods of high

2

Hydrogen-Incorporated Sector-Coupled Smart Grids: A Systematic Review. . .

33

generation and its conversion back into electricity or other forms of energy when demand exceeds supply [39].

2.3.2.2

Power-to-Gas (PtG)

PtG technologies convert excess electrical power into gaseous energy carriers such as hydrogen or methane. The produced gas can be injected into the natural gas grid, stored, or used as a fuel for various applications, including heating, transportation, and industrial processes [40, 41]. PtG allows for the utilization of renewable electricity in sectors that are traditionally reliant on fossil fuels, contributing to decarbonization efforts [42].

2.3.2.3

Power-to-Liquid (PtL)

PtL technologies convert electrical power into liquid energy carriers, such as synthetic fuels. These synthetic fuels, including synthetic diesel, gasoline, or aviation fuel, can be used as drop-in replacements for conventional fossil fuels in transportation or as feedstocks for the chemical industry [43]. PtL enables the direct utilization of renewable electricity in sectors that are challenging to electrify, such as long-haul transportation or aviation [44].

2.3.2.4

Power-to-Heat (PtH)

PtH technologies use excess electrical power to produce heat for various applications, including space heating, district heating, or industrial processes. This can be achieved through resistive heating, heat pumps, or thermal energy storage systems [45]. PtH allows for the efficient use of surplus electricity while meeting the heating demands of residential, commercial, or industrial buildings [46].

2.3.2.5

Power-to-Mobility

Electric vehicles (EVs) play a vital role in sector coupling by acting as mobile energy storage units. EVs can be charged using excess renewable electricity and, in turn, supply power back to the grid or be used for other energy applications when parked [47]. Hydrogen-to-power and hydrogen-to-gas are two different applications of hydrogen as an energy carrier. The following is an overview of each:

34

2.3.2.6

M. M. Hayati et al.

Hydrogen-to-Power (H2-to-Power)

H2-to-power refers to the use of hydrogen as a fuel to generate electricity. This is typically achieved through hydrogen fuel cells, which electrochemically convert hydrogen and oxygen into electricity, with water as the only by-product. Hydrogen fuel cells offer high energy efficiency, zero-emission operation, and can be used in various applications, including transportation (e.g., fuel cell vehicles), stationary power generation, and portable devices. In the context of sector-coupled smart grids, H2-to-power can play a role in storing excess renewable energy and providing on-demand electricity when needed [48].

2.3.2.7

Hydrogen-to-Gas (H2-to-Gas)

H2-to-gas, also known as power-to-gas, involves using surplus renewable electricity to produce hydrogen through electrolysis [42]. The produced hydrogen can then be further converted into synthetic natural gas (methane) through a process called methanation. This synthetic natural gas can be injected into the existing natural gas grid or used as a fuel for various applications, including heating, power generation, and transportation. H2-to-gas allows for the long-term storage of renewable energy and offers a means of integrating renewable electricity with the existing gas infrastructure. It also provides a pathway for utilizing renewable energy in sectors that are currently reliant on natural gas [49, 50]. Both H2-to-power and H2-to-gas are important applications of hydrogen that contribute to the decarbonization of various sectors. H2-to-power enables the direct conversion of hydrogen into electricity, providing a clean and efficient alternative to traditional combustion-based power generation. H2-to-gas, on the other hand, allows for the storage, transportation, and utilization of hydrogen in existing gas infrastructure, providing a means of sectoral integration and enabling the use of renewable energy in sectors traditionally dependent on fossil fuels. Hydrogen-to-industry and hydrogen-to-chemical are two applications of hydrogen in the industrial and chemical sectors. The following is an overview of each:

2.3.2.8

Hydrogen-to-Industry

Hydrogen has various applications in industrial processes. It can be used as a feedstock or fuel in industries such as refining, ammonia production, and metal processing. For example, hydrogen is commonly used in the Haber–Bosch process for ammonia synthesis, which is a key component in fertilizer production. Hydrogen can also be utilized as a reducing agent in metallurgical processes, such as iron and steel production, to remove impurities. Additionally, hydrogen can be employed in the production of various chemicals and materials, including methanol, hydrochloric

2

Hydrogen-Incorporated Sector-Coupled Smart Grids: A Systematic Review. . .

35

acid, and plastics. The use of hydrogen in industry can help reduce carbon emissions and enhance the sustainability of industrial processes [51].

2.3.2.9

Hydrogen-to-Chemical

Hydrogen is a valuable building block in the chemical industry. It serves as a key input in several chemical processes, including hydrogenation, hydrocracking, and methanol synthesis. These processes involve the reaction of hydrogen with other compounds to produce a wide range of chemicals and materials. Hydrogen can be used to produce methanol, which is a versatile chemical that can be further converted into other products such as formaldehyde, acetic acid, or synthetic hydrocarbons. Hydrogen is also used in the production of important chemicals such as ammonia, propylene, and hydrogen peroxide. The availability of hydrogen from renewable sources can contribute to the greening of the chemical industry by reducing reliance on fossil fuel-based hydrogen production methods [52].

2.3.2.10

Hydrogen-to-Fuel

H2-to-fuel refers to the use of hydrogen as a fuel source for various applications such as hydrogen fuel cell vehicles (FCVs) and hydrogen internal combustion engines (HICE), particularly in the transportation sector [53]. Figure 2.3 illustrates the general framework of power-to-X in sector-coupled smart grids. A network in which renewable energy sources and transportation system, gas network and energy storage facilities are actively present, it is expected that future smart grids will be like this.

2.3.3

Energy Management

In HISCSG, effective energy management entails the integration and optimization of hydrogen-based technologies within smart grid systems [14]. The objective is to capitalize on the flexibility and storage capabilities of hydrogen to enhance the overall efficiency, reliability, and sustainability of the energy system. It is crucial to recognize the significance of managing a comprehensive range of energy components, including electricity, heat, and hydrogen, within smart grids to ensure optimal performance. The increasing acknowledgment of electrochemistry’s role within energy systems highlights its capacity to facilitate the clean and efficient conversion of chemical and electrical energy sourced from hydrogen [54]. This underscores hydrogen’s adaptability in addressing various research challenges in the energy system. In a broader context, the term “smart energy system” encompasses control and management systems that extend beyond the scope of smart grids. Presently, the design of energy

36

M. M. Hayati et al.

Fig. 2.3 Power-to-X framework in sector-coupled smart grids

systems reliant on renewable energy sources necessitates meticulous integration of diverse energy sources and carriers, with a specific emphasis on intelligent energy networks [55]. The following are some key aspects of energy management in HISCSG:

2.3.3.1

Demand-Side Management (DSM)

Energy management in HISCSG involves demand-side management strategies to optimize energy consumption and demand patterns. This can include load shifting, demand-response programs, and energy efficiency measures to align energy usage

2

Hydrogen-Incorporated Sector-Coupled Smart Grids: A Systematic Review. . .

37

with hydrogen availability and grid conditions. DSM pertains to the administration of the electricity market, involving both the electrical supply and demand sides, with the objective of enhancing power supply reliability and reducing energy consumption for both parties involved. Additionally, given the intermittent characteristics of renewable energy sources, including hydrogen, it becomes imperative to balance the supply and demand sides within smart grids to ensure superior service quality [56, 57].

2.3.3.2

Demand Response

The design of smart grids must go beyond simply the technology, the importance of the smart users in demand-side management, which actively participate in energy, should be recognized for future development [58, 59]. The electricity demandresponse program (DRP) is the main solution for DSM [60]. Energy management systems in hydrogen-incorporated smart grids can leverage demand-response mechanisms to optimize hydrogen utilization. This involves incentivizing consumers to adjust their energy consumption based on the availability of hydrogen or the overall grid conditions. For example, consumers can be encouraged to shift their electricity usage to periods of high renewable energy generation or when excess hydrogen is available [61].

2.3.3.3

System Optimization

Energy management algorithms need to optimize the overall system operation by considering various factors such as energy generation, storage, conversion, grid stability, and economic considerations [62]. This involves forecasting energy demand, optimizing hydrogen production and storage, and scheduling the dispatch of electricity and hydrogen resources to maximize efficiency and minimize costs [63].

2.3.3.4

Energy Market Integration

HISCSG require appropriate market mechanisms to incentivize the deployment and efficient operation of hydrogen technologies [64]. Energy management systems need to interface with energy markets, enabling the participation of hydrogen producers, consumers, and storage operators in energy trading and ancillary services markets [65, 66].

38

2.3.4

M. M. Hayati et al.

Power Market

The power market in HISCSG requires appropriate pricing and market mechanisms to facilitate efficient resource allocation. This market functions as a mechanism where electricity prices are influenced by factors such as supply, demand, and overall electricity transactions [67, 68]. Furthermore, there is a proposed framework for a localized energy market that incorporates the trading of both electricity and hydrogen. Case studies have shown that this approach enhances the integration of local renewable energy sources while reducing peak demand [69]. An electricity retailer, authorized to procure electricity from the market and distribute it to end-users, plays a key role [70]. To optimize expected profits, electricity retailers determine retail prices for various consumer segments, effectively manage price fluctuations and ensure robust scheduling [71, 72].

2.3.5

Hydrogen Economy

The hydrogen economy represents a prospective economic framework that utilizes hydrogen as a versatile medium for storage, transportation, and conversion [73, 74]. Its inception dates back to the 1970s, and it has gained significant attention due to the sustainable advancements in hydrogen fuel cells and other hydrogenrelated products across multiple industries [75]. By incorporating hydrogen into smart grids, more renewable energy can be effectively integrated, providing energy to all sectors within the energy system without the necessity of costly expansions in grid capacity. Researcher in Ref. [76] have introduced the concept of the “zeroenergy hydrogen economy,” which positions hydrogen as the primary energy vector. In the context of smart grids, literature on the hydrogen economy can be categorized into two main areas: demand-side management (DSM) and electricity market [77, 78]. Consequently, an outline of hydrogen economy fundamentals is drawn in Fig. 2.4.

2.4

Hydrogen-Incorporated Smart Grid Projects

These projects demonstrate the ongoing efforts to integrate hydrogen into smart grid systems, leveraging its potential as a versatile energy carrier. By combining hydrogen production, storage, and utilization with advanced grid technologies, these projects contribute to a more resilient energy infrastructure [79, 80]. While there are several ongoing projects worldwide, the following is an overview of a few notable examples, also summarized in Table 2.1.

2

Hydrogen-Incorporated Sector-Coupled Smart Grids: A Systematic Review. . .

39

Fig. 2.4 The fundamental framework outlining the hydrogen economy within smart grids

Table 2.1 Examples of hydrogen-incorporated power systems around the world Project name Haeolus project

Location Denmark

SmartPowerFlow project

Germany

HyEnergy project

Netherlands

H2Future project

Austria

Hydrogen link project

Japan

Objectives and description Demonstrates large-scale hydrogen production from wind energy, integrating it into the energy system. Uses surplus wind power for electrolysis, injecting hydrogen into the gas grid for vehicle fuel and grid flexibility [81, 82]. Develops a hydrogen-based energy storage system integrated into a smart grid. Combines electrolysis for hydrogen production with fuel cells for storage and conversion. Hydrogen is used for power and heat generation, as well as hydrogen vehicles [83]. Integrates hydrogen into the natural gas grid for energy storage and transportation. Converts surplus renewable electricity to hydrogen via electrolysis, injecting it into the gas grid. The stored hydrogen is versatile for heat, transportation, or reconversion to electricity via fuel cells [84]. Focuses on large-scale hydrogen production using electrolysis powered by excess renewable energy. The produced hydrogen is intended for industrial applications such as steel production, grid stabilization, and mobility solutions [85, 86]. Demonstrates hydrogen and fuel cell integration into a smart grid. Includes hydrogen production, fuel cell power generation, and refueling stations. Showcases hydrogen’s potential for clean energy in power generation, transportation, and residential uses [87].

40

2.4.1

M. M. Hayati et al.

Haeolus Project (Denmark)

The Haeolus project aims to demonstrate the feasibility of large-scale hydrogen production from wind energy and its integration into the energy system. Located in Denmark, the project utilizes surplus wind power to produce hydrogen through electrolysis. The produced hydrogen is then injected into the natural gas grid and used for various purposes, such as fueling hydrogen vehicles and providing flexibility to the grid [81, 82].

2.4.2

SmartPowerFlow Project (Germany)

The SmartPowerFlow project, led by Siemens, focuses on the development of a hydrogen-based energy storage system integrated into a smart grid. The project combines hydrogen production via electrolysis with fuel cells for energy conversion and storage. The stored hydrogen can be used for heat and power generation, as well as for fueling hydrogen vehicles [83].

2.4.3

HyEnergy Project (Netherlands)

The HyEnergy project in the Netherlands explores the integration of hydrogen into the existing natural gas grid for energy storage and transportation. The project aims to convert surplus renewable electricity into hydrogen through electrolysis, which is then injected into the natural gas grid. The stored hydrogen can be utilized for heat production, transportation, or converted back into electricity through fuel cells [84].

2.4.4

H2Future Project (Austria)

The H2Future project is a collaborative effort involving industry partners, research institutions, and energy companies. Located in Austria, the project focuses on largescale hydrogen production using electrolysis, powered by excess renewable energy from wind and hydropower. The produced hydrogen is intended for industrial applications, such as steel production, as well as for grid stabilization and mobility [85, 86].

2

Hydrogen-Incorporated Sector-Coupled Smart Grids: A Systematic Review. . .

2.4.5

41

Hydrogen Link Project (Japan)

The Hydrogen Link project is being carried out in Japan and aims to demonstrate the integration of hydrogen and fuel cell technologies into a smart grid. The project involves the installation of a hydrogen production and supply system, fuel cell power generation units, and hydrogen refueling stations. It aims to showcase the potential of hydrogen as a clean energy carrier for power generation, transportation, and residential applications [87].

2.5

Technological Advancements and Innovations

2.5.1

Emerging Technologies for Hydrogen Production, Storage, and Distribution

As the world continues to transition toward a sustainable and low-carbon future, hydrogen has emerged as a promising clean energy carrier. However, the widespread utilization of hydrogen as a fuel source requires advancements in production, storage, and distribution technologies. In recent years, there have been significant developments in emerging technologies for hydrogen production, storage, and distribution, aimed at improving efficiency, reducing costs, and enhancing the overall viability of hydrogen as an energy solution. These advancements include innovative electrolysis technologies, such as high-temperature and low-temperature electrolysis, as well as novel approaches such as biomass gasification and photoelectrochemical water splitting. Moreover, hydrogen storage options, such as liquid organic hydrogen carriers (LOHC), metal hydrides, and nanostructured materials, are being explored to provide safe and compact storage solutions. Additionally, the establishment of hydrogen infrastructure, including hydrogen refueling stations and networks, is crucial for the seamless integration of hydrogen into existing energy systems. Through ongoing research and development efforts, these emerging technologies hold the potential to unlock the full benefits of hydrogen as a clean and sustainable energy source, enabling its integration into smart grids and the broader energy landscape. Consequently, related emerging technologies are as illustrated in Fig. 2.5.

2.5.1.1

Advanced Electrolysis Technologies

2.5.1.2

High-Temperature Electrolysis (HTE)

HTE involves electrolyzing steam at temperatures above 800 °C using solid oxide electrolysis cells (SOECs). This technology allows for more efficient hydrogen production and enables the utilization of waste heat from industrial processes or concentrated solar power.

42

M. M. Hayati et al.

Advanced Electrolysis Technologies

Renewable Hydrogen

High-Temperature Electrolysis

Liquid Organic Hydrogen Carriers

Hydrogen Pipelines Liquid Hydrogen Carriers

Metal Hydrides Biomass Gasification

Low-Temperature Electrolysis

Bipolar Membrane Electrolysis

Nanostructured Materials

Hydrogen Refueling Stations

Photoelectrochemical (PEC) Water Splitting

Hydrogen Hubs and Networks

Hydrogen Distribution

Hydrogen Infrastructure

Fig. 2.5 Emerging field of HiSGs

2.5.1.3

Low-Temperature Electrolysis (LTE)

LTE includes advancements in proton exchange membrane (PEM) electrolysis, which operates at lower temperatures and pressures. These systems are suitable for decentralized hydrogen production, such as small-scale applications or integration with renewable energy sources.

2.5.1.4

Bipolar Membrane Electrolysis

Bipolar membrane electrolysis utilizes an additional membrane that splits water into hydrogen and oxygen without the need for an external power supply. This technology has the potential to reduce energy consumption and costs in hydrogen production.

2.5.2

Renewable Hydrogen

2.5.2.1

Biomass Gasification

Biomass gasification involves the conversion of organic materials into hydrogenrich synthesis gas (syngas) through a thermochemical process. Syngas can be further processed to extract hydrogen, making it a renewable source of hydrogen production.

2

Hydrogen-Incorporated Sector-Coupled Smart Grids: A Systematic Review. . .

2.5.2.2

43

Photoelectrochemical (PEC) Water Splitting

PEC water splitting utilizes specialized semiconductors to directly convert solar energy into hydrogen through a photoelectrochemical process. Researchers are exploring various materials, such as metal oxides and organic compounds, to enhance the efficiency and stability of PEC systems.

2.5.3

Hydrogen Storage

2.5.3.1

Liquid Organic Hydrogen Carriers (LOHCs)

LOHCs are liquid molecules that can reversibly bind and release hydrogen. This technology allows for safe and compact hydrogen storage and transportation, as hydrogen can be chemically stored in the liquid carrier. Catalysts are used to release hydrogen from the carrier when needed.

2.5.3.2

Metal Hydrides

Metal hydrides, such as complex alloys or intermetallic compounds, can store hydrogen through reversible absorption and desorption reactions. Ongoing research focuses on improving the hydrogen storage capacity, kinetics, and operating conditions of metal hydride systems.

2.5.3.3

Nanostructured Materials

Nanostructured materials, including carbon nanotubes, metal-organic frameworks (MOFs), and porous materials, offer high surface area and tunable properties for enhanced hydrogen storage. These materials can adsorb and release hydrogen more efficiently, enabling compact and lightweight storage options.

2.5.4

Hydrogen Distribution

2.5.4.1

Hydrogen Pipelines

Hydrogen pipelines are being developed and optimized for the safe and efficient transportation of hydrogen over long distances. Specialized materials and corrosionresistant coatings are used to maintain the integrity of the pipelines and prevent hydrogen embrittlement.

44

2.5.4.2

M. M. Hayati et al.

Liquid Hydrogen Carriers

Liquid hydrogen carriers, similar to liquefied natural gas (LNG) carriers, are being explored as a means of transporting hydrogen. Liquid hydrogen has a higher energy density compared to compressed hydrogen gas, allowing for longer transportation distances.

2.5.5

Hydrogen Infrastructure

2.5.5.1

Hydrogen Refueling Stations

Advancements in hydrogen refueling station technologies include fast-filling stations with improved safety features, self-service options, and compatibility with different types of fuel cell vehicles. These developments aim to enhance the convenience and accessibility of hydrogen refueling infrastructure.

2.5.5.2

Hydrogen Hubs and Networks

Hydrogen hubs are regional centers that integrate hydrogen production, storage, and distribution facilities, creating a network for efficient and reliable hydrogen supply. These hubs can support various applications, including transportation, industry, and power generation, promoting the growth of hydrogen economies.

2.5.6

Smart Grid Control and Management Systems for Efficient Hydrogen Utilization

Smart grid control and management systems play a crucial role in enabling the efficient utilization of hydrogen within energy systems [88]. Also, the mentioned technologies can contribute with intelligent control methodologies to have better response in the fields such as industrial sectors [88, 89]. These systems facilitate the integration of hydrogen production, storage, and distribution into the grid, optimizing its deployment and ensuring seamless operation. Therefore, key aspects of smart grid control and management systems for efficient hydrogen utilization are included in Fig. 2.6.

2.5.6.1

Demand Response and Load Management

Smart grid control systems enable demand-response programs, where consumers are incentivized to adjust their energy usage based on supply and demand conditions. By

2

Hydrogen-Incorporated Sector-Coupled Smart Grids: A Systematic Review. . .

45

Demand Response

Grid Integration

Management System Smart Grid Control & Management

Decentralized Control

Intelligent Monitoring

Fig. 2.6 Management Technologies of HiSGs

integrating hydrogen utilization into demand-response strategies, energy consumption patterns can be optimized to align with the availability of hydrogen and renewable energy. This improves the overall efficiency of the grid and enhances the utilization of hydrogen resources.

2.5.6.2

Energy Management Systems

Advanced energy management systems use real-time data and analytics to monitor and control the production, storage, and distribution of hydrogen within a smart grid. These systems optimize the allocation and utilization of hydrogen resources based on various factors such as electricity prices, demand forecasts, and renewable energy availability. By dynamically managing hydrogen utilization, energy management systems ensure efficient and reliable operation of the grid.

2.5.6.3

Grid Integration and Power Balancing

Smart grid control systems enable the seamless integration of hydrogen-based energy systems with the existing electrical grid. By balancing the supply and demand of electricity and hydrogen, these systems ensure stable and reliable

46

M. M. Hayati et al.

operation while maximizing the utilization of renewable resources and hydrogen energy storage. Advanced control algorithms and predictive modeling techniques are used to optimize power flow and maintain grid stability.

2.5.6.4

Intelligent Monitoring and Predictive Maintenance

Smart grid management systems employ artificial intelligence (AI) and machine learning algorithms to monitor the performance of hydrogen production, storage, and distribution infrastructure. By continuously analyzing sensor data, these systems can detect anomalies, predict equipment failures, and optimize maintenance schedules. Intelligent monitoring and predictive maintenance ensure the efficient operation of the system, minimize downtime, and enhance the lifespan of hydrogen infrastructure.

2.5.6.5

Decentralized Control and Peer-to-Peer (P2P) Energy Trading

Smart grid systems facilitate decentralized control and peer-to-peer energy trading, allowing prosumers (consumers who also produce energy) to actively participate in the energy market. By integrating hydrogen utilization into decentralized control mechanisms, locally produced hydrogen can be efficiently distributed and utilized based on local energy demand and supply. Peer-to-peer energy trading enables the exchange of excess renewable energy and hydrogen between users, promoting grid resilience, flexibility, and efficient resource allocation [90, 91]. Blockchain technology is significantly transforming P2P electrical energy trading by introducing decentralized and transparent transactions. Through smart contracts, real-time settlements, and traceable energy provenance, blockchain ensures trust and accountability among participants. Tokenization enables fractional ownership and trading of energy, reducing costs and promoting accessibility. Moreover, P2P trading on the blockchain enhances grid integration, security, and resilience while empowering smaller energy producers. Overall, blockchain’s impact on P2P electrical energy trading encompasses efficient, transparent, and environmentally friendly transactions, revolutionizing the energy landscape.

2.6 2.6.1

Economic and Environmental Considerations Cost Analysis of Hydrogen-Integrated Smart Grid Systems

To assess the viability of hydrogen-integrated smart grid systems, a comprehensive cost analysis is essential. This involves evaluating the costs associated with

2

Hydrogen-Incorporated Sector-Coupled Smart Grids: A Systematic Review. . .

47

hydrogen production, storage, distribution, and integration into the grid. Factors considered in the cost analysis include capital investments, operation and maintenance expenses, energy losses, and infrastructure development. The cost analysis takes into account the specific technologies employed such as electrolysis systems, storage methods, and control systems. It also considers the scale of implementation, as larger deployments often benefit from economies of scale. Additionally, the cost analysis compares hydrogen-integrated smart grid systems with alternative energy storage options to determine their cost competitiveness. Advancements in technologies, economies of scale, and supportive policies are expected to contribute to cost reductions over time, making hydrogen-integrated smart grid systems more economically attractive. However, cost considerations should be balanced with the potential benefits such as grid stability, renewable energy integration, and reduced greenhouse gas emissions.

2.6.2

Evaluation of the Environmental Impacts and Sustainability Aspects

Hydrogen-integrated smart grid systems have the potential to significantly reduce environmental impacts compared to traditional energy systems. Evaluating these impacts involves assessing factors such as greenhouse gas emissions, air pollution, land use, and water consumption throughout the lifecycle of hydrogen production, storage, distribution, and utilization. Life cycle assessments (LCAs) are used to quantify the environmental impacts associated with different stages of the hydrogen supply chain. LCAs consider the energy sources used for hydrogen production, the efficiency of conversion technologies, and the emissions associated with each step. Additionally, the environmental impacts of hydrogen storage materials, transportation methods, and end-use applications are evaluated. Sustainability aspects go beyond environmental considerations and encompass economic and social dimensions. This includes analyzing the long-term availability of hydrogen feedstocks, the impact on local communities, and the potential for job creation and economic growth. To ensure the sustainable development of hydrogenintegrated smart grid systems, efforts are made to minimize environmental impacts through technology advancements, renewable energy integration, and adherence to regulatory standards. Additionally, strategies for recycling, repurposing, or responsibly disposing of hydrogen infrastructure components and materials are explored.

48

M. M. Hayati et al.

Fig. 2.7 Economic incentives of HiSGs

Carbon Pricing

Economic Incentives

R&D Funding

Financial Mechanism

2.6.3

Economic and Financial Incentives for Promoting Hydrogen Integration

Promoting the integration of hydrogen into smart grids often requires economic and financial incentives to overcome initial investment costs and drive market adoption. These incentives can take various forms, which is presented in Fig. 2.7.

2.6.3.1

Carbon Pricing and Emissions Trading

Implementing carbon pricing mechanisms, such as carbon taxes or cap-and-trade systems, can create economic incentives for reducing greenhouse gas emissions. This can indirectly promote the adoption of hydrogen-integrated smart grid systems by making hydrogen a cost-competitive option compared to fossil fuels.

2.6.3.2

Research and Development Funding

Entities can invest in research and development activities to advance hydrogen technologies and reduce costs. Funding research projects, collaborations, and innovation centers can accelerate the commercialization of new technologies and drive down costs through technological advancements.

2.6.3.3

Financing Mechanisms

Access to financing options, such as low-interest loans, guarantees, and venture capital investments, can help overcome financial barriers and facilitate the deployment of hydrogen-integrated smart grid systems. Financial institutions can play a role in developing tailored financial instruments that address the specific needs and risks associated with hydrogen projects.

2

Hydrogen-Incorporated Sector-Coupled Smart Grids: A Systematic Review. . .

49

By providing economic and financial incentives, stakeholders aim to accelerate the adoption of hydrogen-integrated smart grid systems, facilitate the transition to a low-carbon economy, and promote sustainable energy solutions.

2.7

Future Prospects and Research Directions

2.7.1

Identification of Key Challenges and Gaps in Current Hydrogen-Integrated Smart Grid Systems

While hydrogen-integrated smart grid systems show great promise, several challenges and gaps need to be addressed for their widespread adoption. Figure 2.8 includes the following:

2.7.1.1

Cost-Effectiveness

The high capital costs associated with hydrogen production, storage, and distribution infrastructure pose a significant challenge. Continued research and development efforts are needed to reduce costs and improve the cost-effectiveness of hydrogen technologies.

Fig. 2.8 Future prospects of HiSGs Cost-Effectivness

Scalability and Infrastructure Development

Future Prospective

Safety Considerations

Technology Efficiency and Reliability

50

2.7.1.2

M. M. Hayati et al.

Scalability and Infrastructure Development

Scaling up hydrogen production and distribution infrastructure requires substantial investments and careful planning. Developing an extensive hydrogen infrastructure network, including production facilities, refueling stations, and pipelines, remains a challenge that needs to be addressed.

2.7.1.3

Technology Efficiency and Reliability

Improving the efficiency and reliability of hydrogen production, storage, and utilization technologies is crucial for their integration into smart grids. Further advancements are needed to enhance the performance, durability, and lifespan of electrolysis systems, fuel cells, and hydrogen storage materials.

2.7.1.4

Safety Considerations

Safety is a critical aspect of hydrogen utilization. Proper safety measures, codes, and standards must be established to ensure the safe production, storage, and handling of hydrogen. Public perception and acceptance of hydrogen as a safe energy carrier also need to be addressed.

2.7.2

Exploration of Future Prospects and Potential Advancements

The future prospects of hydrogen-integrated smart grid systems are promising, with several potential advancements on the horizon:

2.7.2.1

Advanced Electrolysis Technologies

Continued advancements in high-temperature and low-temperature electrolysis technologies can significantly improve the efficiency and cost-effectiveness of hydrogen production. Innovations in catalyst materials and system designs can enhance performance and durability.

2.7.2.2

Renewable Hydrogen Sources

Research into new and sustainable sources of renewable hydrogen, such as biohydrogen, solar-driven water splitting, and wind-to-hydrogen systems, can

2

Hydrogen-Incorporated Sector-Coupled Smart Grids: A Systematic Review. . .

51

further diversify the feedstock options and increase the sustainability of hydrogen production.

2.7.2.3

Energy Storage and Conversion

Advancements in hydrogen storage materials, such as novel metal hydrides or advanced adsorbents, can enhance storage capacity and kinetics. Further research is also needed to improve the efficiency and durability of fuel cells for various applications, including transportation and power generation.

2.7.2.4

Hydrogen Grid Integration

Integrating hydrogen into existing electricity grids requires the development of advanced control and management systems. This includes the optimization of hydrogen dispatch strategies, grid balancing mechanisms, and demand-response programs to ensure efficient and reliable integration of hydrogen into smart grids.

2.7.3

Research Directions and Areas for Further Investigation

To unlock the full potential of hydrogen-integrated smart grid systems, further research is needed in the following areas:

2.7.3.1

Techno-Economic Analysis

Conducting detailed techno-economic analyses can provide valuable insights into the cost competitiveness and viability of hydrogen-integrated smart grid systems. This analysis should consider various deployment scenarios, scale effects, and the dynamic interaction between hydrogen, electricity, and other energy carriers.

2.7.3.2

Lifecycle Assessments and Sustainability

Comprehensive lifecycle assessments are essential to evaluate the environmental impacts and sustainability of hydrogen-integrated smart grid systems. This includes considering the entire supply chain, from feedstock extraction to end-use applications, and assessing the potential for emissions reduction and resource efficiency.

52

M. M. Hayati et al.

2.7.3.3

System Modeling and Optimization

Developing advanced modeling and optimization tools can support decision-making processes and enable the efficient planning, operation, and control of hydrogenintegrated smart grid systems. This involves considering factors such as energy demand, renewable energy generation profiles, storage capacities, and grid stability requirements.

2.7.3.4

Demonstration Projects and Real-World Applications

Implementing large-scale demonstration projects and real-world applications can provide valuable insights and validate the performance and feasibility of hydrogenintegrated smart grid systems. These projects can help overcome technical, economic, and social barriers and facilitate the adoption of hydrogen technologies. By focusing on these research directions and addressing the key challenges, the future of hydrogen-integrated smart grid systems can be paved with sustainable, efficient, and cost-effective solutions contribute to a clean sustainable electrical energy future.

2.8

Conclusion

A comprehensive review, performed in this chapter, highlights the emerging technologies, economic considerations, and environmental implications of integrating hydrogen into smart grid systems. The findings reveal that advanced electrolysis technologies, renewable hydrogen sources, and innovative storage methods hold promise for efficient and sustainable hydrogen production, storage, and distribution. Moreover, the integration of hydrogen in sector-coupled smart grids has significant implications for renewable energy integration, grid flexibility, decarbonization, and energy security. However, challenges related to cost-effectiveness, infrastructure development, and safety must be addressed to realize the full potential of hydrogen integration. The recommendations for future endeavors include continued research and development, collaborative efforts, supportive policies, and demonstration projects to overcome these challenges and promote the widespread adoption of hydrogen-integrated smart grid systems. By pursuing these avenues, the transition to a sustainable and low-carbon energy future can be accelerated, contributing to global climate change mitigation and energy sector transformation.

2

Hydrogen-Incorporated Sector-Coupled Smart Grids: A Systematic Review. . .

53

References 1. Mahmood, D., Javaid, N., Ahmed, G., Khan, S., & Monteiro, V. (2021). A review on optimization strategies integrating renewable energy sources focusing uncertainty factor–paving path to eco-friendly smart cities. Sustainable Computing: Informatics and Systems, 30, 100559. 2. Mohseni, S., & Brent, A. C. (2022). Quantifying the effects of forecast uncertainty on the role of different battery technologies in grid-connected solar photovoltaic/wind/micro-hydro microgrids: An optimal planning study. Journal of Energy Storage, 51, 104412. 3. Aminlou, A., Hayati, M. M., & Zare, K. (2023). P2P energy trading in a community of individual consumers with the presence of central shared battery energy storage. In V. Vahidinasab & B. Mohammadi-Ivatloo (Eds.), Demand-side peer-to-peer energy trading (pp. 143–159). Springer International Publishing. 4. Ishaq, H., Dincer, I., & Crawford, C. (2022). A review on hydrogen production and utilization: Challenges and opportunities. International Journal of Hydrogen Energy, 47(62), 26238–26264. 5. Hassan, Q., Sameen, A. Z., Olapade, O., Alghoul, M., Salman, H. M., & Jaszczur, M. (2023). Hydrogen fuel as an important element of the energy storage needs for future smart cities. International Journal of Hydrogen Energy. 6. Raghavaiah, N., & Srinivasulu, G. N. (2022). Fuel cells for alternative and sustainable energy systems. Renewable Energy for Sustainable Growth Assessment, 363–388. 7. Al-Mufachi, N. A., & Shah, N. (2022). The role of hydrogen and fuel cell technology in providing security for the UK energy system. Energy Policy, 171, 113286. 8. Kaur, M., & Pal, K. (2019). Review on hydrogen storage materials and methods from an electrochemical viewpoint. Journal of Energy Storage, 23, 234–249. 9. Tao, Y., Qiu, J., Lai, S., & Zhao, J. (2020). Integrated electricity and hydrogen energy sharing in coupled energy systems. IEEE Transactions on Smart Grid, 12(2), 1149–1162. 10. Sun, X., Cao, X., Zeng, B., Zhai, Q., & Guan, X. (2022). Multistage dynamic planning of integrated hydrogen-electrical microgrids under multiscale uncertainties. IEEE Transactions on Smart Grid. 11. Tao, Y., Qiu, J., Lai, S., & Sun, X. (2022). Coordinated planning of electricity and hydrogen networks with hydrogen supply chain for fuel cell electric vehicles. IEEE Transactions on Sustainable Energy. 12. Zheng, J., et al. (2021). Current research trends and perspectives on solid-state nanomaterials in hydrogen storage. Research, 2021. 13. Wan, L., Zhang, W., & Xu, Z. (2020). Overview of key technologies and applications of hydrogen energy storage in integrated energy systems. In 2020 12th IEEE PES Asia-pacific power and energy engineering conference (APPEEC) (pp. 1–5). IEEE. 14. Arsad, A., et al. (2022). Hydrogen energy storage integrated hybrid renewable energy systems: A review analysis for future research directions. International Journal of Hydrogen Energy, 47(39), 17285–17312. 15. Undertaking, H. J. (2019). Hydrogen roadmap Europe: A sustainable pathway for the European energy transition. Sustainability, 13(24), 13681. 16. He, W., et al. (2020). Integration of renewable hydrogen in light-duty vehicle: Nexus between energy security and low carbon emission resources. International Journal of Hydrogen Energy, 45(51), 27958–27968. 17. Wang, Q., Xue, M., Lin, B.-L., Lei, Z., & Zhang, Z. (2020). Well-to-wheel analysis of energy consumption, greenhouse gas and air pollutants emissions of hydrogen fuel cell vehicle in China. Journal of Cleaner Production, 275, 123061. 18. Greene, D. L., Ogden, J. M., & Lin, Z. (2020). Challenges in the designing, planning and deployment of hydrogen refueling infrastructure for fuel cell electric vehicles. ETransportation, 6, 100086.

54

M. M. Hayati et al.

19. Olabi, A., et al. (2021). Large-scale hydrogen production and storage technologies: Current status and future directions. International Journal of Hydrogen Energy, 46(45), 23498–23528. 20. Rath, R., Kumar, P., Mohanty, S., & Nayak, S. K. (2019). Recent advances, unsolved deficiencies, and future perspectives of hydrogen fuel cells in transportation and portable sectors. International Journal of Energy Research, 43(15), 8931–8955. 21. Anvari-Moghaddam, A., Mohammadi-Ivatloo, B., Asadi, S., & Shahidehpour, M. (2022). Sustainable energy systems planning, integration and management (Volume II), vol. 12, ed: MDPI (pp. 10914). 22. Trencher, G. (2020). Strategies to accelerate the production and diffusion of fuel cell electric vehicles: Experiences from California. Energy Reports, 6, 2503–2519. 23. Habibi, M., Vahidinasab, V., Mohammadi-Ivatloo, B., Aghaei, J., & Taylor, P. (2022). Exploring potential gains of mobile sector-coupling energy systems in heavily constrained networks. IEEE Transactions on Sustainable Energy, 13(4), 2092–2105. 24. Sadeghian, O., Oshnoei, A., Mohammadi-Ivatloo, B., & Vahidinasab, V. (2022). Concept, definition, enabling technologies, and challenges of energy integration in whole energy systems to create integrated energy systems. In Whole energy systems: Bridging the gap via vectorcoupling technologies (pp. 1–21). Springer. 25. Daneshvar, M., Mohammadi-Ivatloo, B., & Zare, K. (2022). A novel transactive energy trading model for modernizing energy hubs in the coupled heat and electricity network. Journal of Cleaner Production, 344, 131024. 26. Van Nuffel, L. (2020). Study sector coupling: How can it be enhanced in the EU to foster grid stability and decarbonise? Policy department for economic, scientific and quality of life policies, directorate-general for internal policies. 27. Singh, R., Akram, S. V., Gehlot, A., Buddhi, D., Priyadarshi, N., & Twala, B. (2022). Energy system 4.0: Digitalization of the energy sector with inclination towards sustainability. Sensors, 22(17), 6619. 28. Wu, Y., Wu, Y., Guerrero, J. M., & Vasquez, J. C. (2022). Decentralized transactive energy community in edge grid with positive buildings and interactive electric vehicles. International Journal of Electrical Power & Energy Systems, 135, 107510. 29. Nazari-Heris, M., Abapour, M., & Mohammadi-Ivatloo, B. (2022). An updated review and outlook on electric vehicle aggregators in electric energy networks. Sustainability, 14(23), 15747. 30. Khalid, M., Ahmad, F., Panigrahi, B. K., & Al-Fagih, L. (2022). A comprehensive review on advanced charging topologies and methodologies for electric vehicle battery. Journal of Energy Storage, 53, 105084. 31. Ifaei, P., Nazari-Heris, M., Charmchi, A. S. T., Asadi, S., & Yoo, C. (2022). Sustainable energies and machine learning: An organized review of recent applications and challenges. Energy, 126432. 32. Sorrenti, I., Rasmussen, T. B. H., You, S., & Wu, Q. (2022). The role of power-to-X in hybrid renewable energy systems: A comprehensive review. Renewable and Sustainable Energy Reviews, 112380. 33. Incer-Valverde, J., Patiño-Arévalo, L. J., Tsatsaronis, G., & Morosuk, T. (2022). Hydrogendriven power-to-X: State of the art and multicriteria evaluation of a study case. Energy Conversion and Management, 266, 115814. 34. Osorio-Aravena, J. C., et al. (2023). Synergies of electrical and sectoral integration: Analysing geographical multi-node scenarios with sector coupling variations for a transition towards a fully renewables-based energy system. Energy, 128038. 35. Wu, Y., Wu, Y., Cimen, H., Vasquez, J. C., & Guerrero, J. M. (2022). Towards collective energy community: Potential roles of microgrid and blockchain to go beyond P2P energy trading. Applied Energy, 314, 119003. 36. Feijoo, F., Pfeifer, A., Herc, L., Groppi, D., & Duić, N. (2022). A long-term capacity investment and operational energy planning model with power-to-X and flexibility technologies. Renewable and Sustainable Energy Reviews, 167, 112781.

2

Hydrogen-Incorporated Sector-Coupled Smart Grids: A Systematic Review. . .

55

37. Skordoulias, N., Koytsoumpa, E. I., & Karellas, S. (2022). Techno-economic evaluation of medium scale power to hydrogen to combined heat and power generation systems. International Journal of Hydrogen Energy, 47(63), 26871–26890. 38. Lamers, P., Ghosh, T., Upasani, S., Sacchi, R., & Daioglou, V. (2023). Linking life cycle and integrated assessment modeling to evaluate technologies in an evolving system context: A power-to-hydrogen case study for the United States. Environmental Science & Technology, 57(6), 2464–2473. 39. Schütte, C. et al. (2022). Decarbonization of the metal industry in Hamburg–Demand, efficiency and costs of green hydrogen. In 2022 18th international conference on the european energy market (EEM) (pp. 1–8). IEEE. 40. Norouzi, S., Dadashi, M., Haghifam, S., & Zare, K. (2022). Overview of modern multidimension energy networks. In Coordinated operation and planning of modern heat and electricity incorporated networks (pp. 31–55). 41. Yilmaz, H. Ü., Kimbrough, S. O., van Dinther, C., & Keles, D. (2022). Power-to-gas: Decarbonization of the European electricity system with synthetic methane. Applied Energy, 323, 119538. 42. Genovese, M., Schlüter, A., Scionti, E., Piraino, F., Corigliano, O., & Fragiacomo, P. (2023). Power-to-hydrogen and hydrogen-to-X energy systems for the industry of the future in Europe. International Journal of Hydrogen Energy. 43. Mendoza, J. M. F., & Ibarra, D. (2023). Technology-enabled circular business models for the hybridisation of wind farms: Integrated wind and solar energy, power-to-gas and power-toliquid systems. In Sustainable production and consumption. 44. da Silva Pinto, R. L., et al. (2022). An overview on the production of synthetic fuels from biogas. Bioresource Technology Reports, 18, 101104. 45. Allouhi, A. (2022). Techno-economic and environmental accounting analyses of an innovative power-to-heat concept based on solar PV systems and a geothermal heat pump. Renewable Energy, 191, 649–661. 46. Pastore, L. M., Basso, G. L., Ricciardi, G., & de Santoli, L. (2022). Synergies between powerto-heat and power-to-gas in renewable energy communities. Renewable Energy, 198, 1383–1397. 47. Forndal, L., & Greiff, J. (2022). System study of the techno-economic potential of a hydrogen system: A case study of power to mobility and power to power hydrogen systems, stand-alone or integrated with a CHP, stand-alone or integrated with a CHP, Master Dissertation, Linköping University, Department of Management and Engineering, Energy Systems, https://www.divaportal.org/smash/record.jsf?pid=diva2%3A1666533&dswid=-454 48. Youssef, A.-R., Mallah, M., Ali, A., Shaaban, M. F., & Mohamed, E. E. (2023). Enhancement of microgrid frequency stability based on the combined power-to-hydrogen-to-power technology under high penetration renewable units. Energies, 16(8), 3377. 49. Khaligh, V., Ghezelbash, A., Mazidi, M., Liu, J., Ryu, J.-H., & Na, J. (2022). A stochastic agent-based cooperative scheduling model of a multi-vector microgrid including electricity, hydrogen, and gas sectors. Journal of Power Sources, 546, 231989. 50. Khaligh, V., Ghasemnejad, M. K., Ghezelbash, A., Liu, J., & Won, W. (2023). Risk-constrained energy management of an isolated multi-energy microgrid enhanced with hydrogen storage. Journal of Energy Storage, 63, 107103. 51. Ahmadi, S. E., Sadeghi, D., Marzband, M., Abusorrah, A., & Sedraoui, K. (2022). Decentralized bi-level stochastic optimization approach for multi-agent multi-energy networked micro-grids with multi-energy storage technologies. Energy, 245, 123223. 52. Bridgeland, R., Chapman, A., McLellan, B., Sofronis, P., & Fujii, Y. (2022). Challenges toward achieving a successful hydrogen economy in the US: Potential end-use and infrastructure analysis to the year 2100. Cleaner Production Letters, 3, 100012. 53. Pramuanjaroenkij, A., & Kakaç, S. (2023). The fuel cell electric vehicles: The highlight review. International Journal of Hydrogen Energy, 48(25), 9401–9425.

56

M. M. Hayati et al.

54. Zheng, Y., et al. (2022). Nanotube-based heterostructures for electrochemistry: A mini-review on lithium storage, hydrogen evolution and beyond. Journal of Energy Chemistry, 70, 630–642. 55. Oskouei, M. Z., et al. (2022). A critical review on the impacts of energy storage systems and demand-side management strategies in the economic operation of renewable-based distribution network. Sustainability, 14(4), 2110. 56. Jabari, F., Nazari-heris, M., & Abapour, M. (2023). Implementation and investigation of demand-side management polices in Iran’s industrial and commercial sectors. Journal of Energy Management and Technology, 7(1), 34–42. 57. Niazzadeh, A., Azad, S., Taghi Ameli, M., Nazari-Heris, M., & Asadi, S. (2022). A survey on home energy management systems with viewpoints of concepts, configurations, and infrastructures. In Renewable energy for buildings: Technology, control, and operational techniques (pp. 61–76). Springer. 58. Goulden, M., Bedwell, B., Rennick-Egglestone, S., Rodden, T., & Spence, A. (2014). Smart grids, smart users? The role of the user in demand side management. Energy Research & Social Science, 2, 21–29. 59. Loni, A., Asadi, S., & Nazari-Heris, M. (2023). A peer-to-peer reputation-based mechanism to enhance microgrids’ power exchange quality. In 2023 IEEE texas power and energy conference (TPEC) (pp. 1–6). 60. Nazari-Heris, M., Loni, A., Asadi, S., & Mohammadi-ivatloo, B. (2022). Toward social equity access and mobile charging stations for electric vehicles: A case study in Los Angeles. Applied Energy, 311, 118704. 61. Ghahramani, M., Nazari-Heris, M., Zare, K., & Mohammadi-Ivatloo, B. (2022). A two-point estimate approach for energy management of multi-carrier energy systems incorporating demand response programs. Energy, 249, 123671. 62. Abomazid, A. M., El-Taweel, N. A., & Farag, H. E. (2022). Optimal energy management of hydrogen energy facility using integrated battery energy storage and solar photovoltaic systems. IEEE Transactions on Sustainable Energy, 13(3), 1457–1468. 63. Babaei, R., Ting, D. S., & Carriveau, R. (2022). Optimization of hydrogen-producing sustainable Island microgrids. International Journal of Hydrogen Energy, 47(32), 14375–14392. 64. Lin, R.-H., Zhao, Y.-Y., & Wu, B.-D. (2020). Toward a hydrogen society: Hydrogen and smart grid integration. International Journal of Hydrogen Energy, 45(39), 20164–20175. 65. Zhou, Y., & Lund, P. D. (2023). Peer-to-peer energy sharing and trading of renewable energy in smart communities – trading pricing models, decision-making and agent-based collaboration. Renewable Energy. 66. Hayati, M. M., Aminlou, A., Zare,K., & Abapour, M. (2023). A two-stage stochastic optimization scheduling approach for integrating renewable energy sources and deferrable demand in the spinning reserve market. In 2023 8th international conference on technology and energy management (ICTEM) (pp. 1–7). IEEE. 67. Mansour-Saatloo, A., Mirzaei, M. A., Mohammadi-Ivatloo, B., & Zare, K. (2020). A riskaverse hybrid approach for optimal participation of power-to-hydrogen technology-based multienergy microgrid in multi-energy markets. Sustainable Cities and Society, 63, 102421. 68. Mehdinejad, M., Shayanfar, H., & Mohammadi-Ivatloo, B. (2022). Peer-to-peer decentralized energy trading framework for retailers and prosumers. Applied Energy, 308, 118310. 69. Xiao, Y., Wang, X., Pinson, P., & Wang, X. (2017). A local energy market for electricity and hydrogen. IEEE Transactions on Power Systems, 33(4), 3898–3908. 70. Aminlou, A., Hayati, M. M., & Zare, K. (2023). local peer-to-peer energy trading evaluation in micro-grids with centralized approach. In 2023 8th international conference on technology and energy management (ICTEM) (pp. 1–6). 71. Nojavan, S., Zare, K., & Mohammadi-Ivatloo, B. (2017). Selling price determination by electricity retailer in the smart grid under demand side management in the presence of the electrolyser and fuel cell as hydrogen storage system. International Journal of Hydrogen Energy, 42(5), 3294–3308.

2

Hydrogen-Incorporated Sector-Coupled Smart Grids: A Systematic Review. . .

57

72. Nojavan, S., & Zare, K. (2018). Optimal energy pricing for consumers by electricity retailer. International Journal of Electrical Power & Energy Systems, 102, 401–412. 73. Agabalaye-Rahvar, M., Mansour-Saatloo, A., Mirzaei, M. A., Mohammadi-Ivatloo, B., & Zare, K. (2021). Economic-environmental stochastic scheduling for hydrogen storage-based smart energy hub coordinated with integrated demand response program. International Journal of Energy Research, 45(14), 20232–20257. 74. Aghdam, F. H., Mudiyanselage, M. W., Mohammadi-Ivatloo, B., & Marzband, M. (2023). Optimal scheduling of multi-energy type virtual energy storage system in reconfigurable distribution networks for congestion management. Applied Energy, 333, 120569. 75. Pelaez-Samaniego, M. R., Riveros-Godoy, G., Torres-Contreras, S., Garcia-Perez, T., & Albornoz-Vintimilla, E. (2014). Production and use of electrolytic hydrogen in Ecuador towards a low carbon economy. Energy, 64, 626–631. 76. Alanne, K., & Cao, S. (2017). Zero-energy hydrogen economy (ZEH2E) for buildings and communities including personal mobility. Renewable and Sustainable Energy Reviews, 71, 697–711. 77. Seyyedeh-Barhagh, S., Abapour, M., Mohammadi-ivatloo, B., & Shafie-khah, M. (2023). An overview of implementation of P2P energy trading methods on the electric power systems. In Trading in local energy markets and energy communities: Concepts, structures and technologies (pp. 137–149). 78. Alilou, M., Azami, H., Oshnoei, A., Mohammadi-Ivatloo, B., & Teodorescu, R. (2023). Fractional-order control techniques for renewable energy and energy-storage-integrated power systems: A review. Fractal and Fractional, 7(5), 391. 79. Hjeij, D., Biçer, Y., & Koc, M. (2022). Hydrogen strategy as an energy transition and economic transformation avenue for natural gas exporting countries: Qatar as a case study. International Journal of Hydrogen Energy, 47(8), 4977–5009. 80. Kabiri-Renani, Y., Daneshvar, M., & Mohammadi-Ivatloo, B. (2022). Transactive energy revolution: Innovative leverage for reliable operation of modern energy networks – A critical review. IET Renewable Power Generation, 16(15), 3368–3383. 81. Zenith, F., Flote, M. N., Santos-Mugica, M., Duncan, C. S., Mariani, V., & Marcantonini, C. (2022). Value of green hydrogen when curtailed to provide grid balancing services. International Journal of Hydrogen Energy, 47(84), 35541–35552. 82. Zemite, L., et al. (2023). A comprehensive overview of the Europen and Baltic landscape for hydrogen applications and innovations. Latvian Journal of Physics and Technical Sciences, 60(3), 33–53. 83. Resch, M., Bühler, J., Schachler, B., Kunert, R., Meier, A., & Sumper, A. (2019). Technical and economic comparison of grid supportive vanadium redox flow batteries for primary control reserve and community electricity storage in Germany. International Journal of Energy Research, 43(1), 337–357. 84. Shaikh, I. R. (2020). Chemistry for energy conversion and fossil free sustainable Enterprise. In Sustainable entrepreneurship, renewable energy-based projects, and digitalization (pp. 93–124). CRC Press. 85. Tang, J., Chu, M.-S., Li, F., Feng, C., Liu, Z.-G., & Zhou, Y.-S. (2020). Development and progress on hydrogen metallurgy. International Journal of Minerals, Metallurgy and Materials, 27, 713–723. 86. Zhang, X., Jiao, K., Zhang, J., & Guo, Z. (2021). A review on low carbon emissions projects of steel industry in the world. Journal of Cleaner Production, 306, 127259. 87. Zhao, F., et al. (2020). Hydrogen fuel cell vehicle development in China: An industry chain perspective. Energy Technology, 8(11), 2000179. 88. Safari, A., & Ghavifekr, A. A. (2022). Quantum technology & quantum neural networks in smart grids control: Premier perspectives. In 2022 8th international conference on control, instrumentation and automation (ICCIA) (pp. 1–6).

58

M. M. Hayati et al.

89. Ghavifekr, A. A., Velázquez, R., & Safari, A. (2021). Multirate adaptive inverse dynamics control of 5 DOF industrial gryphon robot. In 2021 9th RSI international conference on robotics and mechatronics (ICRoM) (pp. 255–260). 90. Aminlou, A., Hayati, M., & Zare, K. (2023). ADMM-based fully decentralized peer to peer energy trading considering a shared CAES in a local community. Journal of Energy Management and Technology. 91. Aminlou, A., Mohammadi-Ivatloo, B., Zare, K., Razzaghi, R., & Anvari-Moghaddam, A. (2022). Peer-to-peer decentralized energy trading in industrial town considering central shared energy storage using alternating direction method of multipliers algorithm. IET Renewable Power Generation, 16(12), 2579–2589.

Chapter 3

Techno-Economic Analysis for Centralized GH2 Power Systems Mohammad Mohsen Hayati , Behzad Motallebi Azar Mehdi Abapour , and Kazem Zare

3.1

, Ali Aminlou

,

Introduction

The adoption of hydrogen energy systems (HESs) is growing and developing at the fastest rate, and modern power grids are moving toward decentralization more and more [1]. The emergence of green hydrogen (GH2) technologies and the need to pay attention to renewable energy sources (RES) are the main causes of this energy transition process [2]. In the quest for a low-carbon future, the development and integration of sustainable energy systems have become critical imperatives. As the global community strives to diminish greenhouse gas emissions and mitigate the adverse impacts of climate change, the transition toward a low-carbon economy has gained significant momentum [3, 4]. One promising avenue for achieving this transition lies in the utilization of GH2 as a versatile and carbon-neutral energy carrier, which can facilitate the integration of RESs and address the challenges associated with their intermittent nature [5, 6]. The emergence of centralized green hydrogen power systems offers a promising solution to the challenges of renewable energy integration and energy storage [7, 8]. These systems, which encompass the production, storage, and distribution of green hydrogen, have the potential to revolutionize the energy landscape by providing a reliable and scalable platform for sustainable energy deployment. By leveraging RESs, such as wind and solar, to power the electrolysis process, GH2 can be produced without carbon emissions, thus

M. M. Hayati · A. Aminlou · M. Abapour · K. Zare (✉) Faculty of Electrical and Computer Engineering, Energy Systems Research Institute (ESRI), University of Tabriz, Tabriz, Iran e-mail: [email protected]; [email protected] B. Motallebi Azar Faculty of Electrical Engineering, Sahand University of Technology, Tabriz, Iran e-mail: [email protected] © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 V. Vahidinasab et al. (eds.), Green Hydrogen in Power Systems, Green Energy and Technology, https://doi.org/10.1007/978-3-031-52429-5_3

59

60

M. M. Hayati et al.

offering a pathway to decarbonizing sectors that are difficult to electrify directly such as heavy industry, transportation, and heating. In this context, conducting a comprehensive techno-economic analysis becomes crucial to evaluate the viability and potential benefits of centralized green hydrogen power systems. Such an analysis encompasses various aspects, including the economic feasibility, technological considerations, and environmental implications of adopting this approach. By examining the costs, benefits, and potential risks associated with the deployment of green hydrogen, decision-makers can gain valuable insights into the viability and long-term sustainability of this emerging energy paradigm. Furthermore, the integration of a peer-to-peer (P2P) energy market within the centralized green hydrogen power system holds immense potential [9]. P2P energy exchange allows prosumers (consumers and producers of energy) to exchange excess renewable energy directly, fostering greater market efficiency, grid flexibility, and localized energy resilience [10, 11]. By enabling the dynamic and efficient utilization of renewable energy resources, P2P energy trading can unlock additional economic value and facilitate the transition toward a more decentralized and democratized energy system [12]. Through this chapter, we aim to provide a comprehensive and scientifically grounded analysis of the techno-economic aspects of centralized green hydrogen power systems, with a specific focus on the potential role of P2P energy markets. By examining the economic viability, technological considerations, renewable energy integration, and energy storage aspects, we intend to shed light on the potential of centralized green hydrogen power systems as key catalysts for sustainable energy transitions. Ultimately, this analysis aims to inform policymakers, industry stakeholders, and researchers about the opportunities and challenges associated with adopting a centralized approach to green hydrogen power systems and its implications for a low-carbon future.

3.2

Energy Democracy in Energy Communities

Energy democracy is a concept that advocates for the decentralization and democratization of energy systems, giving individuals and communities more control and decision-making power over their energy sources and consumption [13]. In other words, it is a concept that aims to shift the traditional energy paradigm from centralized, top-down control to a more participatory and decentralized model. Energy communities, in this context, refer to groups of people or organizations that come together to collectively manage and govern their energy resources and infrastructure [14, 15]. Nondiscriminatory access to energy is a fundamental principle of energy democracy, and it emphasizes the importance of ensuring that all consumers have equal opportunities to access and benefit from affordable, reliable, and sustainable energy services. Energy communities are an important component of energy democracy, as they enable local communities to collectively own, generate, and manage their energy resources [16].

3

Techno-Economic Analysis for Centralized GH2 Power Systems

3.3

61

Market Design in Peer-to-Peer Energy Trading

Market design in peer-to-peer energy trading involves creating a framework and rules for energy transaction between individual prosumers within a decentralized network. This type of trading allows participants to buy/sell energy among themselves directly, bypassing traditional intermediaries such as utilities or grid operators [17]. Market designs for peer-to-peer (P2P) energy trading can be classified into three categories: centralized, decentralized, and distributed markets [18, 19]. Centralized markets prioritize global social welfare optimization and offer a higher level of certainty in terms of market outcomes. However, they suffer from challenges such as heavy computational and communication burdens, limited scalability and reliability, as well as increased concerns related to privacy and autonomy, and control performance. On the other hand, decentralized markets possess advantages and disadvantages that are the inverse of centralized markets. They excel in areas where centralized markets fall short [17, 20]. Decentralized markets provide a solution that emphasizes greater scalability, reliability, and reduced computational and communication requirements. Privacy and autonomy concerns are also alleviated to a certain extent [21]. Distributed markets represent a middle ground between centralized and decentralized markets, leveraging the strengths and weaknesses of both approaches. They seek to strike a balance by combining the advantages of centralized and decentralized markets while mitigating their respective drawbacks [22]. This hybrid model aims to optimize market efficiency and effectiveness. One crucial aspect in market design where P2P energy exchange finds significant application is market stability [23]. Ensuring the stability of the market is imperative. Additionally, it is worth emphasizing that the implementation of suitable mechanisms for revenue distribution and pricing is essential for centralized, decentralized, and distributed markets, respectively. These mechanisms play a vital role in augmenting the overall stability of the market [24]. In contrast to the traditional power market characterized by oligopoly and economies of scale, P2P energy trading can be viewed as an embodiment of the “sharing economy” concept [25, 26]. In fact, from the standpoint of power system operators, P2P energy trading offers a prospective solution for effectively handling the anticipated surge in distributed energy resource (DER) integration [27]. By developing appropriate P2P energy trading mechanisms, DERs have the potential to independently promote an enhanced local equilibrium in terms of power and energy, thus mitigating uncertainties for the upstream grid [28, 29]. Looking ahead, the future of market design entails addressing the inherent limitations of each market type. P2P markets typically emerge within low-voltage networks, primarily on a local scale, with the aim of maximizing the utilization of renewable and distributed energy resources [30]. Efforts should focus on overcoming the challenges associated with centralized, decentralized, and distributed markets. Additionally, market design should evolve to encompass broader functions, such as providing ancillary services for power systems. These services could include peak shaving, frequency response, voltage support, and other related capabilities to enhance the overall performance

62

M. M. Hayati et al.

Fig. 3.1 General framework of interactions between markets and other effective aspects of energy trading under the P2P platform

and stability of the power grid [31, 32]. Figure 3.1 illustrates the general framework of interactions between markets and other effective aspects of energy exchange under the P2P platform.

3.3.1

Centralized Peer-to-Peer Energy Market

Centralized P2P energy markets refer to platforms or systems that enable direct transactions of energy between prosumers without the involvement of intermediaries or traditional energy utilities [33, 34]. These markets leverage blockchain or other decentralized technologies to facilitate secure and transparent energy trading. These types of markets have the potential to transform the energy sector by increasing market participation, promoting renewable energy adoption, enhancing energy efficiency, and fostering a more sustainable and resilient energy ecosystem [35]. The opportunities provided by centralized P2P energy markets are numerous. This enables more efficient and cost-effective energy transactions. Centralized peer-to-peer energy markets open up opportunities for small-scale renewable energy producers, such as individual households with solar panels, to participate in the energy market. It promotes a decentralized and democratized energy landscape, empowering individuals and communities. Participants in these markets have the freedom to choose their energy sources and negotiate prices directly [36, 37]. This can foster competition, leading to better prices and options for consumers. Decentralized energy markets encourage the efficient use of energy by enabling real-time energy pricing [38]. Consumers can adjust their consumption patterns based on market signals, incentivizing energy conservation and load shifting. By decentralizing energy production and distribution, P2P energy markets enhance the resilience and reliability of the overall energy system [22, 39]. They can better handle disruptions, such as natural disasters or grid failures, by enabling localized energy sharing and self-sufficiency. Peer-to-peer energy markets facilitate

3

Techno-Economic Analysis for Centralized GH2 Power Systems

63

the integration of RESs into the grid. They enable efficient utilization of locally generated renewable energy, reducing the need for long-distance transmission and improving overall energy sustainability. Blockchain-based platforms used in centralized P2P energy markets provide transparency and immutability of transactions. This enhances trust among participants and helps ensure fair and secure energy trading [40]. These markets encourage innovation and entrepreneurship by creating opportunities for new business models, energy services, and technology solutions [41]. They can spur the development of energy storage, demand response systems, and other advanced energy technologies. Despite their advantages, centralized P2P energy markets face significant challenges. One such challenge is their diminished reliability and scalability. As these markets rely on a centralized authority to coordinate transactions, they are susceptible to potential failures and disruptions [42, 43]. Additionally, as the number of participants and volume of transactions increases, the communication and computational load on the centralized system becomes more burdensome, posing scalability issues. Furthermore, centralized P2P energy markets raise concerns about privacy. Since a central authority is responsible for overseeing and facilitating transactions, there is a risk of compromising individual privacy. The collection and processing of sensitive information by the central entity can raise valid privacy concerns. Moreover, the concentration of power in a centralized authority may lead to issues of autonomy and potential misuse of data, undermining the trust of market participants [44]. To overcome these limitations, future market designs should aim to strike a balance between centralized control and decentralized autonomy in P2P energy markets. Figure 3.2 illustrates an overview of the P2P energy trading framework.

3.3.2

Role of the Aggregators in Energy Communities

In a centralized P2P energy market, aggregators or energy community managers play a significant role in facilitating the efficient functioning of the market [45, 46]. This nonprofit organization, known as the aggregator, plays a vital role in the local energy community by serving as a hub for information sharing. The aggregator exercises direct oversight and authority over all energy transactions involving buyers and sellers, possessing the capability to both supply and manage energy, as well as offer pricing information to peers [47]. In fact, aggregators act as intermediaries between energy producers and consumers within the energy community [48]. They facilitate the P2P energy transactions, matching energy supply with demand and ensuring that the energy generated by local producers is consumed by local consumers whenever possible. Also, energy community managers work toward maximizing the overall efficiency of the energy community by optimizing the utilization of energy resources, minimizing transmission losses, and reducing peak demand [49, 50]. Figure 3.3 shows well the role of aggregators in energy communities.

64

M. M. Hayati et al.

Fig. 3.2 Overview of the P2P energy trading framework

Fig. 3.3 General framework of the role of aggregators in energy communities

3.3.3

Prosumer Preferences in Energy Communities

In energy communities with centralized P2P energy transactions, prosumers play a crucial role in shaping the preferences and dynamics of the system. Prosumers are individuals or businesses that both produce and consume energy within the

3

Techno-Economic Analysis for Centralized GH2 Power Systems

65

community. They typically have RES generation, such as photovoltaic (PV) panels or wind turbines, and are actively involved in trading surplus energy with other community members. Prosumers in energy communities usually prioritize renewable energy generation. They may have invested in solar panels, wind turbines, fuel cells, hydrogen energy storage systems, or other renewable energy technologies to reduce their carbon footprint and contribute to a greener energy system [11]. Prosumers often value the ability to generate and consume their own energy, aiming for self-sufficiency and reduced dependence on the traditional grid. They prefer to generate enough renewable energy to meet their own requirements and potentially store surplus energy for later use. Many prosumers are motivated by the potential cost savings associated with generating their own energy. By participating in centralized P2P energy transactions, they can sell excess energy to other community members and earn credits that offset their own energy consumption costs. Prosumers in energy communities often value community engagement and collaboration. A significant preference for prosumers is reducing their environmental impact. By generating clean, renewable energy, prosumers contribute to mitigating climate change and reducing reliance on fossil fuels. They see their participation in energy communities as a procedure to contribute actively to a more sustainable future. Prosumers value transparency in energy transactions and trust in the centralized P2P energy trading platform. They expect fair and reliable mechanisms for energy pricing, billing, and settlement, ensuring that the energy trading process is efficient, secure, and equitable for all participants. Prosumers are often early adopters of new energy technologies and innovations. They may show interest in smart grid solutions and advanced energy management tools that optimize their energy consumption, generation, and trading capabilities. These key preferences vary among prosumers based on their individual goals, motivations, and circumstances. Therefore, an effective energy community with centralized P2P energy trading should consider these preferences to create a supportive environment that encourages prosumers’ active participation and benefits the entire community [51]. Figure 3.4 illustrates a general framework of potential peer-to-peer billing procedure.

3.4

Feed-in-Tariffs (FiT) Mechanism in Energy Trading Market

The feed-in-tariffs (FiT) mechanism plays a significant role in the energy trading market. This mechanism serves as a supportive policy that encourages the development and integration of RESs into the grid. Under the FiT system, energy producers that generate electricity from renewable sources are offered long-term contracts with fixed, premium prices for the electricity they generate [52]. It should be noted that the states of Florida, Hawaii, Vermont, California, Maine, and Oregon from the United States of America have FiT mechanisms in their energy markets [53]. The FiT mechanism serves multiple purposes. First, it provides a

66

M. M. Hayati et al.

Fig. 3.4 Potential P2P billing procedure

financial incentive for renewable energy producers, ensuring a predictable revenue stream and promoting investment in renewable energy projects. This, in turn, helps achieve renewable energy targets and reduces dependency on fossil fuels [54]. Also, the FiT mechanism supports the growth of a diverse energy mix by facilitating the integration of intermittent RESs such as PV and wind power. By offering guaranteed prices, it mitigates the risks associated with the variability of renewable energy generation, thereby encouraging project developers to invest in these technologies [55]. Additionally, the FiT mechanism promotes energy independence and decentralization by empowering individual energy producers, such as homeowners with rooftop PV panels or small-scale wind farms, to contribute their excess energy to the grid. This allows for a more distributed energy generation system and reduces reliance on centralized power plants [56].

3

Techno-Economic Analysis for Centralized GH2 Power Systems

3.5

67

GH2 System Modeling

The suggested network is made up of a collection of eight prosumers that are arranged in a neighborhood structure. Each prosumer is outfitted with photovoltaic (PV) panels, which operate as the prosumer’s primary source of renewable energy generation. Additionally, each prosumer is equipped with electrical energy storage systems (ESSs). In addition, a subset of these prosumers – specifically, prosumers 1, 2, 5, 6, and 8 – have additional hydrogen facilities such as an electrolyzer, a fuel cell, and a hydrogen tank.

3.6

Problem Formulation

The objective function of this proposed model is shown in Eq. (3.1) to minimize the overall cost. The cost function is composed of three parts, as shown in Eq. (3.2). The three sections of this objective function present the following aspects: the cost analysis of P2P energy transactions, the expenses associated with trading energy with the grid, and an examination of the generation costs. This chapter also considers the T = {t1, t2, . . ., t24} and I = {1, 2, . . ., 8} for showing the set of time intervals with the duration of an hour and peers participating in the P2P market. In this chapter, i shows the market participant’s number. obj = Min

ð3:1Þ

C Total i i2I

C Total = i

G Gen C P2P t,i þ C t,i þ C t,i

ð3:2Þ

t2T

The cost of the P2P market for every prosumer is shown by C P2P t,i and described with the following equations. Equation (3.3) shows the P2P cost for every peer participating in the P2P market. The first part demonstrates the cost of purchasing energy in the P2P market and the second part shows the revenue from selling energy shows the price of energy in the P2P market and is equal to the in this market. λP2P t mean value of the retail and FiT price. P2P B

C P2P t,i =

λP2P t Pt,i,j

P2P S

- λP2P t Pt,j,i

8t 2 T, i 2 N

ð3:3Þ

j2I P2P B

Pt,i,j

P2P B

Pt,i,j

P2P S

= Pt,j,i

P2P S

, Pt,j,i

8t 2 T, i, j 2 N

ð3:4Þ

≥ 0 8t 2 T, i 2 N

ð3:5Þ

68

M. M. Hayati et al. P2P B

≤ Pt

P2P S

≤ Pt

Pt,i,j

Pt,i,j

P2P S

ut,i

T Max P2P B ut,i

ð3:6Þ

T Max P2P S ut,i

ð3:7Þ

P2P B

þ ut,i

≤1

ð3:8Þ

Equation (3.4) demonstrates that the sold energy in the P2P market is equal to the bought energy in this market. In the following, Eqs. (3.5), (3.6), (3.7), and (3.8) ensure that the peers participating in this market do not act as a seller and buyers at P2P S P2P B T Max and ut,i are the binary variable and Pt is the maximum the same time. ut,i trading value. The energy transaction with the upper grid is checked out by Eq. (3.9). The cost of these transactions consists of two parts. The first part is about buying energy with the retail price λret t and the second part is the revenue from the selling energy to the grid with FiT price λFIT t . ret BG FIT SG CG t,i = λt Pt,i - λt Pt,i

ð3:9Þ

BG SG PG t,i = Pt,i - Pt,i

ð3:10Þ

SG PBG t,i,j , Pt,j,i ≥ 0 8t 2 T, i 2 N

ð3:11Þ

TG Max BG ut,i

ð3:12Þ

TG Max SG ut,i

ð3:13Þ

P2P B

ð3:14Þ

PBG t,i ≤ Pt PSG t,i ≤ Pt P2P S

ut,i

þ ut,i

≤1

Equation (3.10) calculates the net demand of prosumers from the grid. Equations (3.11), (3.12), (3.13), and (3.14) ensure that the energy transaction with the grid has positive value and simultaneous buying and selling never happens. PV C Gen t,i = ai Pt,i þ bi

ð3:15Þ

Equation (3.15) shows the generation and investment cost of the solar units. In this equation, ai and bi demonstrate the operation and investment costs. Utilizing hydrogen storage systems for improved energy management is an important aspect. The hydrogen storage system, which comprises an electrolyzer, hydrogen tanks, and a fuel cell, is a crucial part of the smart grid’s energy storage system. When energy prices are low, the electrolyzer turns extra energy into hydrogen, which is kept in the hydrogen tanks. Then, when energy prices are higher during peak times, the fuel cell can use the saved hydrogen to generate energy. This method gives peers more freedom in how they handle energy.

3

Techno-Economic Analysis for Centralized GH2 Power Systems

69

EL EL HyEL t,i = ηi Pt,i

ð3:16Þ

EL EL EL EL uEL t,i Pmin ≤ Pt,i ≤ ut,i Pmax

ð3:17Þ

FC FC PFC t,i = ηi Hyt,i

ð3:18Þ

FC FC FC FC uFC t,i Pmin ≤ Pt,i ≤ ut,i Pmax

ð3:19Þ

EL uFC t,i þ ut,i ≤ 1

ð3:20Þ

Equations (3.16), (3.17), (3.18), (3.19), and (3.20) show the conversion method of electricity energy to hydrogen by the electrolyzer in the hydrogen storage unit. The electrolyzer in the hydrogen storage unit uses an electrical current to split water molecules into hydrogen and oxygen gases. Also, the fuel cell unit can convert the stored hydrogen back into electricity by combining it with oxygen from the air. This process is known as the electrochemical reaction, where hydrogen ions and oxygen ions combine to produce water and release energy in the form of electricity. Furthermore, in this calculation, simultaneous charging and discharging are avoided EL by two binary variables of uFC t,i and ut,i . Finally, the state of charge (SOC) in the hydrogen storage tank is shown by Eq. (3.21). HyS HyS FC SOCHyS þ H EL t,i = SOCt - 1,i 1 - φi t,i - H t,i

ð3:21Þ

This local community is also equipped with a battery energy storage system, which helps store excess energy generated by RESs. The state of charge shown in Eq. (3.22) and the following equations (3.23), (3.25), (3.25), and (3.26) limit the charging/discharging of the battery energy storage system. ES ES Ch PDch þ PCh SOCES t,i = SOCt - 1,i 1 - φi t,i η - t,i =ηDch MIN SOCMAX ≤ SOCES t,i t,i ≤ SOCt,i

ð3:23Þ

Ch MAX Ch ut,i

ð3:24Þ

Dch MAX Dch ut,i

ð3:25Þ

PCh t,i ≤ Pt,i PDch t,i ≤ Pt,i

Ch uDch t,i þ ut,i ≤ 1 P2P B

BG PPV t,i þ Pt,i þ

Pt,i,j j2I

ð3:22Þ

ð3:26Þ P2P S

FC Demand þ PDch þ t,i þ Pt,i = Pt,i

Pt,j,i

EL þ PCh ð3:27Þ t,i þ Pt,i

j2I

Equation (3.27) shows the energy balance constraint for every peer participating in the P2P market. This equation ensures that the produced energy in every time slot shows the prosumer load is equal to the consumed energy. In this equation, PDemand t,i demand.

70

M. M. Hayati et al.

3.7

Simulation and Numerical Results

This section presents the simulations and outcomes of the considered centralized GH2 power system. To this end, we developed two case studies to analyze the optimality of the proposed approach. The first case consists of an optimal operation of centralized P2P energy trading and the second one provides the optimal coalition operation for centralized P2P transactions. All simulations have been performed in the GAMS software as a mixed-integer nonlinear programming (MINLP) to minimize the objective function values on a PC with Intel(R) Core-i7 7700k CPU and 16 GB RAM.

3.7.1

Input Data

The proposed system contains eight prosumers in a local formation. All prosumers have PV panels as their renewable generation system and electrical energy storage systems (ESSs). Also, prosumers 1, 2, 5, 6, and 8 contain hydrogen facilities (electrolyzer, fuel cell, and hydrogen tank). The PV generation output data are displayed in Fig. 3.5 [57]. The demand data of these prosumers are considered constant and shown in Fig. 3.6 [58]. Furthermore, we provided three different i1 i5

i3 i7

i2 i6

i4 i8

250

PV Output (kWh)

200 150 100 50 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

Time(h) Fig. 3.5 PV generation output data

3

Techno-Economic Analysis for Centralized GH2 Power Systems i1 i5

i2 i6

i3 i7

71

i4 i8

Electicity Demand (kWh)

800 700 600 500

400 300 200 100 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

Time(h) Fig. 3.6 Electricity demand data Retail Price

FIT Price

P2P price

Electricity Price ($/kWh)

0.16 0.14 0.12 0.1 0.08 0.06 0.04 0.02 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

Time(h) Fig. 3.7 Electricity prices

electricity prices: retail, feed-in-tariff (FiT), and P2P trading, as illustrated in Fig. 3.7 [59]. Facilities’ capacities and constraints are tabulated in Table 3.1 [60, 61]. Also, the PV units’ operation and investment costs are shown in Table 3.2 [62].

72

M. M. Hayati et al.

Table 3.1 Facilities capacity and constraints

Facilities capacities Parameter Values HyS 100 SOCmax

Units kW

Facilities constraints Parameter Values 60 ηEL i

SOCHyS min

10

kW

ηFC i

50

%

SOCES max SOCES min PEL max PEL min

400

kW

φES i

2

%

35

kW

2

%

62

kW

φHyS i ηCh, Dch

96

15

kW

Pt,i

Ch Max

100

kW

PFC max

60

kW

Pt,i

Dch Max

100

kW

PFC min

5

kW

i7 0.03485 0.12121

i8 0.03359 0.08702

Units %

%

Table 3.2 Operation and investment costs of PV units Prosumer ai ($) bi ($)

i1 0.03278 0.13364

i2 0.03071 0.08702

i3 0.03145 0.05905

i4 0.02442 0.15229

i5 0.02945 0.03418

i6 0.02634 0.10878

Table 3.3 Outputs of optimal operation of centralized P2P trading Prosumers Outputs

3.7.2

i1 ($) 23.76

i2 ($) 15.68

i3 ($) 17.67

i4 ($) -13.57

i5 ($) 24.57

i6 ($) 82.90

i7 ($) 10.79

i8 ($) 331.92

Total 379.5

Simulations Results

This subsection demonstrates the proposed GH2 power system simulation results in each case study, and a brief discussion is provided for each case.

3.7.2.1

Optimal Operation of Centralized P2P Trading

The numerical outputs of the centralized P2P transactions in the 24-h time horizon are displayed in Table 3.3. In this case, peers 1–4 earn revenue by trading electricity. In contrast, peers 5–8 incur costs to fulfill their demands. To better understand peers’ performance, Figs. 3.8 and 3.9 are illustrated, which represent the peer-to-peer and peer-to-grid transactions. As seen from P2P transactions (Fig. 3.8), most electricity exchanges happened at times 12, 15, and 20. This is because of high PV generations in times 12 and 15, high P2P prices in these hours, and high electricity demands in times 15 and 20. In peer-to-grid transactions (Fig. 3.9), the maximum electricity sold to the network is accrued in times 11–14 due to massive PV generations, less electricity demand, and high FiT prices. The

3

Techno-Economic Analysis for Centralized GH2 Power Systems i1

i2

i3

i4

i5

i6

i7

73

i8

200

P2P Transactions (kWh)

150 100 50 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 -50 -100 -150 -200

Time(h)

Fig. 3.8 Peer-to-peer electricity transactions

i1

i2

i3

i4

i5

i6

i7

i8

Peer-to-Grid Transactions (kWh)

1300

800

300

-200

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

-700

-1200

Time(h)

Fig. 3.9 Peer-to-grid electricity transactions

maximum purchased electricity is also accrued in times 2 and 20 due to high electricity demand and less FiT prices. The ESS capacities of prosumers are shown in Fig. 3.10. As can be seen, during periods 7–10, all peers charge their ESSs due to low trading prices and electricity loads. In contrast, during periods 11–15, they discharge ESSs to fulfill their demands and participate in peer-to-peer and peer-to-grid transactions as well. Afterward, they

74

M. M. Hayati et al. i1

i2

i3

i4

i5

i6

i7

i8

2000

Storage Capacity(kWh)

1800 1600 1400 1200 1000

800 600 400 200 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

Time(h) Fig. 3.10 ESSs capacity level

i2

i5

i1

i6

i8

50

Storage Capacity (kWh)

45 40 35 30 25 20 15 10

5 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

Time(h) Fig. 3.11 Hydrogen tank capacity level

repeat the charging procedure between periods 16–19 to supply their requirements in other hours. The hydrogen tank capacities for prosumers 1, 2, 5, 6, and 8 are shown in Fig. 3.11. As illustrated in the figure, all prosumers charged their hydrogen tanks by electrolyzer in off-peak periods and utilized the hydrogen to generate electricity by the fuel cell in their peak demand periods. These periods are presented in Fig. 3.12.

3

Techno-Economic Analysis for Centralized GH2 Power Systems

(a)

i1

i2

i5

i6

(b)

i8

Electricity (kWh)

Electricity (kWh)

i1

15

100 80

75

i2

i5

i6

i8

10

60 40 20

5 0

0 t8

t9

t10

t16

t10

t11

Time(h)

t12 Time(h)

t15

t17

Fig. 3.12 (a) Electricity consumed by electrolyzer and (b) electricity generated by fuel cell

Table 3.4 Outputs of optimal operation of centralized P2P trading Prosumers Scenario 1 Scenario 2

3.7.2.2

i1 ($) -23.75 -24.81

i2 ($) -15.67 -19.28

i3 ($) -17.66 -18.12

i4 ($) -13.57 -18.86

i5 ($) 21.37 22.63

i6 ($) 77.14 78.13

i7 ($) 0.63 2.02

i8 ($) 307.15 308.21

Total ($) 335.64 329.92

Optimal Coalition Operation of Centralized P2P Transactions

In this case, to highlight the coalition operation results, we developed this case in two scenarios to indicate coalition formation advantages/disadvantages to each prosumer: . Scenario 1: All prosumers are considered as a united coalition entity. . Scenario 2: Prosumers are separated into two coalitions. The first coalition is called supplier coalition, which contains peers 1–4, and the second one is called receiver coalition, which contains peers 5–8. The first scenario aims to decrease the total cost of peers participating in the coalition as a united entity without considering prosumers’ preferences. However, in the second scenario, the aim is to diminish the receiver coalition costs and increase the supplier coalition profits separately but correlated to each other. In this regard, supplier coalition can increase its revenue in electricity transactions based on its prosumers’ preferences. The numerical outputs for this case can clarify the differences between these two scenarios, which are tabulated in Table 3.4. The results of the first scenario represent that the costs of peers 5–8 faced a reduction equal to $43.86 (11.55%). However, revenues of peers 1–4 almost remained unchanged. This counts as a disadvantage for this scenario and causes profitable prosumers not to be eager to participate in such coalitions. In the second scenario, after considering peers’ preferences in the form of two coalitions, the profitable peers received a slight revenue equal to $10.42 on their coalition transactions compared to the first scenario. This scenario’s total cost is

76

M. M. Hayati et al.

(a)

i1

i2

i3

i4

i5

i6

i7

i8

500

P2P Transactions (kWh)

400 300

200 100 0 -100

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

-200 -300 -400 -500

(b) 500

Time(h) i1

i2

i3

i4

i5

i6

i7

i8

P2P Transactions (kWh)

400 300 200

100 0 -100

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

-200 -300

-400 -500

Time(h)

Fig. 3.13 Peer-to-peer electricity transactions: (a) Scenario 1 and (b) Scenario 2

reduced by $49.58 (13%) compared to the first case and $5.72 (1.7%) compared to the first scenario. This is right that a slight increscent (i.e., $4.7) accompanies the costs of peers 5–8, but the coalition seems fairer in this situation. Figure 3.13 demonstrates P2P transactions in both coalition scenarios. As can be observed at first sight, the P2P exchanges in the second case are more than the P2P trading rate in the first case in the whole time period (Fig. 3.8). Moreover, some mild differences exist in the prosumers’ trading patterns in each scenario. For instance, in the second scenario, P2P transactions do not happen in hour 2. Instead, the transactions rate sharply increases in hour 3. Similar status has also occurred in time slot 17, in which transaction rates decreased compared to the first scenario and then faced with increments in hours 18 and 19.

3

Techno-Economic Analysis for Centralized GH2 Power Systems

(a)

i1

i2

i3

i4

i5

i6

i7

77

i8

Peer-to-Grid Transactions (kWh)

1000 800 600 400 200 0 -200

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

-400 -600 -800 -1000

(b)

Time(h) i1

i2

i3

i4

i5

i6

i7

i8

Peer-to-Grid Transactions (kWh)

1000 800 600 400 200 0 -200

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

-400 -600 -800 -1000

Time(h)

Fig. 3.14 Peer-to-grid electricity transactions: (a) Scenario 1 and (b) Scenario 2

Peer-to-grid exchange output is shown in Fig. 3.14. In the first scenario, the purchased electricity from the main network decreased in hours 6, 7, 13, 14, and 20 compared to the first case. The main reason for these diminutions is the increments in P2P exchanges. In contrast, electricity purchased from the primary grid has increased in hour 4 due to low retail prices. Furthermore, electricity sold to the grid in hours 11, 13, and 22 decreased. In the second scenario, electricity purchased increased in hours 2, 4, 20, and 23 and decreased in hours 6, 7, and 13 compared to the first case. The primary reason is the retail prices in these periods. In addition, electricity sold in hours 11, 14, and 21 is lessened. The ESS charge/discharge patterns in both scenarios, as shown in Fig. 3.15, are similar. The differences are only in the hourly quantities of the capacity level of the ESSs. In the hydrogen system, hydrogen tanks’ capacities are illustrated in Fig. 3.16.

M. M. Hayati et al.

78

Storage Capacity (kWh)

(a)

i1

i2

i3

i4

i5

i6

i7

i8

2000 1800 1600 1400 1200 1000 800 600 400 200 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

Time(h)

Storage Capacity (kWh)

(b)

i1

i2

i3

i4

i5

i6

i7

i8

2000 1800 1600 1400 1200 1000 800 600 400 200 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

Time(h) Fig. 3.15 ESSs capacity level: (a) Scenario 1 and (b) Scenario 2

In the first scenario, the prosumers’ charge/discharge patterns are not the same as in the first case. Peers prefer to charge their hydrogen tanks at peak PV generation periods instead of off-peak periods. The reason can be related to the formation of the united coalition entity. However, in the second scenario, prosumers charge their hydrogen tanks in off-peak hours, just similar to the first case. As explained earlier, the electrolyzer converts electricity to hydrogen and utilizes it to generate electricity by fuel cell in required periods. This procedure is displayed in Fig. 3.17. As noticed earlier, electrolyzers in the first scenario convert electricity to hydrogen at peak PV generation times (Fig. 3.17a) and utilize it to generate electricity at the peak demand hours (Fig. 3.17b). Unlike the first scenario, electrolyzers in the second scenario do this procedure in off-peak periods (Fig. 3.17c) and fuel cells deliver the electricity in peak periods (Fig. 3.17d).

3

Techno-Economic Analysis for Centralized GH2 Power Systems

(a)

i2

i5

i1

i6

79 i8

Storage Capacity (kWh)

70 60 50

40 30 20 10 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

Time(h) (b)

i2

i5

i1

i6

i8

30

Storage Capacity (kWh)

25

20 15 10 5 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

Time(h) Fig. 3.16 Hydrogen tank capacity level: (a) Scenario 1 and (b) Scenario 2

3.8

Conclusion

This chapter considers eight prosumers connected to each other with a low-voltage network in a local structure. Also, each peer is outfitted with photovoltaic (PV) panels, which operate as the prosumer’s primary source of renewable energy generation. Additionally, each peer is equipped with ESSs. The hydrogen storing systems in the proposed network include hydrogen tanks, water electrolysis, and fuel cells, which bring more flexibility in meeting energy needs. This local community is also equipped with a battery energy storage system, which helps store excess electricity generated by renewable units. The results of this study showed that P2P market players significantly reduced their costs in a centralized trading approach. In this study, two scenarios are considered for peers. First, all prosumers are considered as a united coalition entity,

80

M. M. Hayati et al.

(a)

i1

i2

i5

i6

(b)

i8

i1

60

100

Electricity (kWh)

Electricity (kWh)

120

80 60 40 20

i2

i6

i8

50 40 30 20 10 0

0 t5

t6

t8

t10

t16

t18

t6

t19

t7

(c)

i1

i2

i5

i6

(d)

i8

i1

15

Electricity (kWh)

80

t20

Time(h)

Time(h)

Electricity (kWh)

i5

60 40 20

i2

i5

i6

i8

10 5 0

0 t4

t5

t6

t8

Time(h)

t10

t16

t5

t6

t7

t20

Time(h)

Fig. 3.17 (a) Electricity consumed by electrolyzer in Scenario 1, (b) electricity generated by fuel cell in Scenario 1, (c) electricity consumed by electrolyzer in Scenario 2, and (d) electricity generated by fuel cell in Scenario 2

and then in the second scenario, prosumers are separated into two coalitions. The results of the first scenario showed that the costs of peers are 11.55%, and the total cost of the second scenario has decreased by 1.7% compared to the first scenario. In the first scenario, prosumers prefer to charge their hydrogen tanks at peak PV generation periods instead of off-peak periods. In the second scenario, prosumers charge their hydrogen tanks during off-peak hours.

References 1. Arsad, A., et al. (2022). Hydrogen energy storage integrated hybrid renewable energy systems: A review analysis for future research directions. International Journal of Hydrogen Energy, 47(39), 17285–17312. 2. Kovač, A., Paranos, M., & Marciuš, D. (2021). Hydrogen in energy transition: A review. International Journal of Hydrogen Energy, 46(16), 10016–10035. 3. Khaligh, V., Ghezelbash, A., Akhtar, M. S., Zarei, M., Liu, J., & Won, W. (2023). Optimal integration of a low-carbon energy system – A circular hydrogen economy perspective. Energy Conversion and Management, 292, 117354. 2023/09/15. 4. Alilou, M., Azami, H., Oshnoei, A., Mohammadi-Ivatloo, B., & Teodorescu, R. (2023). Fractional-order control techniques for renewable energy and energy-storage-integrated power systems: a review. Fractal and Fractional, 7(5), 391. 5. Daneshvar, M., & Mohammadi-Ivatloo, B. (2023). 1 – Overview of Internet of Things-based multi-energy management of cleaner multi-energy mix. In M. Daneshvar, B. Mohammadi-

3

Techno-Economic Analysis for Centralized GH2 Power Systems

81

Ivatloo, K. Zare, & A. Anvari-Moghaddam (Eds.), IoT enabled multi-energy systems (pp. 1–8). Academic. 6. Van Antwerpen, J., et al. (2023). A model for assessing pathways to integrate intermittent renewable energy for e-methanol production. International Journal of Hydrogen Energy. 7. Ebrahimi, M., & Sheikhi, A. (2023). A local integrated electricity-heat market design among multi Smart Energy Hubs with renewable energy generation uncertainty. Electric Power Systems Research, 218, 109217. 2023/05/01. 8. Aghdam, F. H., Mudiyanselage, M. W., Mohammadi-Ivatloo, B., & Marzband, M. (2023). Optimal scheduling of multi-energy type virtual energy storage system in reconfigurable distribution networks for congestion management. Applied Energy, 333, 120569. 9. Aminlou, A., Hayati, M. M., & Zare, K. (2023). Local peer-to-peer energy trading evaluation in micro-grids with centralized approach. In 2023 8th international conference on technology and energy management (ICTEM) (pp. 1–6). IEEE. 10. Aminlou, A., Mohammadi-Ivatloo, B., Zare, K., Razzaghi, R., & Anvari-Moghaddam, A. (2022). Peer-to-peer decentralized energy trading in industrial town considering central shared energy storage using alternating direction method of multipliers algorithm. IET Renewable Power Generation, 16(12), 2579–2589. 11. Aminlou, A., Nourollahi, R., & Zare, K. (2023). Optimal scheduling of local peer-to-peer energy trading considering hydrogen storage system. In V. Vahidinasab & B. MohammadiIvatloo (Eds.), Demand-side peer-to-peer energy trading (pp. 203–218). Springer. 12. Mehdinejad, M., Shayanfar, H., & Mohammadi-Ivatloo, B. (2022). Peer-to-peer decentralized energy trading framework for retailers and prosumers. Applied Energy, 308, 118310. 13. Wahlund, M., & Palm, J. (2022). The role of energy democracy and energy citizenship for participatory energy transitions: A comprehensive review. Energy Research & Social Science, 87, 102482. 14. Toubeau, J.-F., Teng, F., Morstyn, T., Von Krannichfeldt, L., & Wang, Y. (2022). Privacypreserving probabilistic voltage forecasting in local energy communities. IEEE Transactions on Smart Grid, 14(1), 798–809. 15. Saatloo, A. M., Mirzaei, M. A., & Mohammadi-Ivatloo, B. (2022). A robust decentralized peerto-peer energy trading in community of flexible microgrids. IEEE Systems Journal, 17(1), 640–651. 16. Mehdinejad, M., Shayanfar, H. A., Mohammadi-Ivatloo, B., & Nafisi, H. (2022). Designing a robust decentralized energy transactions framework for active prosumers in peer-to-peer local electricity markets. IEEE Access, 10, 26743–26755. 17. Bukar, A. L., et al. (2023). Peer-to-peer electricity trading: A systematic review on currents development and perspectives. Renewable Energy Focus. 18. Khorasany, M., Mishra, Y., & Ledwich, G. (2019). A decentralized bilateral energy trading system for peer-to-peer electricity markets. IEEE Transactions on Industrial Electronics, 67(6), 4646–4657. 19. Seyedhossein, S. S., & Moeini-Aghtaie, M. (2022). Risk management framework of peer-topeer electricity markets. Energy, 261, 125264. 20. Liu, C., & Li, Z. (2022). Comparison of centralized and peer-to-peer decentralized market designs for community markets. IEEE Transactions on Industry Applications, 58(1), 67–77. 21. AlSkaif, T., Crespo-Vazquez, J. L., Sekuloski, M., van Leeuwen, G., & Catalao, J. P. (2021). Blockchain-based fully peer-to-peer energy trading strategies for residential energy systems. IEEE Transactions on Industrial Informatics, 18(1), 231–241. 22. Tushar, W., et al. (2021). Peer-to-peer energy systems for connected communities: A review of recent advances and emerging challenges. Applied Energy, 282, 116131. 23. Capper, T., et al. (2022). Peer-to-peer, community self-consumption, and transactive energy: A systematic literature review of local energy market models. Renewable and Sustainable Energy Reviews, 162, 112403. 24. Parag, Y., & Sovacool, B. K. (2016). Electricity market design for the prosumer era. Nature Energy, 1(4), 1–6.

82

M. M. Hayati et al.

25. Cuenca, J. J., Jamil, E., & Hayes, B. (2021). State of the art in energy communities and sharing economy concepts in the electricity sector. IEEE Transactions on Industry Applications, 57(6), 5737–5746. 26. Schneiders, A., Fell, M. J., & Nolden, C. (2022). Peer-to-peer electricity trading and the sharing economy: Social, markets and regulatory perspectives. Energy Sources, Part B: Economics, Planning, and Policy, 17(1), 2050849. 27. Zhou, Y., Wu, J., & Long, C. (2018). Evaluation of peer-to-peer energy sharing mechanisms based on a multiagent simulation framework. Applied Energy, 222, 993–1022. 28. Hayati, M. M., Aminlou, A., Zare, K., & Abapour, M. (2023). A two-stage stochastic optimization scheduling approach for integrating renewable energy sources and deferrable demand in the spinning reserve market. In 2023 8th international conference on technology and energy management (ICTEM) (pp. 1–7). IEEE. 29. Zhang, C., Wu, J., Zhou, Y., Cheng, M., & Long, C. (2018). Peer-to-peer energy trading in a Microgrid. Applied Energy, 220, 1–12. 30. Aminlou, A., Hayati, M. M., & Zare, K. (2023). ADMM-based fully decentralized peer to peer energy trading considering a shared CAES in a local community. Journal of Energy Management and Technology. 31. Farrokhabadi, M., et al. (2017). Energy storage in microgrids: Compensating for generation and demand fluctuations while providing ancillary services. IEEE Power and Energy Magazine, 15(5), 81–91. 32. Villanueva-Rosario, J. A., Santos-García, F., Aybar-Mejía, M. E., Mendoza-Araya, P., & Molina-García, A. (2022). Coordinated ancillary services, market participation and communication of multi-microgrids: A review. Applied Energy, 308, 118332. 33. Gbadega, P. A., & Sun, Y. (2022). Centralized peer-to-peer transactive energy market approach in a prosumer-centric residential smart grid environment. Energy Reports, 8, 105–116. 34. Aminlou, A., Hayati, M. M., & Zare, K. (2023). P2P energy trading in a community of individual consumers with the presence of central shared battery energy storage. In V. Vahidinasab & B. Mohammadi-Ivatloo (Eds.), Demand-side peer-to-peer energy trading (pp. 143–159). Springer. 35. Mehdinejad, M., Shayanfar, H., & Mohammadi-Ivatloo, B. (2022). Decentralized blockchainbased peer-to-peer energy-backed token trading for active prosumers. Energy, 244, 122713. 36. Giotitsas, C., Pazaitis, A., & Kostakis, V. (2015). A peer-to-peer approach to energy production. Technology in Society, 42, 28–38. 37. Strbac, G., et al. (2019). Cost-effective decarbonization in a decentralized market: The benefits of using flexible technologies and resources. IEEE Power and Energy Magazine, 17(2), 25–36. 38. Bahrami, S., Amini, M. H., Shafie-khah, M., & Catalao, J. P. (2017). A decentralized electricity market scheme enabling demand response deployment. IEEE Transactions on Power Systems, 33(4), 4218–4227. 39. Shahidehpour, M., Yan, M., Shikhar, P., Bahramirad, S., & Paaso, A. (2020). Blockchain for peer-to-peer transactive energy trading in networked microgrids: Providing an effective and decentralized strategy. IEEE Electrification Magazine, 8(4), 80–90. 40. Kavousi-Fard, A., Almutairi, A., Al-Sumaiti, A., Farughian, A., & Alyami, S. (2021). An effective secured peer-to-peer energy market based on blockchain architecture for the interconnected microgrid and smart grid. International Journal of Electrical Power & Energy Systems, 132, 107171. 41. Mezquita, Y., Gil-González, A. B., Martín del Rey, A., Prieto, J., & Corchado, J. M. (2022). Towards a blockchain-based peer-to-peer energy marketplace. Energies, 15(9), 3046. 42. Guerrero, J., Chapman, A. C., & Verbič, G. (2019). Local energy markets in LV networks: Community based and decentralized P2P approaches. In 2019 IEEE Milan PowerTech (pp. 1–6). IEEE. 43. Azar, B. M., Mohammadian-Alirezachaei, H., & Kazemzadeh, R. (2023). Transactive energy and peer-to-peer trading applications in energy systems: An overview. In V. Vahidinasab &

3

Techno-Economic Analysis for Centralized GH2 Power Systems

83

B. Mohammadi-Ivatloo (Eds.), Demand-side peer-to-peer energy trading (pp. 47–75). Springer. 44. Bjarghov, S., et al. (2021). Developments and challenges in local electricity markets: A comprehensive review. IEEE Access, 9, 58910–58943. 45. Adeyemi, A., Yan, M., Shahidehpour, M., Bahramirad, S., & Paaso, A. (2020). Transactive energy markets for managing energy exchanges in power distribution systems. The Electricity Journal, 33(9), 106868. 46. Wu, Y., Wu, Y., Cimen, H., Vasquez, J. C., & Guerrero, J. M. (2022). Towards collective energy Community: Potential roles of microgrid and blockchain to go beyond P2P energy trading. Applied Energy, 314, 119003. 47. Khorasany, M., Mishra, Y., & Ledwich, G. (2020). Hybrid trading scheme for peer-to-peer energy trading in transactive energy markets. IET Generation, Transmission and Distribution, 14(2), 245–253. 48. Moret, F., & Pinson, P. (2018). Energy collectives: A community and fairness based approach to future electricity markets. IEEE Transactions on Power Systems, 34(5), 3994–4004. 49. Sousa, T., Soares, T., Pinson, P., Moret, F., Baroche, T., & Sorin, E. (2019). Peer-to-peer and community-based markets: A comprehensive review. Renewable and Sustainable Energy Reviews, 104, 367–378. 50. Koirala, B. P., Koliou, E., Friege, J., Hakvoort, R. A., & Herder, P. M. (2016). Energetic communities for community energy: A review of key issues and trends shaping integrated community energy systems. Renewable and Sustainable Energy Reviews, 56, 722–744. 51. Mahmoud, M., & Slama, S. B. (2023). Peer-to-peer energy trading case study using an AI-powered community energy management system. Applied Sciences, 13(13), 7838. 52. Parandeh, K., Bagheri, A., & Jadid, S. (2023). Optimal day-ahead dynamic pricing of gridconnected residential renewable energy resources under different metering mechanisms. Journal of Modern Power Systems and Clean Energy, 11(1), 168–178. 53. Fink, S., Porter, K., & Rogers, J. (2010). Relevance of generation interconnection procedures to feed-in tariffs in the United States. National Renewable Energy Lab (NREL). 54. Mendonça, M. (2009). Feed-in tariffs: Accelerating the deployment of renewable energy. Routledge. 55. Mabee, W. E., Mannion, J., & Carpenter, T. (2012). Comparing the feed-in tariff incentives for renewable electricity in Ontario and Germany. Energy Policy, 40, 480–489. 56. Branker, K., Shackles, E., & Pearce, J. M. (2011). Peer-to-peer financing mechanisms to accelerate renewable energy deployment. Journal of Sustainable Finance & Investment, 1(2), 138–155. 57. PVOutput. Available: https://pvoutput.org/. Accessed 11 Nov 2020. 58. PVOutput. Available: https://pvoutput.org/. Accessed 20 Dec 2022. 59. Rakipour, D., & Barati, H. (2019). Probabilistic optimization in operation of energy hub with participation of renewable energy resources and demand response. Energy, 173, 384–399. 60. Motallebi Azar, B., Mirzapour-Kamanaj, A., Kazemzadeh, R., Mohammadi-Ivatloo, B., & Zare, K. (2022). Optimal coalition operation of interconnected hybrid energy systems containing local energy conversion technologies, renewable energy resources, and energy storage systems. In Whole energy systems: Bridging the gap via vector-coupling technologies (pp. 169–198). Springer. 61. Nojavan, S., Zare, K., & Mohammadi-Ivatloo, B. (2017). Application of fuel cell and electrolyzer as hydrogen energy storage system in energy management of electricity energy retailer in the presence of the renewable energy sources and plug-in electric vehicles. Energy Conversion and Management, 136, 404–417. 2017/03/15. 62. Yang, J., Paudel, A., Gooi, H. B., & Nguyen, H. D. (2021). A Proof-of-Stake public blockchain based pricing scheme for peer-to-peer energy trading. Applied Energy, 298, 117154. 2021/09/ 15.

Chapter 4

Techno-Economic Analysis for Decentralized GH2 Power Systems Ali Aminlou, Mohammad Mohsen Hayati, Hassan Majidi-Garehnaz, Hossein Biabani, Kazem Zare, and Mehdi Abapour

4.1

Introduction

The global energy landscape is undergoing a profound transformation driven by the urgent need to mitigate climate change and transition toward a sustainable lowcarbon future [1, 2]. In this context, the integration of renewable energy sources and the development of decentralized energy systems are of paramount importance [3, 4]. The decentralized green hydrogen (GH2) power system, coupled with transactive energy (TE) and peer-to-peer (P2P) energy trading markets, emerges as a promising solution for enhancing the penetration of renewable energy and achieving a sustainable energy system [5, 6]. A sustainable energy system aims to minimize greenhouse gas emissions, promote the efficient use of resources, and ensure long-term energy security [7]. As fossil fuel dependency and carbon emissions continue to pose significant environmental and economic challenges, the transition to low-carbon alternatives has become imperative [8]. Renewable energy sources, such as solar, wind, and hydro, offer abundant and clean energy potential. However, their intermittent nature and variable power output create challenges in terms of grid stability and reliability [9]. To address these challenges, decentralized GH2 power systems have gained attention as a viable solution. These systems utilize renewable energy sources to produce green hydrogen through electrolysis, which can then be stored and utilized when renewable energy generation is insufficient [10]. By enabling the coupling of renewable energy integration and energy storage, decentralized GH2 power systems offer a reliable and dispatchable energy supply, thereby enhancing the stability and resilience of the grid [11, 12].

A. Aminlou (✉) · M. M. Hayati · H. Majidi-Garehnaz · H. Biabani · K. Zare · M. Abapour Faculty of Electrical and Computer Engineering, Energy Systems Research Institute (ESRI), Smart Energy Systems Lab, University of Tabriz, Tabriz, Iran e-mail: [email protected]; [email protected] © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 V. Vahidinasab et al. (eds.), Green Hydrogen in Power Systems, Green Energy and Technology, https://doi.org/10.1007/978-3-031-52429-5_4

85

86

A. Aminlou et al.

Furthermore, the incorporation of TE and P2P energy trading markets in decentralized GH2 power systems provides a novel approach to optimize energy generation, consumption, and distribution [13]. Transactive energy systems allow for real-time coordination of energy transactions between various participants, including prosumers (consumers who also produce energy) and energy service providers. This enables the efficient utilization of distributed energy resources, encourages demand response, and promotes local energy autonomy [14, 15]. Conducting a comprehensive techno-economic analysis is essential to assess the feasibility and viability of decentralized GH2 power systems. In this context, this chapter aims to provide a professional and scientific examination of the technoeconomic aspects of decentralized GH2 power systems and their integration with transactive energy and P2P energy trading markets. Additionally, it will highlight the key considerations and methodologies required for conducting a robust technoeconomic analysis in the context of sustainable energy systems. By shedding light on the techno-economic dimensions of decentralized GH2 power systems, this study seeks to contribute to the ongoing efforts toward a sustainable and low-carbon energy future. This research focuses on the design and optimization of a decentralized P2P energy trading system comprising eight interconnected peers. These peers can trade energy with each other and the upstream grid in a decentralized manner. The system incorporates various energy resources and storage technologies, including electric vehicles, hydrogen vehicles, battery energy storage systems (BESS), hydrogen energy storage systems, electrolyzers, and fuel cells. Also, in this study, the discussed optimization problem is solved with the help of the alternating direction method of multipliers (ADMM) algorithm. Within the realm of distributed optimization methods, researchers commonly favor the use of the ADMM due to its straightforward formulation and proven convergence for convex problems [16]. This algorithm addresses convex optimization problems by decomposing them into smaller, more manageable subproblems, facilitating their solution compared to the original problem.

4.2

Technological and Policy Approaches for the Integration of GH2 Resources in Renewable Energy Systems

Technological advancements have led to continuous efforts aimed at reducing the expenses associated with renewable energy sources. Numerous governments, including the United States of America and the European Union, have initiated measures to enhance the adoption and integration of renewable energy sources, particularly green hydrogen technology, into modern power systems. These initiatives involve the implementation of policies and support schemes. The most prevalent and effective approach to encourage the widespread use of renewable energy

4

Techno-Economic Analysis for Decentralized GH2 Power Systems

87

sources is the feed-in-tariffs (FiT) mechanism. Under this scheme, owners of distributed energy resources (DER) guarantee a minimum price for electricity generation [17]. The FiT mechanism in energy trading is a policy approach that encourages the adoption and generation of renewable energy. Under this mechanism, the owner of a renewable energy system, such as solar panels or wind turbines, is guaranteed a fixed price for the electricity they generate and feed into the grid. The governments or trustee institutions typically set the tariff, which is often higher than the market price for conventional energy sources. For instance, the British government proposed a reduction in solar PV production tariffs for new applicants starting in 2016 [18]. It should be noted that changes in FiT rates directly impact the benefits received by DER owners, leading to a decrease [19]. In the meantime, a new energy trading mechanism known as P2P energy trading has emerged, which can be considered as a successor to the FiT scheme [20]. P2P energy trading empowers DER owners to sell their excess energy to fellow consumers or sellers at more favorable rates in comparison to the FiT program. Consequently, consumers can access energy at lower costs within a competitive market, where they have the freedom to choose their energy supplier. This process ensures stable power supply to consumers. Moreover, peer-to-peer energy trading proves effective in reducing line losses and deferring costly network upgrades [21, 22]. As a result, local energy markets based on peer-to-peer energy trading bring numerous benefits to both consumers and DER owners. In the presence of green hydrogen resources, P2P energy trading can play a significant role in facilitating the exchange of energy generated from these renewable sources. With the integration of green hydrogen resources, P2P energy trading offers several benefits. First, it allows for the efficient utilization of surplus green hydrogen generated from renewable sources. Producers of green hydrogen can directly sell their excess supply to consumers or other users in the P2P energy trading network, promoting its widespread adoption and reducing waste. Second, P2P energy trading can provide a platform for consumers to access and purchase green hydrogen at competitive prices. Through direct transactions with hydrogen producers, consumers have the freedom to choose their energy supplier and have a transparent understanding of the source and quality of the green hydrogen they are purchasing.

4.3

Peer-to-Peer (P2P) Energy Trading Concept

Peer-to-peer (P2P) energy trading is gaining prominence as a viable management strategy for individuals who both produce and consume renewable energy within distribution grids. This new concept revolutionizes the traditional energy market by enabling direct transactions of electricity between consumers and prosumers (consumers who also produce energy) without the need for intermediaries [23]. In a P2P energy trading system, prosumers with renewable energy generation capabilities, such as solar panels or wind turbines, can sell their excess electricity directly to nearby consumers in real-time. This peer-to-peer interaction allows for more

88

A. Aminlou et al.

efficient utilization of distributed energy resources, promotes local energy production and consumption, and fosters community engagement. The concept of P2P energy trading aligns with the principles of transactive energy, where energy transactions occur based on real-time supply and demand dynamics. Through the integration of smart meters, IoT devices, and data analytics, participants can monitor their energy production and consumption patterns, negotiate prices, and autonomously execute transactions [24]. P2P energy trading offers several benefits to different stakeholders in the energy ecosystem. Prosumers have the opportunity to monetize their excess energy production, reduce reliance on the centralized grid, and gain more control over their energy choices [25]. Consumers can access cleaner and potentially cheaper energy options, support local renewable energy generation, and have greater visibility into the origin and environmental impact of the electricity they consume. Furthermore, P2P energy trading contributes to the integration of renewable energy sources into the grid, as it enables the efficient management of intermittent and decentralized generation. By allowing energy to be consumed closer to its source, transmission losses can be minimized, and grid congestion can be alleviated. This decentralized approach also enhances grid resilience and reliability by enabling local energy sharing during grid disruptions or emergencies [5, 26].

4.3.1

Decentralized P2P Trading Mechanism

As we discussed in the previous section, P2P energy trading introduces a novel approach within the smart grid domain, facilitating the establishment of local energy markets. To fully harness the advantages offered by P2P energy-sharing platforms, a growing number of prosumers and consumers should actively engage in local energy trading [27]. Prosumers interested in participating in the forward P2P energy market must initially forecast their local generation and demand, while market clearing mechanisms prioritize maximizing social welfare while considering network limitations [28, 29]. In decentralized P2P transactions, participants engage in trading without the involvement of an energy community manager, distinguishing it from centralized transactions [30]. Decentralized energy markets lack a central entity responsible for monitoring business activities. Within this market structure, participants have the freedom to determine transaction prices and peers possess ample control and independence in decision-making processes [31, 32]. Consequently, trading occurs directly between individual peers, ensuring participants’ privacy [33]. Nonetheless, the absence of the distribution operator’s participation in the transaction process can present challenges for the company in effectively planning resources, potentially leading to the reduced operational efficiency of the network. Additionally, this approach may result in a decline in social welfare for the P2P community as individual interests take precedence over community welfare [34].

4

Techno-Economic Analysis for Decentralized GH2 Power Systems

4.4

89

Hydrogen Energy Storage System (HESS)

In recent years, there has been a significant acceleration in the advancement of hydrogen-related technologies [35]. A hydrogen energy storage system (HESS) is a technology that utilizes hydrogen as a means to store and release energy [36, 37]. It involves converting electrical energy into hydrogen through the process of electrolysis and then storing the hydrogen for later use. When energy is needed, the stored hydrogen is converted back into electricity through a fuel cell or other hydrogenbased technology [38]. Consequently, the implementation of energy storage is regarded as a viable strategy to stabilize electricity supply, ensuring a harmonious equilibrium between generation and demand, and ensuring uninterrupted energy provision to consumers. In contrast to alternative energy storage systems, the main characteristic of HESSs is the versatility of utilizing the stored hydrogen in hydrogen-dependent industries or injecting it into the natural gas network to cater to gas consumers [39]. HESSs, similar to other energy storage technologies such as pumped storage units [40], compressed air energy storage (CAES) units, batteries, and electric vehicles [41], play a significant role in optimizing the equilibrium between electricity generation and consumption.

4.5

Problem Formulation

The objective of this research is to develop an optimal energy trading mechanism for interconnected peers. The system should enable efficient energy exchange and management, taking into account the availability and capabilities of the different energy resources and storage technologies. The system prioritizes the use of renewable energy sources to minimize reliance on fossil fuels and reduce carbon emissions. The system optimizes the utilization of battery energy storage, hydrogen energy storage, and other storage technologies to balance energy supply and demand, enhance system resilience, and enable continuous energy trading. The interconnected peers have the autonomy to engage in P2P energy trading individually. They can sell their excess energy production on the P2P market or to the upper network, providing flexibility in their energy trading activities. The system also supports the integration of electric vehicles, hydrogen vehicles, and various energy storage technologies, enhancing overall flexibility. To address the challenges and optimize the P2P energy trading system, the research proposes the use of advanced optimization algorithms such as the ADMM. This algorithm decomposes the main optimization problem into subproblems.

90

4.5.1

A. Aminlou et al.

Cost Modeling

In the proposed P2P market system, every peer has the freedom to participate in the market during each period. The ADMM algorithm, known for its decomposability, solves the optimization problems for each peer. As a result, each peer can make participation decisions based on conditions such as price, load, and PV generation. This freedom of choice empowers peers to make informed decisions regarding their involvement in the P2P market. This model considers the T = {t1, t2, . . ., t24} for the time interval set with the duration of 1 h. Furthermore, the set of I = {1, 2, . . ., 8} indicates the peers participating in the P2P market and i shows the market participants. ObjM = min

PtB E tBfG i - PtS EtS2G i

ð4:1Þ

t2T i2I

The objective function (4.1) in the proposed model represents the optimization goal. It consists of two components that quantify the energy exchange between the peers and the grid. The first part of the objective function represents the energy bought from the grid (E tBfG i ). It shows the amount of energy that the peers need to buy from the grid to meet their energy demands. PtB reflects the price of purchasing energy from the grid. The second part of the objective function represents the energy sold to the grid (EtS2G i). It measures the surplus energy generated by the peers, which is supplied back to the grid. This component reflects the potential revenue earned by the peers through selling their excess energy to the grid. By formulating the objective function in this manner, the model seeks to optimize the net energy flow between the peers and the grid. The objective is to minimize the energy bought from the grid, thereby reducing costs, while maximizing the energy sold to the grid, thereby generating revenue. This approach incentivizes the peers to generate surplus energy and actively participate in energy trading, promoting selfsufficiency and potentially improving economic outcomes.

4.5.1.1

Power Balance Constraint

Equation (4.2) represents a complex constraint that establishes the peer-to-peer energy trading system’s interconnectedness. This limitation assures that the energy bought by consumers from producers (E tBfP i,j ) is equal to the energy sold by producers (EtS2P j,i ) in the traditional decentralized system. This complicated constraint is incorporated into the decomposed objective function of each peer via the ADMM algorithm. The ADMM algorithm decomposes the global optimization problem into local subproblems, allowing each peer to optimize its own energy trading decisions. The ADMM algorithm offers effective coordination and optimization among interconnected peers. This strategy protects information privacy and individual

4

Techno-Economic Analysis for Decentralized GH2 Power Systems

91

peers’ autonomy by allowing them to autonomously select their energy trading strategies based on local factors such as price, load, and PV generation. This decentralized solution with the ADMM algorithm addresses the issues associated with complicated limitations in the P2P market while also increasing individual peers’ autonomy and decision-making capacities. In the following, Eq. (4.3) shows the energy balance constraint for each customer. In this equation, EtREN i and EtLoad i shows the renewable energy generation output and the demanded load. E tBfP i,j = E tS2P j,i 8t 2 T, j, i 2 I E tREN i þ

ð4:2Þ

EtBfP i,j þ EtBfG i þ EtEVDch i þ E tFC i = E tLoad i þ j2N j≠i

E tS2G i þ

E tS2P i,j þ EtEVCh i þ E tELC i 8t 2 T, i 2 I

ð4:3Þ

j2N j≠i

4.5.1.2

Plug-in Electric Vehicle Modeling

Equations (4.4)–(4.9) demonstrate the technical limits of plug-in electric vehicles (PEVs). The state of charge (SOC) constraint for EVs in the P2P energy trading system is represented by Eq. (4.4). This constraint (4.5) ensures that the battery SOC remains within a specified range to preserve battery health and performance. The model optimizes EVs as flexible energy resources by regulating the SOC, charging, and discharging activities to maximize benefits for EV owners. Equation (4.6) shows the energy consumed by the EV in the travel times. Eqs. (4.7) and (4.8) evaluate the power limitations of PEV in the charging and discharging times. These equations describe the maximum power thresholds for charging and discharging the PEV’s battery, respectively. In addition, Eq. (4.9) incorporates the PEV’s charge, discharge, and travel times as binary variables. This equation indicates that the PEV’s battery is prohibited from charging and discharging at the same time. This constraint ensures that the PEV operates either in a charging mode or a discharging mode at any given time. -1 SOCtEV i = SOCtEV i þ ηEV

t Ch E EVCh i

-

1 Et - E tTRVL i 8t 2 T, i 2 N ηEV Dch EVDch i

0 ≤ SOCtEV i ≤ SOCmax EV 8t 2 T, j, i 2 I

ð4:4Þ ð4:5Þ

92

A. Aminlou et al.

EtTRVL i = ΔtD i ΩEV i utEV E tEVCh i ≤ utEVCh i E Ch E tEVDch i

TVL i

ð4:6Þ ð4:7Þ

max i

≤ utEVDch i E Dch max i

utEVDch i þ utEVCh i þ utEV

4.5.1.3

TVL i

≤1

ð4:8Þ ð4:9Þ

Hydrogen Storage Unit Modeling

Hydrogen, a light, plentiful element, may help solve energy and environmental issues. For high-energy, short-term applications, compressed hydrogen gas is ideal. In this study, two number of units have a hydrogen storage unit, and that these units constitute a new set H = {i6, i7}. Electrolysis splits water into hydrogen and oxygen, which produces hydrogen. An electric current splits water molecules into oxygen and hydrogen at the anode and cathode, respectively. The electrolyzer mode uses off-peak electricity to produce hydrogen molecules that saves in the hydrogen tanks. Constraints (4.10) and (4.11) set maximum and minimum electrolyzer power consumption. Constraint (4.12) limits the electrolyzer’s hydrogen molar production. This constraint keeps hydrogen production within the system’s capability. This hydrogen may be used subsequently for power generation, transportation, or hydrogen storage. Furthermore, constraint (4.13) demonstrates that the generated hydrogen molar is a function of the electrolyzer’s used power. t EtELC i ≤ E max ELC i uELC i 8t 2 T, i 2 H

ð4:10Þ

t EtELC i ≥ E min ELC i uELC i 8t 2 T, i 2 H

ð4:11Þ

t N tELC i ≤ N max ELC i uELC i 8t 2 T, i 2 H

ð4:12Þ

N tELC i =

E tELC i ηELC LHV H2

8t 2 T, i 2 H

ð4:13Þ

During peak moments of high demand for electric energy, the stored hydrogen is used to produce power in the fuel cell. Constraint (4.14) specifies the maximum number of hydrogen molecules that the fuel cell may consume. This limitation guarantees that the fuel cell functions within a certain range of hydrogen consumption, limiting excessive use beyond its capacity. Also, constraints (4.15) and (4.16) limits the fuel cell maximum and minimum power generation. Furthermore, constraint (4.17) demonstrates the consumed hydrogen molar in the fuel cell. t N tFC i ≤ N max FC i uFC i 8t 2 T, i 2 H

ð4:14Þ

4

Techno-Economic Analysis for Decentralized GH2 Power Systems

93

t EtFC i ≤ E max FC i uFC i 8t 2 T, i 2 H

ð4:15Þ

t E tFC i ≥ Emin FC i uF i 8t 2 T, i 2 H

ð4:16Þ

N tFC i =

E tFC i 8t 2 T, i 2 H LHV H2 ηFC

ð4:17Þ

In Eq. (4.18), utFC i and utELC i show the binary variable of the fuel cell and electrolyzer, respectively. This binary variable prevents the simultaneous working of the fuel cell and electrolyzer. utFC i þ utELC i ≤ 18t 2 T, i 2 H -1 PtH2 i = PtH2 iþ

ℜΤH2 t N ELC i - N tFC i - N tH2V V H2 P0H2 i = Pinit H2 i i 2 H

i

8t 2 T, i 2 H

ð4:18Þ ð4:19Þ ð4:20Þ

Equation (4.19) also provides the dynamic concept of hydrogen storage pressure. This equation covers the pressure variations and behavior of the hydrogen storage system over time. This research studies the changes in pressure based on the volume, and temperature inside the hydrogen storage system when the gas is charged or released by integrating the ideal gas law. This provides for a more precise knowledge of the system’s behavior and aids in the development of effective storage techniques. Finally, Eq. (4.20) shows the initial state of pressure in the storage tank for H2 vehicle.

4.5.1.4

ADMM Algorithm Implementation

The ADMM algorithm is a distributed way to iteratively solving decentralized optimization problems. This algorithm iteratively solves the complicated problem and decompose the main optimization problem into several subproblems using augmented Lagrangian method. In this method, the ADMM multipliers updates in each iteration. In each iteration, we consider the estimated energy transaction data as a parameter that is not included in decision variables, and they are represented by a hat in the formulation. The aforementioned information is used to update the ADMM multipliers [42]. The objective function (4.1) for each prosumer i is decomposed using augmented Lagrangian technique by applying the ADMM algorithm to the model. This technique considers the objective functions (4.1) and (4.2) to decompose the main problem to the independent subproblems, which is shown in Eq. (4.21).

94

A. Aminlou et al.

Obji,k = min

PtB EtBfG i - PtS E tS2G i þ t2T

j2N

j2N

j≠i

j≠i

t

γ i = ρi=2

E k,t S2P i,j þ γ i 8i 2 N ð4:21Þ

k,t k,t λk,t j,i E BfP i,j - λi,j

EtBfP i,j - E S2P j,i

2

t

þ

EtS2P i,j - EBfP j,i

j2N

j2N

j≠i

j≠i

2

ð4:22Þ

Equation (4.22) shows the penalty value. The penalty value is reduced until it gets close to zero as the process converges. In this method, constraints (4.3)–(4.20) involve the decomposed objective function of each peer. By using the ADMM algorithm, every peer can solve its optimization problem individually and update the ADMM multilayers by Eq. (4.23), which shows the energy transaction price.

kþ1,t = λk,t λi,j i,j þ ρi

E k,t S2P i,j 8t 2 T, i 2 N

E k,t BfP j,i j2N

j2N

j≠i

j≠i

Pt þPt λ0,t i,j = ð B S Þ=28i, j 2 N, t 2 T

r k,t P,i =

k,t EBfP j,i -

k,t ES2P i,j 8t 2 T, i 2 N

j2N

j2N

j≠i

j≠i

ð4:23Þ

ð4:24Þ ð4:25Þ

The initial value of the ADMM multiplier is set to the average of the selling and buying price of the upper grid as shown in Eq. (4.24). The ADMM algorithm iteration continues until the convergence criteria are achieved. The residual value shown in Eq. (4.25) shows the convergence rate of the algorithm and if this value becomes less than a small value (ε), the iteration stops and the result is published between the peers.

4.6

Result and Analysis

Numerical analysis was performed in this work to assess the peer-to-peer (P2P) energy trading system in a local community. The results illustrate the potential for decentralized energy management to decrease total costs by establishing peer-topeer energy sharing networks. Furthermore, the investigation evaluates the

4

Techno-Economic Analysis for Decentralized GH2 Power Systems

95

performance of a new technology such as PEV and hydrogen electrolyzer system to store hydrogen when the electricity price is high and meet the hydrogen vehicle demand. This model can also generate electricity from the fuel cell technology when the electricity price is high.

4.6.1

Model Implementation and Data

In this work, we consider eight peers located in a local community. These peers are interconnected and can participate in P2P energy trading market (Fig. 4.1). In this model, every peer equipped with rooftop solar energy generation system that generated power is shown in Fig. 4.2. This figure shows that every peer has different energy generation in size and time because of installed system size and shadow effect [43]. These peers have different load profile, as shown in Fig. 4.3, which shows the different types of participants. By combining different energy resources and storage technologies, the system intends to promote efficient energy exchange and management. These include plugin electric cars, hydrogen vehicles, hydrogen energy storage systems, electrolyzers, and fuel cells. In this model, every peer has freedom to choose the energy source for supply their demand. Participant in the P2P market can use the potential of this market to reduce the overall cost. These peers can join the P2P energy trading market to earn more money as a seller and reduce their cost as a buyer. The ADMM model was implemented using the GAMS software (V24.9.1), which was then solved using the MINLP (mixed-integer nonlinear programming) model and BARON solver. The simulation lasted 24 h and was broken into 1-h time

Fig. 4.1 PV schematic of the prosumers participate in the P2P market

96

A. Aminlou et al. 25 Peer1

Peer2

Peer3

Peer4

Peer5

Peer6

Peer7

Peer8

Output of PV System (kW)

20

15

10

5

0 1

2

3

4

5

6

7

8

9

10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

Time (hour)

Fig. 4.2 PV generation profile of each prosumer 80

Forecasted Load Profile (kW)

70

Peer 1

Peer 2

Peer 3

Peer 4

Peer 5

Peer 6

Peer 7

Peer 8

60 50 40 30 20 10 0 1

2

3

4

5

6

7

8

9

10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

Time (hour)

Fig. 4.3 Forecasted load curve of every prosumer

periods. The goal of this experiment was to assess the impact of peer-to-peer energy trading on different microgrid features. In this proposed model, all peers except peer 6 and peer 7 have the PEV technology. These electric vehicles are connected to the grid and have the ability to charge and discharge energy to increase the flexibility of their peers. This PEV system acts like a battery energy storage system when they are connected to the grid but the difference appears when they leave the charging station. During this time the vehicle used the stored energy in its batteries. So, the minimum state of charge is required to leave the charging station. In this chapter, we consider that the peer 1, peer 2, and peer 8 electric vehicle leave the charging station at 8, 7, and 9 o’clock, respectively, and they arrive at 16, 17, and 15 o’clock. The state of charge (SOC) of PEV is shown in Fig. 4.4. This figure shows that the peers charge their batteries

4

Techno-Economic Analysis for Decentralized GH2 Power Systems

97

25 Peer 1 Peer 3 Peer 5

SOC of PEV(kWh)

20

Peer 2 Peer 4 Peer 8

15

10

5

0 1

2

3

4

5

6

7

8

9

10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

Time (hour)

Fig. 4.4 SOC of PEV 10 Charging Power of PEV(kW)

Peer 1

Peer 2

Peer 3

Peer 4

Peer 5

Peer 8

7.5

5

2.5

0 1

2

3

4

5

6

7

8

9

10 11 12 13 14 15 16 17 18 19 20 21 22 23 Time (hour)

Fig. 4.5 Charging powers of PEV

before departure and it is clear that the SOC level of the PEV do not change while they are stopped out of the charging station. These electric vehicles used 0.1667 kW/ km. PEVs help the peers energy management system by discharging in the higher electricity price. These PEVs reduce the overall cost by charging in lower energy price. Figures 4.5 and 4.6 shows the charging and discharging power from the PEV. As we described, the PEV owners cannot charge and discharge simultaneously. As shown in Fig. 4.1, the hydrogen storage unit consists of a hydrogen tank, water electrolyzer, and fuel cell. The hydrogen tank is responsible for storing the hydrogen gas, which can be used as a fuel source for various applications. This chapter considers the maximum pressure of the hydrogen tank to be 13.8 bar. The water electrolyzer plays a crucial role in the system by using electricity to split water molecules into hydrogen and oxygen gases. This process, known as electrolysis, allows for the production of hydrogen gas on demand. In this chapter, we consider a

98

A. Aminlou et al. 17.5

Discharging Power of PEV(kW)

Peer 1

Peer 2

Peer 3

Peer 4

Peer 5

Peer 8

15 12.5 10 7.5 5 2.5 0 1

2

3

4

5

6

7

8

9

10 11 12 13 14 15 16 17 18 19 20 21 22 23 Time (hour)

Fig. 4.6 Discharging powers of PEV 16 Peer 8 electrolyzer Peer 7 electrolyzer Peer 7 pressure Peer 8 pressure

Hydrogen Tank Pressure (kpa)

1200 1000

14 12 10

800 8 600 6 400

4

200

Electrolyzer Power (KW)

1400

2 0

0 1

2

3

4

5

6

7

8

9

10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

Time (hour)

Fig. 4.7 Hydrogen energy storage used power and stored hydrogen pressure

hydrogen vehicle with a demand of 1511 g of hydrogen before leaving the home. The fuel cell converts the stored hydrogen gas into electricity through an electrochemical reaction with oxygen from the air. This electricity can then be used to power vehicles, homes, or any other device that requires energy. This chapter considers that the water electrolyzer and fuel cell efficiency is 55 and 45%, respectively. Figure 4.7 shows the pressure of the hydrogen tank and electrolyzer power demand. Peer 6 and peer 7 are the only customers equipped with hydrogen storage units. This figure demonstrates that based on the energy price, peers try to meet their demand by increasing the hydrogen storage pressure. In the P2P market, every peer is an independent player that can manage its energy trading. This means that each peer in the P2P market has the ability to control and participate in the trading of energy. They are not dependent on any central authority or intermediary to facilitate their transactions. So, based on the energy price and generated energy, every peer decides to buy or sell energy to other peers in the

Techno-Economic Analysis for Decentralized GH2 Power Systems

Peer 7

Peer 7

Peer 6

Peer 6

Peer 4

Peer 4

Peer 3

Peer 3

Peer 2

10 9 8 7 6 5 4 3 2 1 0

Peer 2

Peer 1 Peer 5 Peer 6 Peer 7 Peer 8 Peer 5 Peer 6 Peer 8 Peer 5 Peer 6 Peer 8 Peer 5 Peer 6 Peer 8 Peer 5 Peer 6 Peer 8 Peer 5 Peer 6 Peer 8 Peer 5 Peer 8 Peer 8 Peer 5 Peer 6 Peer 8 Peer 5 Peer 8 Peer 1 Peer 1 Peer 5 Peer 1 Peer 1 Peer 2 Peer 4 Peer 6 Peer 7 Peer 8 Peer 2 Peer 4 Peer 5 Peer 6 Peer 7 Peer 8 Peer 2 Peer 5 Peer 6 Peer 7 Peer 8 Peer 2 Peer 3 Peer 6 Peer 7 Peer 8

Energy Bought from P2P Market

Peer 1

99

Peer 1

4

6t 7t 7t 7t 7t 8t 8t 8t 9t 9t 9t 10t 10t 10t 11t 11t 11t 12t 12t 12t 13t 13t 14t 15t 15t 15t 16t 16t 17t 18t 18t 19t 20t 21t 21t 21t 21t 21t 22t 22t 22t 22t 22t 22t 23t 23t 23t 23t 23t 24t 24t 24t 24t 24t

Time (hour)

Fig. 4.8 Energy transaction profile in the P2P market

market. This decentralized approach allows for greater flexibility and efficiency in energy trading, as peers can directly negotiate and agree on prices and quantities without the need for a middleman. Additionally, this peer-to-peer model promotes renewable energy generation and consumption, as it enables individuals with excess energy to sell it to those in need. Figure 4.8 demonstrates the various types of energy transactions that occur in the peer-to-peer market. This decentralized system fosters innovation and competition among peers, driving down prices and encouraging the adoption of cleaner energy solutions. Furthermore, it empowers individuals to take control of their energy choices and contribute to a greener future. The residual value of the ADMM method is shown in Fig. 4.9. The residual value is information used to assess the convergence of an optimization process, such as the ADMM. It represents how close the algorithm is to discovering the optimal solution. The residual value is commonly described in the context of decentralized energy trading as the difference between total sold energy and total purchased energy from the P2P market. It measures the amount of change between iterations and is employed to evaluate the algorithm’s progress toward convergence. The residual value is often shown across iterations to illustrate the algorithm’s convergence behavior. The x-axis in Fig. 4.9 depicts the number of iterations, while the y-axis represents the residual value. As the algorithm advances, the residual value should decrease, suggesting that the solution is approaching the optimum. As shown in this figure, the presented approach satisfies the stopping criteria in four iterations.

4.7

Conclusion

This chapter considers eight peers connected to each other with the low voltage network. These peers also have secure communication infrastructure. These peers are individual players in a local community. They have their own renewable energy sources such as solar panels. With the help of P2P energy trading systems, these individuals can now sell their surplus energy directly to their neighbors who may

100

A. Aminlou et al. 60000

Residual Value

50000 40000 30000 20000 10000 0 n1

n2

n3

n4

Iterations Fig. 4.9 The residual value of the ADMM algorithm in decentralized energy trading

have a higher demand. This not only promotes the adoption of renewable energy sources but also reduces the reliance on centralized power grids and fossil fuels. By enabling direct transactions between peers, these systems eliminate the need for intermediaries and associated costs, resulting in significant cost reductions for both buyers and sellers. Moreover, individuals gain more control over their energy choices as they can decide whom to buy from or sell to based on factors such as price, renewable energy generation, or even personal preferences. Furthermore, the addition of hydrogen storage units, which included a hydrogen tank, water electrolyzer, and fuel cell, gave further flexibility in meeting energy needs. The model aimed to optimize energy exchange and management by integrating various energy resources and storage technologies such as plug-in electric vehicles (PEVs) and hydrogen storage systems. The results demonstrated that P2P market players may substantially decrease total costs by utilizing the possibilities of this decentralized trading platform. PEV technology allows peers to operate as flexible energy storage devices, charging during low-priced times and discharging during high-priced periods. This strategy reduced costs and simplified overall energy trading between the peers. Compared to conventional energy markets, the activation of the P2P market will reduce costs by 4%. This cost reduction is a significant advantage that attracts both consumers and producers to participate in peer-to-peer energy trading.

4

Techno-Economic Analysis for Decentralized GH2 Power Systems

101

References 1. Daneshvar, M., Asadi, S., & Mohammadi-Ivatloo, B. (2021). “Overview of the grid modernization and smart grids,” Grid modernization– future energy network infrastructure: Overview, uncertainties, modelling, optimization, and analysis, pp. 1–31. 2. Kabiri-Renani, Y., Daneshvar, M., & Mohammadi-Ivatloo, B. (2022). Transactive energy revolution: Innovative leverage for reliable operation of modern energy networks – A critical review. IET Renewable Power Generation, 16(15), 3368–3383. 3. Ahmadi, E., McLellan, B., Mohammadi-Ivatloo, B., & Tezuka, T. (2020). The role of renewable energy resources in sustainability of water desalination as a potential fresh-water source: An updated review. Sustainability, 12(13), 5233. 4. Daneshvar, M., Mohammadi-Ivatloo, B., Zare, K., & Asadi, S. (2021). Transactive energy management for optimal scheduling of interconnected microgrids with hydrogen energy storage. International Journal of Hydrogen Energy, 46(30), 16267–16278. 5. Aminlou, A., Mohammadi-Ivatloo, B., Zare, K., Razzaghi, R., & Anvari-Moghaddam, A. (2022). Peer-to-peer decentralized energy trading in industrial town considering central shared energy storage using alternating direction method of multipliers algorithm. IET Renewable Power Generation, 16(12), 2579–2589. 6. Daneshvar, M., Mohammadi-Ivatloo, B., & Zare, K. (2022). A novel transactive energy trading model for modernizing energy hubs in the coupled heat and electricity network. Journal of Cleaner Production, 344, 131024. 7. Dincer, I., & Aydin, M. I. (2023). New paradigms in sustainable energy systems with hydrogen. Energy Conversion and Management, 283, 116950. 8. Zhou, Y. (2022). Low-carbon transition in smart city with sustainable airport energy ecosystems and hydrogen-based renewable-grid-storage-flexibility. Energy Reviews, 1, 100001. 9. Hayati, M. M., Aminlou, A., Zare, K., & Abapour, M. (2023). A two-stage stochastic optimization scheduling approach for integrating renewable energy sources and deferrable demand in the spinning reserve market. In 2023 8th international conference on technology and energy management (ICTEM) (pp. 1–7). IEEE. 10. Lebrouhi, B., Djoupo, J., Lamrani, B., Benabdelaziz, K., & Kousksou, T. (2022). Global hydrogen development-A technological and geopolitical overview. International Journal of Hydrogen Energy, 47, 7016. 11. Oraibi, W. A., Mohammadi-Ivatloo, B., Hosseini, S. H., & Abapour, M. (2023). Multi microgrid framework for resilience enhancement considering mobile energy storage systems and parking lots. Applied Sciences, 13(3), 1285. 12. Oraibi, W. A., Mohammadi-Ivatloo, B., Hosseini, S. H., & Abapour, M. (2023). A resilienceoriented optimal planning of energy storage systems in high renewable energy penetrated systems. Journal of Energy Storage, 67, 107500. 13. Abdolahinia, H., Lesani, H., & Moeini-Aghtaie, M. (2023). Decentralized transactive energy market framework under network constraints to deal with technical issues in a radial distribution network. Electric Power Systems Research, 223, 109416. 14. Zou, Y., Xu, Y., Feng, X., Naayagi, R., & Soong, B. H. (2022). Transactive energy systems in active distribution networks: A comprehensive review. CSEE Journal of Power and Energy Systems, 8(5), 1302–1317. 15. Gupta, N., Prusty, B. R., Alrumayh, O., Almutairi, A., & Alharbi, T. (2022). The role of transactive energy in the future energy industry: A critical review. Energies, 15(21), 8047. 16. Munsing, E., & Moura, S. (2018). Cybersecurity in distributed and fully-decentralized optimization: Distortions, noise injection, and ADMM. arXiv preprint arXiv:1805.11194. 17. Gipe, P. (2006). Renewable energy policy mechanisms. Renewable Energy. 18. Jenkins, N., Long, C., & Wu, J. (2015). An overview of the smart grid in Great Britain. Engineering, 1(4), 413–421.

102

A. Aminlou et al.

19. Tushar, W., Yuen, C., Mohsenian-Rad, H., Saha, T., Poor, H. V., & Wood, K. L. (2018). Transforming energy networks via peer-to-peer energy trading: The potential of game-theoretic approaches. IEEE Signal Processing Magazine, 35(4), 90–111. 20. Zhou, Y., Wu, J., Long, C., & Ming, W. (2020). State-of-the-art analysis and perspectives for peer-to-peer energy trading. Engineering, 6(7), 739–753. 21. Morstyn, T., Farrell, N., Darby, S. J., & McCulloch, M. D. (2018). Using peer-to-peer energytrading platforms to incentivize prosumers to form federated power plants. Nature Energy, 3(2), 94–101. 22. Shrestha, A., et al. (2019). Peer-to-peer energy trading in micro/mini-grids for local energy communities: A review and case study of Nepal. IEEE Access, 7, 131911–131928. 23. Aminlou, A., Hayati, M. M., & Zare, K. (2023). Local peer-to-peer energy trading evaluation in micro-grids with centralized approach. In 2023 8th international conference on technology and energy management (ICTEM) (pp. 1–6). 24. Zhang, C., Wu, J., Zhou, Y., Cheng, M., & Long, C. (2018). Peer-to-peer energy trading in a microgrid. Applied Energy, 220, 1–12. 25. Xia, Y., Xu, Q., Huang, Y., & Du, P. (2023). Peer-to-peer energy trading market considering renewable energy uncertainty and participants’ individual preferences. International Journal of Electrical Power & Energy Systems, 148, 108931. 26. Mehdinejad, M., Shayanfar, H., & Mohammadi-Ivatloo, B. (2022). Peer-to-peer decentralized energy trading framework for retailers and prosumers. Applied Energy, 308, 118310. 27. Aminlou, A., Hayati, M., & Zare, K. (2023). ADMM-based fully decentralized peer to peer energy trading considering a shared CAES in a local community. Journal of Energy Management and Technology, 100, 105021. 28. Suthar, S., Cherukuri, S. H. C., & Pindoriya, N. M. (2023). Peer-to-peer energy trading in smart grid: Frameworks, implementation methodologies, and demonstration projects. Electric Power Systems Research, 214, 108907. 29. Seyfi, M., Mehdinejad, M., Mohammad-Ivatloo, B., & Shayanfar, H. (2021). Decentralized peer-to-peer energy trading for prosumers considering demand response program. In 2021 11th smart grid conference (SGC) (pp. 1–5). IEEE. 30. Aminlou, A., Hayati, M. M., & Zare, K. (2023). P2P energy trading in a community of individual consumers with the presence of central shared battery energy storage. In V. Vahidinasab & B. Mohammadi-Ivatloo (Eds.), Demand-side peer-to-peer energy trading (pp. 143–159). Springer International Publishing. 31. Mehdinejad, M., Shayanfar, H. A., Mohammadi-Ivatloo, B., & Nafisi, H. (2022). Designing a robust decentralized energy transactions framework for active prosumers in peer-to-peer local electricity markets. IEEE Access, 10, 26743–26755. 32. Mehdinejad, M., Shayanfar, H., & Mohammadi-Ivatloo, B. (2022). Decentralized blockchainbased peer-to-peer energy-backed token trading for active prosumers. Energy, 244, 122713. 33. Luo, F., Dong, Z. Y., Liang, G., Murata, J., & Xu, Z. (2019). A distributed electricity trading system in active distribution networks based on multi-agent coalition and Blockchain. IEEE Transactions on Power Systems, 34(5), 4097–4108. 34. Morstyn, T., Teytelboym, A., & Mcculloch, M. D. (2019). Bilateral contract networks for peerto-peer energy trading. IEEE Transactions on Smart Grid, 10(2), 2026–2035. 35. Arsad, A. Z., et al. (2022). Hydrogen energy storage integrated hybrid renewable energy systems: A review analysis for future research directions. International Journal of Hydrogen Energy, 47(39), 17285–17312. 36. Oskouei, M. Z., et al. (2022). A critical review on the impacts of energy storage systems and demand-side management strategies in the economic operation of renewable-based distribution network. Sustainability, 14(4), 2110. 37. Aminlou, A., Nourollahi, R., & Zare, K. (2023). Optimal scheduling of local peer-to-peer energy trading considering hydrogen storage system. In V. Vahidinasab & B. MohammadiIvatloo (Eds.), Demand-side peer-to-peer energy trading (pp. 203–218). Springer International Publishing.

4

Techno-Economic Analysis for Decentralized GH2 Power Systems

103

38. Nojavan, S., Zare, K., & Mohammadi-Ivatloo, B. (2017). Application of fuel cell and electrolyzer as hydrogen energy storage system in energy management of electricity energy retailer in the presence of the renewable energy sources and plug-in electric vehicles. Energy Conversion and Management, 136, 404–417. 39. Mirzaei, M. A., Sadeghi Yazdankhah, A., & Mohammadi-Ivatloo, B. (2019). Stochastic security-constrained operation of wind and hydrogen energy storage systems integrated with price-based demand response. International Journal of Hydrogen Energy, 44(27), 14217–14227. 40. Nazari-Heris, M., Madadi, S., & Mohammadi-Ivatloo, B. (2018). Chapter 15: Optimal management of hydrothermal-based micro-grids employing robust optimization method. In A. F. Zobaa, S. H. E. Abdel Aleem, & A. Y. Abdelaziz (Eds.), Classical and recent aspects of power system optimization (pp. 407–420). Academic Press. 41. Heydarian-Forushani, E., Golshan, M. E. H., Shafie-khah, M., & Siano, P. (2018). Optimal operation of emerging flexible resources considering sub-hourly flexible ramp product. IEEE Transactions on Sustainable Energy, 9(2), 916–929. 42. Boyd, S., Parikh, N., Chu, E., Peleato, B., & Eckstein, J. (2011). Distributed optimization and statistical learning via the alternating direction method of multipliers. Foundations and Trends® in Machine learning, 3(1), 1–122. 43. “PVOutput.” (2020). Available: https://pvoutput.org/. Accessed 11 Nov 2020.

Chapter 5

Hydrogenation from Renewable Energy Sources for Developing a Carbon-Free Society: Methods, Real Cases, and Standards Mehdi Talaie, Farkhondeh Jabari, and Asghar Akbari Foroud

5.1

Status Quo, Challenges, and Outlook

Hydrogen is the simplest known chemical element. Hydrogen is one of the most abundant elements on earth. This element does not exist in nature in pure form and is always found in combination with other elements. It also requires a lot of energy to separate it from other elements. Among the most conventional sources of hydrogen on earth, we can mention water, biomass, and fossil sources. Hydrogen is a permanent, stable, indestructible, universal, and renewable system due to its independence from primary energy sources. It is expected that in the near future, hydrogen production and consumption will develop due to the lack of polluting gas production, the lack of carbon dioxide production, and the lack of impact on global warming, as well as high energy density. Hydrogen can be obtained from various primary sources such as primary energy sources, secondary energy sources, renewable sources, and various processes such as water electrolysis, natural gas reforming, and partial oxidation of hydrocarbons. Currently, hydrogen production in the world is mainly based on fossil fuels and the conversion of hydrocarbons or the water splitting by these sources, and most of the hydrogen produced around the world is obtained from natural gas and steam methane reforming (SMR) process.

M. Talaie · A. A. Foroud (✉) Electrical and Computer Engineering Faculty, Semnan University, Semnan, Iran e-mail: [email protected]; [email protected] F. Jabari Power Systems Operation and Planning Research Department, Niroo Research Institute (NRI), Tehran, Iran e-mail: [email protected] © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 V. Vahidinasab et al. (eds.), Green Hydrogen in Power Systems, Green Energy and Technology, https://doi.org/10.1007/978-3-031-52429-5_5

105

106

M. Talaie et al.

Nowadays, the hydrogen production from fossil fuels is considered the most economical way to produce large amounts of hydrogen. On the other hand, fossil fuels have endangered the environment due to polluting gases production. For example, the amount of CO2 released from the process of producing one ton of hydrogen through the reformation of natural gas is about 14 tons, and if coal is used, the emission of carbon dioxide will be doubled. Therefore, this method is not desirable for hydrogen production, and the need for a stable energy source with less pollution is felt more than ever. Other challenges and obstacles in the field of hydrogen production include the following: • Currently, the cost of producing and transporting hydrogen is higher compared to other fuels. • Low demand for hydrogen prevents the increase of production capacity. • Hydrogen storage and transportation technologies should be developed. • Electrode materials are very valuable in the electrolysis process, and electrolysis technology requires high quality water. • The costs related to the initial investment of renewable energies are high. Therefore, the cost of electricity produced from them is high, which has reduced the desire to use them compared to fossil fuels. Therefore, in order to have more economic technologies, with higher efficiency and less environmental pollution, many researches and studies have been done to develop hydrogen production technologies. The vision is that in the coming years, hydrogen produced from clean and renewable sources will be used as a suitable and cheap fuel, and with the increase in the knowledge of using renewable energy, the amount of use of this type of energy will increase continuously. Therefore, green hydrogen and hydrogen production using renewable sources as an alternative process have been assigned a huge part of current studies.

5.2

Hydrogenation Using Renewable Energy

Due to the increase in world population, fossil fuel sources are decreasing. Fossil fuels have put the ecosystem at risk due to the production of environmental pollutants. Therefore, the need for sustainable energy sources with less pollution is felt more than before [1]. The process of renewing fossil fuel resources takes millions of years, and considering the increase in energy consumption, these resources will not be able to provide the required energy in the coming years. Therefore, these energy sources are considered nonrenewable. On the other hand, renewable energies are constantly being regenerated and their renewal process is very fast. Therefore, the huge RES, such as solar, wind, and geothermal energy, which are considered clean sources of energy, can be an ideal alternative to fossil fuels, and ultimately benefiting from these resources leads to improve the environmental situation.

5

Hydrogenation from Renewable Energy Sources for Developing a Carbon-Free. . .

107

In today’s world, hydrogen is more and more considered as a sustainable energy source [1]. Among the features that distinguish hydrogen from other energy carriers are its abundance on the surface of the earth, almost unique consumption, high efficiency of energy conversion, the possibility of storage in different ways (including gas, liquid, or in metal hydrides), the possibility of moving over long distances, the simplicity of converting to other types of energy, having a high heating value and a low heating value higher than most conventional fossil fuels, very little emission of pollutants, the reversibility of its production cycle, and reduction of greenhouse effects and environmental superiority compared to other fuels [2, 3]. The hydrogen is a permanent, stable, indestructible, comprehensive, and renewable system due to its independence from the primary sources of energy. It is expected that in the near future, the supply and demand of hydrogen as an energy carrier will spread throughout the world economy. Due to the fact that hydrogen is the lightest element in the world, this gas does not exist on earth and tends to move away from earth. The most conventional source of hydrogen on earth is water and other sources are coal, fossil fuel, and natural gases. In the generation of hydrogen, many agents such as availability of source, cost, quantity, and purity of hydrogen are very important. Hydrogen is obtained by various methods such as passing steam over hot carbon, thermal decomposition of hydrocarbons, reaction of sodium hydroxide or potassium on aluminum, and electrolysis of water. In an ideal hydrogen-based energy system with the aim of ensuring the security of energy supply, protecting the environment and improving the efficiency of the energy system, hydrogen is produced from RES and is used in various applications after being stored and transported to consumption locations. Hydrogen produced based on RES is called green hydrogen. Green hydrogen can reduce the fluctuations of variable renewable energy by acting as energy storage [4]. Because most hydrogen production methods are immature, the cost of hydrogen production is high and the production efficiency is low. Nearly 50% of the global demand for hydrogen is produced by natural gas reforming, and the remaining 30% is produced by oil reforming, 18% by coal gasification, 3.9% by water electrolysis, and 0.1% by other methods. In order to eliminate the effect of using fossil fuels on the environment, human health, and climate, hydrogen production should be developed from environmentally friendly methods and renewable energies. Green hydrogen refers to the generation of hydrogen based on environmentally friendly methods and renewable energies. Various methods can be used to produce hydrogen based on renewable energies, as shown in Fig. 5.1 [5].

5.2.1

Water Splitting

One of the methods of hydrogen generation is the decomposition of water in thermochemical cycles. The production of hydrogen by thermochemical method through water splitting is a chemical process that ultimately leads to its

108

M. Talaie et al.

Fig. 5.1 Hydrogen production methods based on RES

decomposition into two elements, hydrogen and oxygen. The process of dissociating water molecules into hydrogen and oxygen requires a lot of energy and is an energyconsuming process. The methods of hydrogen generation from the splitting of water molecules into its components are divided into three methods: electrolysis, thermolysis, and photocatalyst.

5.2.1.1

Hydrogen Production Using Electrolysis Method

The electrolysis method was first investigated and studied in 1830 by the English scientist, Michel Faraday. In the water electrolysis process, water molecules on the surface of the electrode are converted into their constituent elements, namely oxygen and hydrogen. Equation (5.1) shows the general reaction of water electrolysis [6]. 1 H2 O þ Energy → H2 þ O2 2

ð5:1Þ

In order to produce 1 mole of hydrogen, 1 mole of water is needed. Therefore, the transient negative electrical charges from the external circuit, i.e., electrons, must be equal to the positive electrical charges, transient ions from the electrolyte. Each electrolysis cell contains two electrodes called anode and cathode, which are immersed in the electrolyte solution. These two electrodes are connected to a direct current power source. The electrode connected to the positive pole of the power source is considered as anode and the electrode connected to the negative pole is considered as cathode. In a water electrolysis process, oxygen is produced at the

5

Hydrogenation from Renewable Energy Sources for Developing a Carbon-Free. . .

109

Fig. 5.2 Overview of an electrolysis cell

Fig. 5.3 Division of electrolyzers according to electrolyte

anode and hydrogen at the cathode. Figure 5.2 shows a general view of an electrolysis cell. Electrolyzers can be categorized according to the electrolyte used in them as shown in Fig. 5.3. Positive ions or cations move toward the cathode under the influence of the electric field. On the other hand, negative ions or anions pass through the electrolyte in the electric field and move toward the anode. Depending on the type of material used, electrolytes have the ability to pass anion, cation, or

110

M. Talaie et al.

both. In acidic electrolytes, cations, and in alkaline electrolytes, anions carry electric charge. Hydrogen production by electrolysis has many advantages. Among the advantages of hydrogen production by electrolysis, we can mention the high purity percentage (99.9%) of hydrogen produced, environmental protection, and the reduction of CO2 emissions, as well as benefit of excess renewable energy for hydrogen generation [7]. The electrolysis process is cost-effective when cheap electricity is available, and we also need hydrogen with a very high degree of purity. The most widely used electrolysis technologies are: (a) Proton exchange membrane (PEM), (b) Alkaline water electrolysis (AWE), and. (c) Solid oxide electrolysis cells (SOED), which will be discussed below. (a) Proton exchange membrane (PEM) This method uses an acidic polymer electrolyte to produce hydrogen, which allows only positive ions or cations to pass through this electrolyte, and anions and electrons cannot pass through it. Regarding the electrolysis of water, the hydrogen ions produced in the anode pass through the membrane and become hydrogen molecules by capturing electrons in the cathode. The general reaction performed in the water electrolysis of the proton exchange membrane is according to the relations (5.2), (5.3) and (5.4): Anode

1 H2 O → 2Hþ þ O2 þ 2e 2

Cathode Cell

2Hþ þ 2e - → H2 1 H2 O → 2Hþ þ O2 2

ð5:2Þ ð5:3Þ ð5:4Þ

The working temperature of this electrolyzer is limited from 25 to 80 °C because at higher temperatures the polymer chains of the electrolyte will start to break and the membrane will be destroyed. Normally, the amount of electricity required for this process is 23–26 MJ per 1 M3 of hydrogen production. The purity of hydrogen generated in this process reaches 99.99% [8]. The electrodes used in this type of electrolyzer must be very resistant to corrosion because the existing polymer membrane creates an acidic environment. In this regard, the type of electrodes is usually selected from the platinum group, which is very expensive. Therefore, making the electrode in this type of electrolyzer is one of the most expensive parts. Among the advantages of hydrogen generation by electrolysis method based on PEM electrolyzer, we can mention hydrogen production with high purity, fast response, high current density, small size, operation at low temperature, and high

5

Hydrogenation from Renewable Energy Sources for Developing a Carbon-Free. . .

111

efficiency [8]. The most important disadvantage of PEM is the high costs of this process. (b) Alkaline water electrolysis (AWE) This technology has been used since 1920 and is one of the best options for producing green hydrogen on a large scale. Hydrogen production by alkaline electrolysis has the simplest and most complete technology. This electrolyzer uses alkaline electrolyte with a pH greater than 7. In an alkaline electrolyte, the transient ionic species are the anions of the hydroxyl group. In these types of electrolyzers, the electrolyte is usually a liquid solution of sodium hydroxide or potassium hydroxide. Eqs. (5.5), (5.6) and (5.7) show electrochemical reactions in alkaline electrolyzer: Anode Cathode Cell

1 2OH - → H2 O þ O2 þ 2e 2

ð5:5Þ

2H2 O þ 2e - → H2 þ 2OH -

ð5:6Þ

1 H2 O → 2Hþ þ O2 2

ð5:7Þ

The purity of water used is one of the major concerns in the electrolysis process. The presence of metal atoms such as calcium and magnesium causes unwanted reactions and metal deposits on the electrodes, which reduce the activity of the electrodes over time. Also, the presence of salts such as sodium chloride or other sources containing chlorine ions in water causes the production of chlorine gas in the cathode, which is highly corrosive. Most of the commercial alkaline electrolyzers operate at a temperature between 70 and 90 °C and its density is between 1000 and 3000 A/m2. The generation of hydrogen by alkaline electrolysis method has the following advantages [9]: • • • •

The purity of hydrogen produced in this technology is approximately 99%. The investment cost in AWE technology is low. Alkaline electrolysis has the simplest and most mature technology. It does not depend on noble metals and includes performance at low pressures.

The disadvantages of this method include low current density and slow start-up of the electrolyzer, which negatively affects the size, capacity of the system to adapt to renewable energy fluctuations, and production costs [9]. (c) Solid oxide electrolysis cells (SOEC) The solid oxide electrolyzer is a developing process in which water vapor is converted into oxygen and hydrogen at high temperatures. The electrolyte used in this method is a nonporous metal oxide capable of transporting oxygen or hydrogen ions. Water vapor enters from the side of the porous cathode. By applying an electric current, water molecules penetrate into the reaction sites and turn into hydrogen gas and oxygen ions in the cathode–electrolyte interface. The produced hydrogen

112

M. Talaie et al.

penetrates the surface of the cathode and is collected. Oxygen ions migrate from the electrolyte to the anode. In the anode part, oxygen ions are oxidized to oxygen gas and penetrate to its surface from inside the porous anode. The electrochemical reactions in the solid oxide electrolyzer are as follows: Anode Cathode Cell

2O2 - → O2 þ 4e H2 O þ 2e - → H2 þ O2 1 H2 O → 2Hþ þ O2 2

ð5:8Þ ð5:9Þ ð5:10Þ

The solid oxide electrolyzer operates at a temperature between 500 and 900 °C using water vapor. The efficiency of this technology is economically higher compared to PEM and AWE. Also, solid oxide electrolysis cells in reverse mode can work as a fuel cell or co-electrolysis mode.

5.2.1.1.1

Advantages and Disadvantages of Electrolyzer Methods

• Electrolysis allows the use of excess renewable energy to produce hydrogen and increases the amount of hydrogen production and reduces its production cost. • For small and medium scale applications, PEM technologies and AWE are more suitable than solid oxide electrolysis cells [10]. • For applications with a scale larger than 1 MW, solid oxide electrolysis cell technology is more suitable than PEM and AWE technologies. Because this technology reuses thermal energy and waste heat, its efficiency increases. • Among the PEM electrolyzer technologies, solid oxide electrolysis cell, and alkaline water electrolysis, the most suitable technology for working with renewable energy sources and green hydrogen production is PEM electrolyzer technology [11–13]. • Due to the use of expensive materials and equipment in the manufacture of proton exchange membrane electrolyzer, this technology is more expensive than other technologies. • Alkaline water electrolysis cost is lower than PEM electrolysis technologies and solid oxide electrolysis cell. However, it is prone to catalyst degradation problems [12]. • The operation of SOEC technology is difficult due to the need and access to high temperature during the electrolysis process, and hence it is not a promising technology.

5.2.1.2

Hydrogen Production Using Photocatalyst Method

Green hydrogen generation by electrolysis requires electricity obtained from RES (wind, water, or solar energy), while solar energy can be directly used for

5

Hydrogenation from Renewable Energy Sources for Developing a Carbon-Free. . .

113

photocatalytic hydrogen generation [14]. Therefore, energy consumption for the hydrogen production process can be effectively reduced by this technology. Although both photocatalytic and electrocatalytic processes produce green hydrogen, photocatalytic hydrogen production seems to be a very clean and renewable hydrogen production method compared to electrocatalytic hydrogen. Hydrogen production by photocatalytic process resulting from water decomposition is a very suitable approach for green hydrogen production due to low energy consumption, low cost, and compatibility with the environment. The basis of the photocatalytic hydrogen process is a semiconductor photocatalyst that harnesses solar energy to decompose water. In this process, when the sunlight (photon) reaches an electron of the semiconductor located at the capacitance level, it transfers it to the conduction band if the energy gap is suitable, and a hole or empty space is created in the capacitance band. This generated energy gap can be used for water splitting, provided that the semiconductor energy gap is suitable. The energy gap should be enough to decompose both water and light to excite the electron. Based on whether the semiconductor is of n or p type, one of the following two situations occurs: p‐type n‐type

2H2 O þ 2e - → 2OH - þ H2

ð5:11Þ

1 H2 O þ 2h - → 2Hþ þ O2 2

ð5:12Þ

Scientists consider the acceptable energy gap for sunlight to be less than 2.2 electron volts. Unfortunately, semiconductors with this energy gap are unstable in water. Therefore, some semiconducting oxides are used in this field, because semiconducting oxides have the highest stability in water, but their energy gap should be reduced as much as possible, which is actually the main challenge of this method.

5.2.1.3

Hydrogen Production Using Thermolysis of Water

Thermolysis of water, also known as single-step separation of water, has the following reaction: 1 H2 O þ Heat → H2 þ O2 2

ð5:13Þ

This process requires a heat source that can provide up to 2500 K. One of the challenges of this hydrogen production method is to separate O2 and H2 gases. Existing semipermeable membranes can be used up to 2500 K and thus the mixture must be cooled before reaching the separation process. The results of the solar water thermolysis experiment conducted by Baykara showed that if the gaseous product is cooled quickly (in a few milliseconds) to a temperature of 1500–2000 K, the recombination of oxygen and hydrogen can be used to separate the hydrogen through a palladium membrane [15].

114

M. Talaie et al.

Table 5.1 Advantages, disadvantages of any technologies of hydrogen production from the breakdown of water molecules Renewable energy sources

Water splitting

Methods Electrolysis

PEM

AWE

SOEC

Photocatalysis

Thermolysis

Advantage Hydrogen production with high purity (99.99%) Fast response High current density Operation at low temperature High efficiency One of the best options for producing green hydrogen on a large scale Simplest and most mature technology The investment cost in AWE technology is low Performance at low pressures It does not depend on noble metals Higher efficiency compared to PEM and AWE Work as a fuel cell or co-electrolysis No need for electricity and direct use of solar energy A very clean and renewable hydrogen production method Low energy consumption, low cost, and compatibility with the environment It has environmental benefits

Disadvantage High costs of this process

Low current density Slow start-up of the electrolyzer

The need for high temperatures during the electrolysis process Low rate of hydrogen production

Requires a significant amount of heat source Low efficiency (about 30% due to photovoltaic conversion efficiency)

According to the mentioned content, a summary of the advantages and disadvantages of each hydrogen production technology from the breakdown of water molecules is shown in Table 5.1.

5

Hydrogenation from Renewable Energy Sources for Developing a Carbon-Free. . .

5.2.2

115

Biohydrogen (Production of Hydrogen from Biomass)

Water and biomass have been mentioned as the most important and abundant renewable resources for the production of green hydrogen. Biomass refers to resources that originate from natural organic materials that store sunlight. According to the definition of the United Nations Climate Change Working Group, biomass is: nonfossil and degradable organic matter that originates from plants, animals, or microorganisms. A suitable way to produce green hydrogen from RES is biohydrogen production technology. The generation of green hydrogen from biomass is very momentous compared to other hydrogen production technologies and has many advantages, among which the following can be mentioned [16, 17]: • Biomass as a carbon neutral energy source. It is also considered as a renewable resource. • Biohydrogen production has a relatively high conversion efficiency. • Energy use in the generation of biohydrogen is low. • Using biomass to produce green hydrogen, in addition to producing clean fuel, also reduces the volume of waste. • Using biomass resources for hydrogen production can reduce production costs. It also provides the possibility of more accessible and cheaper hydrogen gas production. • Biomass resources are available in large quantities and widely. Among the sources of biomass, as an example, we can mention agricultural products and waste, horticulture and food industry, animal waste, forests and forest waste, sewage and waste, and so on. It is possible to produce green hydrogen from biomass through thermochemical or biological processes [18, 19]. Hydrogen production through biological process is classified into the following ways: • • • •

water photolysis photo-fermentation dark-fermentation dark–photo co-fermentation

Also, the generation of green hydrogen from biomass through a chemical process is classified into the following four methods: • • • •

biomass gasification pyrolysis-reforming supercritical water conversion catalytic-reforming of small organic molecules Figure 5.4 presents the classification of biohydrogen production.

116

M. Talaie et al.

Fig. 5.4 Classification of biohydrogen production

There are also different classifications for biomass sources. The US Department of Energy has classified biomass resources into three categories: primary, secondary, and tertiary materials as follows [20]: • Primary materials: all land plants that are produced by photosynthesis and exist on land and water; • Secondary materials: includes all waste and by-products of food industries, wood, forestry, and animal waste; • Tertiary materials: includes all wastes and postconsumer waste such as fats, urban solid waste, wood waste from urban environments, packaging waste, sewage, and landfill gas. Also, according to the studies conducted to determine the potential of biomass resources, the National Renewable Energy Laboratory of America has presented the following classification [21]: • Agricultural waste including agronomic waste and methane from animal excreta • Wood waste including forest waste, wood industries, and urban wood waste (gardens, branches, and wood waste) • Urban waste includes landfill gas, sewage gas • Energetic plants As mentioned, the production of hydrogen from biomass is called biohydrogen. This method provides hydrogen production by chemical or biological processes [18, 19], which we will describe below.

5

Hydrogenation from Renewable Energy Sources for Developing a Carbon-Free. . .

5.2.2.1

117

Biological Methods

In biological technology, microbial metabolism and various types of micro-organisms are used to react and decompose biomass into required products. Common biological technologies include water photolysis, photo-fermentation, dark-fermentation and dark–photo co-fermentation.

5.2.2.1.1

Water Photolysis

Hydrogen generation through photolysis of water is done by two methods, direct biophotolysis and indirect biophotolysis, both of which use sunlight. Direct production of hydrogen from water using sunlight energy through biological systems is called direct biophotolysis. In other words, this technology uses microorganisms to break down water through photosynthesis and produce hydrogen and oxygen [22]. In the direct biophotolysis process, microalgae are used to convert sunlight into hydrogen. Two photosynthetic systems work in this process. The primary photo system reduces CO2 and the secondary photo system decomposes water. The advantage of this process is the production of hydrogen from water under standard temperature and pressure in an aqueous environment. The speed of the hydrogen production process through photolysis of water is influenced by many factors, including raw materials, catalyst, temperature, and light intensity. Among the disadvantages of this process, we can mention the low efficiency of this process in hydrogen production, the small volume of hydrogen production, the sensitivity of enzymes involved in biological hydrogen production, and the long cycle of hydrogen production [23].

5.2.2.1.2

Photo-Fermentation

This process is an anaerobic fermentation in which hydrogen production is done under anaerobic conditions using light energy and reducing the ability of small molecular organic matter. In the photo-fermentation process, various types of organic substrates and reduced compounds such as organic acids are oxidized under anaerobic conditions and light and produce carbon dioxide and hydrogen. Nitrogenase is the basic enzyme of this process. Nitrogenase can produce hydrogen in the absence of nitrogen by using light energy and reduced compounds as electron donors and ferredoxin as an electron transporter [24]. Green and purple sulfur bacteria, non-sulfur bacteria, and some green algae are able to perform these processes. In order to produce hydrogen through bacteria, soybean wastewater, dairy wastewater, and starch wastewater can be used as ideal substrates. All routes of hydrogen production in this method can be according to Eq. (5.14):

118

M. Talaie et al.

ðCH2 OÞx → Ferredoxin → Dinitrogenase → H2

ð5:14Þ

The speed of the hydrogen generation in this process is influenced by the pH value of the solution, the concentration of the substrate, the kind of fermentation microorganism, and the light intensity [25].

5.2.2.1.3

Dark-Fermentation

In this process, anaerobic bacteria are very effective in producing biohydrogen. These bacteria produce biohydrogen by using biochemical energy stored in organic materials under anaerobic conditions and independent of light. In other words, in this biochemical process, the energy stored in organic materials is converted into another form of energy by anaerobic bacteria in the dark, i.e., conditions independent of light (or in low light). Since the bioreactors used in dark-fermentation do not need to provide light, they are much simpler and cheaper than photo-fermentation reactors [26]. This process has received special attention due to its simplicity of operation, flexibility in cultivation, stability of reliable hydrogen production, simultaneous production of hydrogen and consumption of organic waste, and no need for light. In this process, various organic substrates are converted into hydrogen by facultative or absolute anaerobic bacteria in anaerobic and dark conditions. The energy required in this process is provided by the oxidation of organic substrates instead of sunlight. In this process, various metabolic products are produced, the amount and composition of which is different according to the type of microorganism and process conditions. Low efficiency, low production capacity per unit of hydrogen production, and the need for a large area are the biggest challenges of this process.

5.2.2.1.4

Dark–Photo Co-fermentation

Due to the fact that this technology simultaneously uses the unique advantages of hydrogen production methods in photo-fermentation and dark-fermentation processes [27], it has a higher efficiency than the other two processes. Due to this advantage, hydrogen production by dark–photo co-fermentation technology can be significantly increased. The increase in efficiency and the increase in the production rate make the cost of hydrogen production lower compared to the other two technologies.

5

Hydrogenation from Renewable Energy Sources for Developing a Carbon-Free. . .

5.2.2.2

119

Chemical Methods

In chemical methods, biomass is first converted into hydrogen-rich gas through thermochemical processes, then pure hydrogen is obtained by separating hydrogen from this gas. In this method, hydrogen can be produced directly from biomass or intermediate products of biomass polymerization. As mentioned in the above lines, the chemical technology of hydrogen production from biomass includes the process of biomass gasification, pyrolysis-reforming, conversion in supercritical water, and catalytic-reforming of small organic molecules. Below is a brief description of each of these processes.

5.2.2.2.1

Pyrolysis Reforming

The word pyrolysis is made from the combination of two Greek words pyro (meaning fire) and lysis (meaning separation). Pyrolysis reforming is based on thermochemical process. In this process, biomass is decomposed into its components due to heat and in the absence of gasification agent. The advantage of the thermal decomposition process is the production of hydrogen with high purity. This process is irreversible. The reforming process can be done in five ways: water phase reforming method, photocatalytic reforming method, steam reforming method, chemical chain reforming method, and autothermal reforming method [28].

5.2.2.2.2

Supercritical Water Conversion

The meaning of conversion to gas in supercritical water conditions (T ≥ 374.2 °C, P ≥ 22.1 MPa) is that the desired biomass in the reactor is converted into hydrogenrich gas by water at a pressure and temperature higher than the supercritical point [29]. In this technology, due to the high solubility of water in the supercritical state, the biomass is hydrolyzed and finally turns into a gaseous state. Also, due to the high pressure of the process, separation becomes easy. In this technology, hydrogen production efficiency depends on temperature and pressure, reaction time, biomass concentration, oxidant concentration, and type of catalysts. This technology has a higher reaction efficiency than the conventional gasification process. Therefore, hydrogen production increases in this process and the cost of hydrogen generation is reduced and more affordable. The chemical reaction in supercritical water conversion technology is as follows [29]: CHn Om þ

1 n 1 H O→ þ H þ CO m 2 2 m 2

ð5:15Þ

120

5.2.2.2.3

M. Talaie et al.

CO þ H2 O → CO2 þ H2

ð5:16Þ

CO þ 3H2 → CH4 þ H2 O

ð5:17Þ

CO2 þ 4H2 → CH4 þ 2H2 O

ð5:18Þ

Biomass Gasification

The process of compressing and converting biomass raw materials into combustible gas containing hydrogen by a gas conversion agent (such as air and steam) is called the process of producing hydrogen through biomass gasification [30]. Therefore, gasification agents are needed in this process. The amount of hydrogen production using biomass to gas conversion technology depends on many indicators, including catalyst, selection of biomass raw materials, and gas conversion agent. Also, the intensity of the hydrogen production process is influenced by the temperature and the retention time of gas conversion [31]. The chemical reaction temperature in the hydrogen production method through biomass gasification is usually 1000–450 °C. In order to meet relatively high temperature conditions and purify the gaseous agent, huge energy consumption is required, which increases the cost [32]. There are many chemical reactions to produce hydrogen through biomass gasification, which are given as follows: C þ O2 → CO2

ð5:19Þ

2C þ O2 → 2CO

ð5:20Þ

C þ CO2 → 2CO

ð5:21Þ

C þ H2 O → CO þ H2

ð5:22Þ

C þ 2H2 O → CO2 þ 2H2

ð5:23Þ

CO þ H2 O → CO2 þ H2

ð5:24Þ

CH4 þ H2 O → CO þ 3H2

ð5:25Þ

The reactions show that H2, CO, CO2, CH4, and other small molecular carbides are produced during the hydrogen production process.

5.2.2.2.4

Catalytic Reforming of Small Organic Molecules

Biohydrogen production can create technical problems for hydrogenation stations such as pipeline transportation, hydrogen storage, and compression. In order to solve these problems, biomass can be converted into small molecular intermediates and they can be distributed directly to the hydrogenation stations [33]. In addition to

5

Hydrogenation from Renewable Energy Sources for Developing a Carbon-Free. . .

121

solving the mentioned problems, this process can reduce the final cost of hydrogen production and distribution. Also, catalytic reforming of small organic molecules is a very endothermic process and requires external heat. The advantages and disadvantages of hydrogen generation technologies from biomass are summarized in Table 5.2.

5.3

Applications of Hydrogen in Power and Energy Systems

Discussions about low-carbon energy systems often lead to issues such as flexibility and system integration. The category of flexibility is more related to energy supply. In an obvious example, fossil fuels are potential sources of energy stored in the layers of the earth, which can be used everywhere and at any time, and their high density can be used in all three states of gas, liquid, and solid if necessary. This feature allows them to be transported efficiently over long distances. This inherent feature provides great flexibility for the energy system. Meanwhile, in a low-carbon energy system based on RES, this time and place flexibility to adjust energy supply according to demand is limited. In terms of electricity production, renewable carriers are often provided by wind, bright light, sea waves, and wind sources. However, these resources do not response all the demands of subscribers at any time and place. This causes periods of surplus and deficit in supply to be different from one place to another. In addition, the fluctuation of output energy production as a result of weather variability can lead to fluctuations in supply. Since in the power grid, the required amount of energy must be supplied immediately and the balance between supply and demand must always be maintained, the fluctuation of energy production is considered a challenge. Many options are available to overcome the imbalance of variable power generation and consumption. Grid infrastructure, flexible load and generation, demand side management and energy storage can all be used to balance supply and demand. Hydrogen produced from electricity and water can be transported over long distances or stored in large quantities and converted into electricity or other energy. The stored hydrogen has many applications, including it can be injected into the natural gas network or converted into methane and sent to the power plant or sold to the transportation sector as fuel for fuel cell electric vehicles (FCVs). In this way, in a general view, stored hydrogen can be divided into major applications, some of which are mentioned in Fig. 5.5. On the other hand, fuel cells are a suitable alternative for consumers who used batteries to supply their needs. Specifically, the battery-powered electric vehicle, which was noticed in the past decades and efforts were made to commercialize it, has found a more suitable replacement today, and the new generation of transportation consists of fuel cell electric vehicles with clean fuels. Problems such as battery weight, maximum distance traveled, and recharging time on one hand and using unlimited hydrogen resources on the other hand changed the perspective of researchers working in this field from batteries to fuel cells.

122

M. Talaie et al.

Table 5.2 Advantages and disadvantages of any technologies of hydrogen generation from biomass Methods Renewable energy sources

Biohydrogen

Water photolysis

Photofermentation

Darkfermentation

Dark–photo co-fermentation

Advantage Production of hydrogen from water under standard temperature and pressure It is economically more affordable Using sunlight energy It has almost no environmental pollution Extensive raw materials It has almost no environmental pollution Extensive raw materials

Much simpler and cheaper than photofermentation reactors Flexibility in cultivation Stability of reliable hydrogen production Simultaneous production of hydrogen and consumption of organic waste No need for light It has almost no environmental pollution Extensive raw materials Higher efficiency Increasing the amount of production and reducing hydrogen production cost compared to the other two technologies It has almost no environmental pollution Extensive raw materials

Disadvantage Low efficiency Small volume of hydrogen production The sensitivity of enzymes involved in biological hydrogen production Long cycle of hydrogen production

The speed of this process is influenced by the pH value of the solution, the concentration of the substrate, the kind of fermentation microorganism, and the light intensity Low efficiency Low production capacity per unit of hydrogen production Need for a large area

It is not suitable for large-scale hydrogen production The cost of production is high Unstable hydrogen production rate

(continued)

5

Hydrogenation from Renewable Energy Sources for Developing a Carbon-Free. . .

123

Table 5.2 (continued) Methods Biomass gasification

Pyrolysisreforming

Supercritical water conversion

Catalyticreforming of small organic molecules

Advantage Extensive raw materials Low level of pollution

High purity hydrogen production No need for gasification agent Extensive raw materials Higher reaction efficiency compared to gasification process Higher production rate compared to gasification process Lower production cost compared to the gasification process Extensive raw materials Strong adaptability Extensive raw materials

Disadvantage It depends on the catalyst, raw materials, and gasification agent Requires a lot of energy Cost of production is high Remaining solid carbon The need for high temperature Complex reaction Hydrogen production efficiency depends on temperature and pressure, reaction time, biomass concentration, oxidant concentration, and catalyst High investment and maintenance costs

The process is very endothermic and requires external heat Hydrogen production rate depends on temperature and catalyst

In an ideal energy system based on green hydrogen with the aim of ensuring the security of energy supply, environmental protection and increasing the efficiency of the energy system, hydrogen is produced from electricity produced from RES such as wind and sun. This energy can be used after storage and transfer to the required places in various applications such as transportation industries and power plant industries, facilitating the exploitation of storage resources and their integration, methane gas production, and so on. Some of the most important applications of hydrogen in power and energy systems are given in the following subsections.

124

M. Talaie et al.

Fig. 5.5 Hydrogen and its different types of main applications

5.3.1

Facilitating the Operation of Renewable Resources and Their Integration

The dependence of the amount of renewable energy production on weather conditions, load changes, and faults in the lines can cause challenges in the analysis and planning of production and consumption of the system. In order to minimize these challenges during the process of planning, optimizing, and operating the system, the benefits of the energy storage system can be used to provide additional reserve capacity to cover such unforeseen fluctuations in production as well as errors occurring in the network. Therefore, the need to store electric energy with the aim of benefiting from its advantages and the stability of the electric system is evident. Considering the fuel value of hydrogen and its advantages, including the possibility of storing and distributing it in gas pipelines or retransformation into electrical energy by fuel cell, hydrogen can be considered as a sustainable source for renewable energy storage. These energy storage resources are very effective in integrating distributed generation units and are very valuable for power system management in order to increase security and reliability in operation. Also, energy storage resources help balance supply and demand, especially during peak consumption. Green hydrogen can be produced in the period of low load and through the excess energy of renewable energy production or production units with low pollution production. It is also possible to use the green hydrogen produced when the production rate decreases or during peak hours instead of using production units with high emissions. Therefore, this system has many technical and economic

5

Hydrogenation from Renewable Energy Sources for Developing a Carbon-Free. . .

125

advantages such as reducing the cost of operation, improving the reliability of the system, reducing peak consumption, and reducing the cost due to power outages. The storage and exploitation of green hydrogen reduces the intermittency of renewable resources and the use of battery storage systems, which impose a heavy cost on the power grid and have many disadvantages for the environment.

5.3.2

Reducing the Emission of Polluting Gases

The largest demand of primary energy in the world is related to electricity production and transportation. Most of the global demand for coal is related to electricity generation and most of the global demand for oil is related to transportation. Electric transportation, such as FCVs, can significantly reduce dependence on oil derivatives and air pollutant emissions and increase the use of RES. Likewise, RES are being developed to replace fossil fuel-based electricity production in order to reduce greenhouse gas emissions and other pollutants. Hydrogen production through the integration of electrolyzers and renewable energies significantly reduces the world’s dependence on fossil fuels, and as a result, greenhouse gas emissions are also reduced [34].

5.3.3

Fuel Cell

The use of hydrogen along with the development of fuel cells creates a very bright perspective for the energy sector. It is possible to use hydrogen in engines and fuel cells. In fuel cells, the chemical potential of hydrogen is directly converted into electrical energy and its by-product is water and heat. The efficiency of fuel cells is about three times compared to internal combustion engines, and the efficiency of fuel cells is about 43–63%. The applications of fuel cells are very wide. Compared to other power generators, fuel cells have many advantages, such as no air pollution, ability to operate at low pressure, high fuel efficiency, long life, high power density, reduction of noise pollution, absence of moving parts, and reduction of maintenance costs. The optimal performance conditions for a fuel cell depend on fuel purity, gas humidity, water management, the amount of air and fuel entering the fuel cell series, temperature control, and finally the gas pressure in the system and fuel cell series. Fuel cells are classified into the following types based on the type of electrolyte, operating temperature, type of fuel, and scope of application: • Polymer electrolyte fuel cell or proton exchange membrane fuel cell (PEMFC) • Alkaline fuel cell (AFC) • Phosphoric acid fuel cell (PAFC)

126

M. Talaie et al.

• Solid oxide fuel cell (SOFC) • Direct methanol fuel cell (DMFC)

5.3.4

Use of Green Hydrogen in the Transportation System

The possibility of storing excess electricity produced from renewable sources, through the production of hydrogen from electrolyzed water, provides the possibility of “load shifting” of electricity production. This makes it possible to use hydrogen as an energy carrier and makes it more widely used as a carbon-free fuel for land, air, and sea transportation. As an example, we can mention the use of green hydrogen in FCVs. These cars are a type of electric vehicle in which a fuel cell produces electricity through a chemical process [35]. FCVs have a (side) fuel tank such as natural gas or hydrogen. These cars can either rely entirely on fuel cells or be designed in combination with batteries such as HEV or PHEV. The future vision of the hydrogen economy requires the operation of FCVs in the transportation system. In other words, green hydrogen obtained from electrolysis of water is a potential for using renewable resources, and by using green hydrogen in FCVs, renewable energy sources can be used well.

5.4

Hydrogen Storage Technologies and Issues and Problems of Hydrogen Storage and Transfer

Two important indicators for using hydrogen as a fuel are specific energy and energy density. Compared to hydrocarbons, hydrogen has a suitable weight density, but an unfavorable volume density. Unfavorable volume density is the main problem in energy storage and transfer. To solve this problem and improve the volume density of hydrogen and use it as a fuel, the physical state of hydrogen must be changed. Therefore, one of the main challenges in the field of hydrogen is its storage. When choosing a hydrogen storage technology, various features should be considered. Among these features, we can mention safety, capacity, efficiency, durability, light weight, and costs of storage technology [36]. The relative importance of these indicators can be different according to the intended use of the system. Among these indicators, safety is of high priority because the flammability of hydrogen is very high and a small leak can lead to a major accident [37].

5.4.1

Storage Systems

Today, hydrogen storage systems are in the following forms:

5

Hydrogenation from Renewable Energy Sources for Developing a Carbon-Free. . .

• • • •

Hydrogen storage by gas or compression Hydrogen storage by liquefaction Hydrogen storage by physical absorption (physisorption) Hydrogen storage with the help of chemical absorption (chemisorption)

127

During hydrogen storage, about 15% of the total storage energy is used for hydrogen compression, and about 33–43% of it is used for hydrogen liquefaction. In super-cold systems where hydrogen is converted into liquid and stored, it must be completely isolated and must not have the least heat exchange with the surrounding environment. One of the problems of using liquid hydrogen is its liquefaction process, which requires a lot of energy. Due to the abundant use of liquid hydrogen, hydrogen liquefaction processes are very important. Another method of hydrogen storage is storage in solid format. Hydrogen can be stored as a hydride by combining with metals and metal alloys. Hydride storage tanks are usually used in small dimensions and mainly in the automotive industry. At the same time, the hydrides used must have multiple charge and discharge capabilities.

5.4.1.1

Hydrogen Storage by Gas or Compression

Hydrogen storage with the help of gas or condensation is the most common and simplest way to store natural hydrogen. This method has no evaporation loss. In this process, with the same technology that natural gas is compressed, hydrogen can also be compressed inside the tank. Hydrogen storage pressure is usually in the range of 20–25 MPa, and the study process indicates that in the future it will be possible to store up to 70 MPa inside the tanks [38]. Hydrogen is stored in a gaseous state in steel cylinders and very light composite tanks that can withstand high pressure. Cylinders should have high tensile strength and low density. It also does not react with hydrogen and does not release hydrogen from them. Most high-pressure cylinders are made of stainless steel, copper, or high-strength alloys such as titanium. Currently, composite tanks have many advantages, which are as follows: • The weight of composite tanks is much lighter than steel tanks. • Resistance to high pressures, around 350–700 bar. • These materials have high toughness and have very high safety at the moment of explosion, and it does not break into pieces when it explodes. The disadvantages of composite tanks are as follows: • It requires a lot of energy to store hydrogen with high pressure in cylinders. • They have little resistance to sunlight. • The high price of compactor and cylinder prevents this method from being commercially acceptable.

128

M. Talaie et al.

Hydrogen is compressed by devices that work mechanically. The amount of theoretical work required for the isothermal compression of hydrogen is obtained from Eq. (5.26). ΔG = R:TLn

P P0

ð5:26Þ

In this equation, R is the general gas constant, T is the absolute temperature, and P0 and P are the initial and final pressures. The advantages of storing hydrogen in the form of gas under pressure are as follows: • Compared to other methods, it has a simpler technology. • Hydrogen storage time is unlimited. • The degree of purity of hydrogen does not create a limitation in the storage process. One of the disadvantages of this process is that isothermal hydrogen compression requires a lot of energy. Another problem is the security of compressed gas cylinders. Cylinders can explode.

5.4.1.2

Hydrogen Storage by Liquefaction

Currently, Hydrogen is stored in liquid form at a temperature of 21.2 K and at ambient pressure in tanks. The critical temperature of hydrogen is 33.2 K and above this point there is no liquid hydrogen. Currently, there are several issues regarding hydrogen storage in liquid form, which are as follows: • In order to liquefy hydrogen gas, about 30–33% of the total energy of hydrogen is consumed. • Procurement of materials used in the construction of storage tanks is expensive. • Safety of storage tanks. • Evaporation loss in the storage tank (depending on the size of the tank, evaporation loss usually varies between 9.1% and 9% per day). Therefore, hydrogen storage by liquefaction method is often not suitable on a large scale and especially for mobile applications such as vehicles, although it occupies less space due to the high storage density.

5.4.1.3

Hydrogen Storage by Physical Absorption

Hydrogen can be molecularly attached to surfaces with a weak bond (mainly van der Waals), which is called physical adsorption. Hydrogen storage is possible through physical absorption because the absorbed gas is reversibly released. There are

5

Hydrogenation from Renewable Energy Sources for Developing a Carbon-Free. . .

129

different methods for physical adsorption according to the geometry of the absorber and the absorption temperature. Physical absorption is a type of absorption process in which the effective forces are intermolecular van der Waals forces, and as a result of this absorption process, there is no noticeable change in the pattern of the atoms involved. This type of absorption lacks activation energy. Therefore, due to the lack of activation energy, the speed of physical absorption is high. The energy released is 1–10 kJ/mol. Due to the fact that it has weak interactions, physical absorption is observed at temperatures lower than 273 K. In the physical absorption of hydrogen, due to the fact that the energy released in this process is not enough to separate the hydrogen molecule, the hydrogen molecule is not separated and the covalent bond remains intact. Materials that store hydrogen in this way are carbon materials, organometallic compounds, keratins, impregnated polymers, and so on. Some metals and alloys are able to absorb hydrogen and can desorb it by applying heat. These metals and alloys are known as metal hydrides. Due to high safety, smaller volume and relatively high density of hydrogen storage in metal hydrides, which is also associated with relative ease of hydrogen absorption, these materials are of great importance as a substrate for hydrogen storage. This process is considered to be one of the most common and simplest processes of hydrogen absorption. There are various metallic elements that are capable of storing hydrogen, but only a few of them can store hydrogen at the suitable temperature and pressure. Carbon-based materials (including activated carbon, carbon nanotubes, and fullerenes) are considered promising materials for hydrogen storage. Hydrogen storage with the help of physical absorption has many advantages. Among these advantages, performance at ambient pressure and temperature, high volumetric capacity, high storage efficiency, high purity of absorbed hydrogen, easy maintenance, solid state, safety, reversibility, ease of transportation, and ease of manufacturing can be mentioned.

5.4.1.3.1

Carbon Materials

Diamond and graphite are two allotropes of carbon. Each carbon atom in diamond is bonded to four other atoms and the hybridization of carbon atoms in this structure is SP3. In the graphic, regular carbon hexagons create layers. The layers are placed on top of each other and connected to each other by weak van der Waals bonds. Carbon hybridization is SP2 in graphic. Other forms of carbon such as fullerenes and nanotubes are the newest carbon structures with special properties. Activated carbons are also included in this category. These materials have shown the highest efficiency in hydrogen storage. (a) Fullerenes Fullerenes are a new class of aromatic carbon compounds with an unusual structure. The basis of fullerenes is the planes in the graphic. With the difference that in the atomic structure of fullerene, instead of the regular hexagons found in the graphic

130

M. Talaie et al.

pages, there are some hexagons and pentagons, which are placed together in the middle and form the fullerene sphere. The most stable and abundant fullerenes are C60 and C70. These materials can combine with hydrogen according to the following reaction: C60 þ xH2 O þ xe - → C60 Hx þ xOH -

ð5:27Þ

The most stable products of this reaction are C60H24, C60H36, and C60H48. Generally, the bonds between carbon and hydrogen are very stable and a temperature above 400 °C is necessary for desorption of hydrogen. (b) Carbon nanotubes Carbon nanotubes are classified into single-walled carbon nanotubes and multiwalled carbon nanotubes. A nanotube is actually a graphic that is in the form of a tube. When the graphic layers are intertwined, they form carbon nanotubes. There is only one graphite tube in a single-walled nanotube, but the multi-walled nanotube contains a number of concentric tubes. Nanotubes are hollow structures, so it is possible to put foreign materials inside them. Therefore, they can rapidly absorb hydrogen at room temperature and atmospheric pressure. Hydrogen is stored in nanotubes through physical absorption and electrochemically. Multi-walled carbon nanotubes absorb hydrogen between single-walled nanotubes, but hydrogen absorption causes the radius of the tubes to increase. Therefore, the multi-walled nanotube is unstable. The mass percentage of hydrogen stored in these nanotubes depends on the number of layers and the diameter of the tube. The results showed that carbon nanotubes can store relatively large amounts of hydrogen. (c) Activated carbon Bulk carbon with high surface area is called activated carbon. Activated carbon is a type of carbon structure that is able to absorb hydrogen in its microscopic pores. Pores are classified into three groups based on the pore size: macrometric pores, meso pores and micro pores. A total of 2.5 wt.% of hydrogen is absorbed in activated carbons at cooling temperature and pressure of 45–60 bar..

5.4.1.4

Hydrogen Storage with Chemical Absorption

Hydrogen can be absorbed atomically with a strong bond in hydride materials (alloys and metals), which is known as chemical absorption. In this type of absorption, the interaction between the absorber and the adsorbed is strong, which changes the nature of the absorber or the adsorbed, because it is possible to exchange electrons between the absorber and the adsorbed. These interactions are actually strong electrostatic attractions. Due to the strong interaction between the absorber and the adsorbed, chemical absorption releases more heat than physical absorption. In

5

Hydrogenation from Renewable Energy Sources for Developing a Carbon-Free. . .

131

chemical absorption, the heat of absorption is more than 20 kJ/mol. In this process, the activation energy is high. Molecules have the right time to be absorbed in their places, so there is selectivity in chemical absorption. In the chemical absorption of hydrogen, chemical interactions with the surface lead to the separation of the hydrogen molecule into atomic hydrogen and a surface bond with the absorber is formed. Chemical absorption of hydrogen is possible through two ways, active and passive. In the inactive path, the hydrogen molecule is separated spontaneously, but in the active path, the hydrogen molecule passes through the energy barrier before it is separated. Materials that store hydrogen in this way are metal hydrides, carbohydrates, ammonia, boraneamine complexes, formic acid, phosphonium borate, and imidazolium ionic liquid.

5.4.2

The Best Hydrogen Storage Technology

From the discussed topics, it appears that the method of storing hydrogen physically or chemically (metal hydrides) is the most suitable storage method. One of the most important reasons for using this method is the relatively high hydrogen storage capacity, decomposition temperature, and low heat of formation, which makes it easily formed and easily decomposed. The advantages of this technology include high safety and higher hydrogen storage density compared to gaseous and liquid hydrogen, lower operating and maintenance costs, and stability against water and oxygen. Comparing this method with other storage methods, it should be acknowledged that there is a possibility of hydrogen leakage in other hydrogen storage methods. Also, the storage capacity in other methods is low and in terms of safety, gas storage in these methods is very dangerous, but in the physical or chemical method, in addition to very high safety and very suitable storage capacity, the charging process can easily be done with replace charged rods with discharged rods.

5.4.3

Hydrogen Transmission and Distribution Technology

Hydrogen transport has a great impact on increasing the cost, distribution, and consumption of energy. Therefore, the methods used to transfer hydrogen will have significant technical and economic effects on the process of hydrogen production and consumption [39]. Three common methods for hydrogen transfer can be used, which are: • Transmission by pipelines • Transportation by road and rail • Transportation by sea

132

M. Talaie et al.

5.4.3.1

Transmission by Pipelines

It is possible to transfer hydrogen as gas or liquid in the pipe. The liquid hydrogen is transported for short distances in pipelines insulated with a vacuum jacket from the place of production to the place of consumption. However, these insulated pipes cannot be used for liquid hydrogen for long distances. The problem with hydrogen is the reactivity of hydrogen with metals, which causes cracking of metals. Another problem is the low volumetric energy density of hydrogen. In terms of volume, the energy density of hydrogen is only one-third of the energy density of natural gas. Therefore, hydrogen gas must be pumped three times as much as natural gas in a pipeline to transfer an equivalent amount of energy. Therefore, installing pumps and compressors is mandatory. In addition, all equipment must be improved to adapt to the unique characteristics of hydrogen. Hydrogen transportation through pipelines has been successfully carried out in several countries including France, the United States, and Canada.

5.4.3.2

Transportation by Road and Rail or by Sea

Although this method is a developed technique, it is not very useful and can only be used to transport gaseous hydrogen in small volumes. At present, small trucks transport liquid hydrogen in medium quantities on roads. Liquid hydrogen is usually transported by rail in tankers with a capacity of 9500–28,000 gallons. These tankers have insulation layers and vacuum jackets. Currently, there are ships dedicated to liquid hydrogen or ships carrying liquid natural gas tanks that transport all natural gas. Hydrogen is usually transported as a compressed gas or high-density liquid. Gaseous hydrogen is often transported by pipeline, but liquid hydrogen is transported by vehicle or by ships. In the transition process, all aspects of technology, cost, safety, reliability, and environmental impact should be evaluated. According to studies, hydrogen transportation by pipeline may be the cheapest method of transportation, but a high cost is required to build the pipeline. In other words, large-scale hydrogen transport by pipeline and small-scale hydrogen transport by road transport are suitable [39].

5.5

Green Hydrogen Standard

In technical terms, the development of guidelines, rules, and regulations in order to improve the quality and desirable characteristics of a product or service is called a standard. Compliance with standards is often voluntary, unless they are mandated as a requirement of law or regulation, or as part of a formal contract.

5

Hydrogenation from Renewable Energy Sources for Developing a Carbon-Free. . .

133

As mentioned, in recent years, green hydrogen has gradually attracted the attention of international communities. In global markets, hydrogen is considered as a commercial commodity and significant amounts of hydrogen are currently used in industry. Hydrogen can be produced from biological or non-biological raw materials and different energy paths (renewable or nonrenewable). According to this, hydrogen is named after the raw materials from which it is produced, renewable or nonrenewable production sources, and the process used to separate it. Among others, we can mention black hydrogen or brown hydrogen, which can be defined as hydrogen produced from coal through gasification. If in this process the primary source is natural gas, the produced hydrogen is called gray hydrogen. Also, if CO2 released in the gray hydrogen production process is absorbed and stored to a large extent (80–90%), the produced hydrogen is called blue hydrogen or low-carbon hydrogen. There are different viewpoints and various criteria to define green hydrogen. Different definitions for green hydrogen have also been proposed. Therefore, a global standard for green hydrogen is necessary to facilitate and increase global trade, create trust and confidence among investors, producers, and consumers, and increase investment [40]. Due to the lack of a global agreement on green hydrogen, an international standard for green hydrogen has not been developed yet. The lack of a uniform definition and a global standard makes international trade more difficult. In order to facilitate the development of global hydrogen trade, the desire to write national and transregional green hydrogen standards has increased. The development of national standards helps transregional standard organizations (such as ISO and IEC) in developing standards. The alignment between national and international standards will facilitate trade. In order to develop national or transregional green hydrogen standards, first of all, the definition of the structure, supply, and demand of hydrogen in the region or country should be defined. The main differences between green hydrogen standards often depend on the following factors.

5.5.1

Definition of Green Hydrogen

As mentioned, the definitions of green hydrogen are different regarding the type of raw materials, renewable or nonrenewable production resources, and production technology. Therefore, the following topics are usually discussed in the definition of green hydrogen: • It evaluates the type of hydrogen source based on renewable or nonrenewable production sources. • The role of the amount of greenhouse gas emissions in the definition of hydrogen is investigated. • The qualification of hydrogen production technology can be considered in the definition of hydrogen.

134

M. Talaie et al.

For example, the definitions of green hydrogen in some standards or organizations for issuing guarantee of origin certificates are as follows: • The term green hydrogen was first used in 1995 by the National Renewable Energy Laboratory (NREL). This reference considers the term green hydrogen as synonymous with the term renewable hydrogen (hydrogen produced from renewable energies). • The French Association of Hydrogen and Fuel Cells (AFHYPAC) has also defined hydrogen produced from RES as green hydrogen and considers it equivalent to renewable hydrogen. • CERTIFHY introduces the hydrogen obtained from any renewable path with a purity threshold of 99.5% or more as green hydrogen. Some standards have emphasized more on the reduction of greenhouse gas consumption and have proposed a broader definition of green hydrogen. Definitions based on the intensity of greenhouse gas emissions are more popular. • The European Union has provided the definition of green hydrogen standard based on CERTIFHY. In this standard, the emission threshold of greenhouse gases (at the POP) is proposed to be 36.4 gCO2e/MJH2 (CEN/CENELEC CLS JCT6). When this hydrogen is produced from nonrenewable energies by complying with this threshold, it is called low-carbon hydrogen. • The Green Hydrogen Organization (GH2) proposes hydrogen produced through electrolysis of water with 100% or near 100% renewable energy with nearly zero greenhouse gas emissions as green hydrogen. This standard sets a maximum threshold for greenhouse gas emissions of 1 kg CO2 per kg H2 for a 12-month period. Also, this standard focuses on the use of renewable energy technologies, which is the best option for the development of green hydrogen production [40]. • The TÜV SÜD CMS 70 standard identifies hydrogen from RES and/or waste as green hydrogen (TÜV SÜD, 2011b). According to the German TÜV SÜD CMS 70 standard, three processes are accepted in the production of green hydrogen [41]: – Water electrolysis using renewable electricity and greenhouse gas emission threshold of 83.8–89.7 CO2eq/MJ. – Steam reforming of biomethane with a greenhouse gas emission threshold of 83.8 gCO2eq/MJ. – Reforming and pyro-reforming of glycerin (when it is a by-product of biodiesel production) with a greenhouse gas emission threshold of 89.7 gCO2eq/ MJ [41]. • In 2020, the China Hydrogen Union defined low-carbon hydrogen, clean hydrogen, and renewable hydrogen based on the quality criteria of hydrogen production from different routes or sources. In this standard, renewable hydrogen is equivalent to green hydrogen. In this standard, the emission threshold of greenhouse gases is defined for clean hydrogen, low-carbon hydrogen, and renewable hydrogen or green hydrogen.

5

Hydrogenation from Renewable Energy Sources for Developing a Carbon-Free. . .

135

Table 5.3 System boundary CEN/

Chinese

TÜV SÜD

AFHYPAC

Low Carbon

CENELEC

standard

(Germany)

(France)

Standard

Fuel Standard

(international)

System

POU

California

GH2

POP

POP

POU

POP

(United Kingdom)

CERTIFHY POP

BEIS

POP

POP

boundary

5.5.2

System Boundary

The system boundary for calculating greenhouse gas emissions is defined in terms of hydrogen supply stages and the effect of these stages on greenhouse gas emissions. Based on the reduction of greenhouse gas emissions, different boundaries have been proposed for hydrogen. In the references, system boundaries are categorized into point of production (POP) and point of use (POU). The system boundary at the point of use is very broad and includes the calculation of downstream greenhouse gases such as transportation, storage system, supply, and fuel losses, and also includes the calculation of upstream greenhouse gases until its delivery to the final customer. In the system boundary at the point of production, the boundaries of greenhouse gas calculation are limited to the point of production only. In this border, due to the lack of calculation of downstream greenhouse gases, this system has a better performance. The boundaries of greenhouse gas calculation are limited to the point of production at the system boundary at the point of production. This system has a better performance due to the lack of calculation of downstream greenhouse gases. The system boundary in some standards or organizations of issuing a guarantee of origin is presented in Table 5.3.

5.5.3

Greenhouse Gases Emissions Threshold

Various interpretations have been given by standard organizations regarding the selection criteria for greenhouse gases emissions threshold, and there is no global consensus on it. The greenhouse gases emission threshold depends on the system boundary, hydrogen supply method, and consumption structure. Due to the different structures of hydrogen supply and consumption, the greenhouse gases emission threshold is also different. The selection criterion of threshold depends on the hydrogen supply structure when the system boundary is POP. When POU is defined as the system boundary, the choice of criterion depends on hydrogen consumption structure. Also, the standards differ according to the choice of a relative threshold (compared to the emission of fossil fuels) or an absolute threshold.

136

5.5.4

M. Talaie et al.

Qualification Level and Qualifying Technical Route

Global acceptance of renewable hydrogen is much higher than green hydrogen. Because determining the qualification level criteria as well as the qualifying technical paths of renewable resources is limited and excludes carbon capture and storage (CCS). Greenhouse gases emissions threshold, qualification level, and qualifying Technical Route in some standards are presented in Table 5.4.

Table 5.4 Greenhouse gases emissions threshold, qualification level, and qualifying Technical Route Reference AFHYPAC (France)

Qualification level Must be 100% renewable

CERTIFHY

At least 60% reduction of greenhouse gas emissions compared to hydrogen produced using SMR (this value is ≤36.4 gCO2e/MJ H2 for the last 12 months) It is equal to 83.8–89.7 gCO2e/MJ according to the production process

TÜV SÜD (Germany)

Greenhouse gases emissions threshold None

Greenhouse gas emissions from hydrogen produced through SMR of natural gas

Greenhouse gas emissions from hydrogen produced through SMR of natural gas Adapted from the standard Not determined

CEN/ CENELEC GH2

Adapted from the standard

California Low Carbon Fuel Standard

30% lower greenhouse gases and 50% lower NOX emissions for fuel cell electric vehicles

Well to wheel emissions from new gasoline vehicles

BEIS (United Kingdom) Cha (chain)

To be determined

Not determined

Emission of greenhouse gases less than 4.90 kgCO2eq/kgH2

Hydrogen produced from coal through gasification

Maximum 1 kg CO2 per kg H2 for a 12-month period

Qualifying Technical Routes Any technical pathway of hydrogen production from RES, including electrolysis powered by waste Hydrogen from any renewable pathway with a threshold of 99.5% purity or greater

Renewable electrolysis; biomethane SMR; pyroreforming of glycerin Adapted from the standard Electrolysis of water with 100% or near 100% renewable energy with nearly zero greenhouse gas emissions Renewable electrolysis, catalytic cracking of SMR of biomethane or production of hydrogen from biomass Neutral technology

The raw materials for hydrogen production should be derived from renewable energies

5

Hydrogenation from Renewable Energy Sources for Developing a Carbon-Free. . .

5.6

137

Conclusion

Considering the many advantages of hydrogen, including the high efficiency of energy conversion, the possibility of storage in different ways, the possibility of moving over long distances, the simplicity of conversion to other energies, the reversibility of its production cycle, and the reduction of the effects of greenhouse gases, being sustainable and renewable. It is expected that hydrogen will be established as an energy carrier in the coming years. Currently, for economic reasons, hydrogen production in the world is mainly based on fossil fuels and the conversion of hydrocarbons or the water splitting by these sources. These energy sources are nonrenewable and they harm the environment due to the production of large amounts of polluting gases. Therefore, this method is not suitable for hydrogen production and the need for a stable energy source with less pollution is felt more than ever. In an ideal hydrogen-based energy system, hydrogen production from RES and environmentally friendly methods should be developed in order to protect the environment and improve the efficiency of the system. The hydrogen produced based on this process is called green hydrogen. Among the green hydrogen production methods, we can mention biohydrogen and water splitting. The methods of hydrogen generation from water splitting are divided into three methods: electrolysis, thermolysis, and photocatalysis. The electrolysis method has advantages such as hydrogen produced with high purity (99.9%) and reducing CO2 emissions, benefiting from excess renewable energy and using cheap electricity. The most widely used electrolysis technologies are PEM, AWE, and SOEC. PEM electrolyzer is the most suitable option for green hydrogen production. Photocatalytic hydrogen from water splitting is a potential approach for green hydrogen production due to its low cost, low energy consumption, and environmental friendliness. The most abundant renewable resources for green hydrogen production are water and biomass. A suitable way to produce green hydrogen from RES is biohydrogen production technology through thermochemical or biological processes. Hydrogen production through biological process is classified into water photolysis, photofermentation, dark-fermentation, and dark–photo co-fermentation. Dark–photo co-fermentation technology has higher efficiency and lower production cost compared to the other two processes due to the use of unique advantages of hydrogen production methods in photo-fermentation and dark-fermentation processes. Also, chemical processes include methods of biomass gasification, pyrolysis reforming, supercritical water conversion, and catalytic-reforming of small organic molecules. Supercritical water conversion technology has a higher efficiency compared to the gasification process. Green hydrogen in power systems facilitates the exploitation of storage resources, their integration, and the reduction of polluting gas emissions. Also, by reconverting green hydrogen into electrical energy by fuel cell, hydrogen can be considered as a sustainable source for renewable energy storage. These storage resources help to

138

M. Talaie et al.

integrate distributed generation resource, increase network security, balance supply and demand, reduce the cost of operation, and improve the reliability of the system and peak shaving. Also, green hydrogen produced from water electrolysis is a potential for renewable resources that can be used as a carbon-free fuel for land, air, and sea transportation. One of the important challenges in the field of hydrogen is its storage. Today, hydrogen storage is possible through gas or compression, liquefaction, physisorption, and chemisorption processes. The gasification process is the most common and simple method of hydrogen storage. In hydrogen storage, about 15% of the energy is used for the compression process and about 33–43% is used for the hydrogen liquefaction process. In the liquefaction process, issues such as the high cost of materials used in the storage tank, the safety of the storage tanks and evaporation losses in the storage tank are discussed. Studies show that physisorption and chemisorption processes are the most suitable storage methods. Because these processes have high safety and higher hydrogen storage density and lower operating and maintenance costs compared to gaseous and liquid hydrogen. It is possible to transfer hydrogen in gas, liquid, and solid forms. In order to transfer hydrogen, it is possible to use pipelines, transportation by road and railway, and transportation by sea. Pipeline transportation is the most economical way to transport hydrogen, but pipeline construction costs are high. Transportation by road and railway is suitable for transporting hydrogen in small volumes and it is not very beneficial. Hydrogen is named according to the type of raw materials, renewable or nonrenewable production sources, and production process. Today, green hydrogen has attracted the attention of the world community. However, due to the lack of global agreement, an international standard for green hydrogen has not been developed yet. The main differences in the standards are often related to the definition of green hydrogen, system boundary, greenhouse gases emissions threshold, qualification level, and qualifying Technical Route. The boundary of the system is defined in terms of hydrogen supply stages and the impact of these stages on greenhouse gas emissions and is divided into two parts: POP and POU. The greenhouse gases emissions threshold depends on the system boundary, hydrogen supply method, and consumption structure. Due to the different structures of hydrogen supply and consumption, the greenhouse gases emissions threshold is also different.

References 1. Widera, B. (2020). Renewable hydrogen implementations for combined energy storage, transportation and stationary applications. Thermal Science and Engineering Progress, 16, 100460. 2. Davoodabadi, A., Mahmoudi, A., & Ghasemi, H. (2021). The potential of hydrogen hydrate as a future hydrogen storage medium. iScience, 24(1), 101907.

5

Hydrogenation from Renewable Energy Sources for Developing a Carbon-Free. . .

139

3. Bockris, J. O. M. (2013). The hydrogen economy: Its history. International Journal of Hydrogen Energy, 38(6), 2579–2588. 4. Worighi, I., Maach, A., Hafid, A., Hegazy, O., & Van Mierlo, J. (2019). Integrating renewable energy in smart grid system: Architecture, virtualization and analysis. Sustainable Energy, Grids and Networks, 18, 100226. 5. Ji, M., & Wang, J. (2021). Review and comparison of various hydrogen production methods based on costs and life cycle impact assessment indicators. International Journal of Hydrogen Energy, 46(78), 38612–38635. 6. Dinçer, İ., & Zamfirescu, C. (2016). Sustainable hydrogen production. Elsevier. 7. Chi, J., & Yu, H. (2018). Water electrolysis based on renewable energy for hydrogen production. Chinese Journal of Catalysis, 39(3), 390–394. 8. Kumar, S. S., & Himabindu, V. (2019). Hydrogen production by PEM water electrolysis – A review. Materials Science for Energy Technologies, 2(3), 442–454. 9. Bhattacharyya, R., Misra, A., & Sandeep, K. C. (2017). Photovoltaic solar energy conversion for hydrogen production by alkaline water electrolysis: Conceptual design and analysis. Energy Conversion and Management, 133, 1–13. 10. Wappler, M., Unguder, D., Lu, X., Ohlmeyer, H., Teschke, H., & Lueke, W. (2022). Building the green hydrogen market – Current state and outlook on green hydrogen demand and electrolyzer manufacturing. International Journal of Hydrogen Energy, 47(79), 33551–33570. 11. Kojima, H., Nagasawa, K., Todoroki, N., Ito, Y., Matsui, T., & Nakajima, R. (2023). Influence of renewable energy power fluctuations on water electrolysis for green hydrogen production. International Journal of Hydrogen Energy, 48(12), 4572–4593. 12. Marocco, P., Ferrero, D., Lanzini, A., & Santarelli, M. (2021). Optimal design of stand-alone solutions based on RES+ hydrogen storage feeding off-grid communities. Energy Conversion and Management, 238, 114147. 13. Jang, D., Kim, J., Kim, D., Han, W. B., & Kang, S. (2022). Techno-economic analysis and Monte Carlo simulation of green hydrogen production technology through various water electrolysis technologies. Energy Conversion and Management, 258, 115499. 14. Yao, Y., Gao, X., & Meng, X. (2021). Recent advances on electrocatalytic and photocatalytic seawater splitting for hydrogen evolution. International Journal of Hydrogen Energy, 46(13), 9087–9100. 15. Baykara, S. Z. (2004). Experimental solar water thermolysis. International Journal of Hydrogen Energy, 29(14), 1459–1469. 16. Zhang, Z., Li, Y., Zhang, H., He, C., & Zhang, Q. (2017). Potential use and the energy conversion efficiency analysis of fermentation effluents from photo and dark fermentative bio-hydrogen production. Bioresource Technology, 245, 884–889. 17. Kumar, R., Kumar, A., & Pal, A. (2021). An overview of conventional and non-conventional hydrogen production methods. Materials Today: Proceedings, 46, 5353–5359. 18. Etemoglu, A. B., & Can, M. (2007). Classification of geothermal resources in Turkey by exergy analysis. Renewable and Sustainable Energy Reviews, 11(7), 1596–1606. 19. Guban, D., Muritala, I. K., Roeb, M., & Sattler, C. (2020). Assessment of sustainable high temperature hydrogen production technologies. International Journal of Hydrogen Energy, 45(49), 26156–26165. 20. Wright, L., & Boundy, B. (2006, September). Biomass energy data book (1st ed.). Oak Ridge National Laboratory Office of Planning, Budget and Analysis, Energy Efficiency and Renewable Energy, US Department of Energy. 21. Milbrandt, A. (2005). Geographic perspective on the current biomass resource availability in the United States (No. NREL/TP-560-39181). National Renewable Energy Lab. (NREL). 22. Shi, Y., Yang, A. F., Cao, C. S., & Zhao, B. (2019). Applications of MOFs: Recent advances in photocatalytic hydrogen production from water. Coordination Chemistry Reviews, 390, 50–75. 23. Çelik, D., & Yıldız, M. (2017). Investigation of hydrogen production methods in accordance with green chemistry principles. International Journal of Hydrogen Energy, 42(36), 23395–23401.

140

M. Talaie et al.

24. Argun, H., & Kargi, F. (2011). Bio-hydrogen production by different operational modes of dark and photo-fermentation: An overview. International Journal of Hydrogen Energy, 36(13), 7443–7459. 25. Akhlaghi, N., & Najafpour-Darzi, G. (2020). A comprehensive review on biological hydrogen production. International Journal of Hydrogen Energy, 45(43), 22492–22512. 26. Kumar, G., Sivagurunathan, P., Pugazhendhi, A., Thi, N. B. D., Zhen, G., Chandrasekhar, K., & Kadier, A. (2017). A comprehensive overview on light independent fermentative hydrogen production from wastewater feedstock and possible integrative options. Energy Conversion and Management, 141, 390–402. 27. Zhang, T., Jiang, D., Zhang, H., Jing, Y., Tahir, N., Zhang, Y., & Zhang, Q. (2020). Comparative study on bio-hydrogen production from corn Stover: Photo-fermentation, dark-fermentation and dark-photo co-fermentation. International Journal of Hydrogen Energy, 45(6), 3807–3814. 28. Zeng, L., & Gong, J. L. (2015). Advances in chemical looping reforming for direct hydrogen production. CIE Journal, 66(8), 2854–2862. 29. Research progress on catalysts for hydrogen production from biomass supercritical water gasification. (2013). Chemistry Research Applied, 25, 7–12. 30. Zhang, Y., Xu, P., Liang, S., Liu, B., Shuai, Y., & Li, B. (2019). Exergy analysis of hydrogen production from steam gasification of biomass: A review. International Journal of Hydrogen Energy, 44(28), 14290–14302. 31. Amin, M., Shah, H. H., Fareed, A. G., Khan, W. U., Chung, E., Zia, A., Farooqi, Z. U. R., & Lee, C. (2022). Hydrogen production through renewable and non-renewable energy processes and their impact on climate change. International Journal of Hydrogen Energy, 47(77), 33112–33134. 32. Correa, C.R. and Kruse, A., 2018. Supercritical water gasification of biomass for hydrogen production–Review. The Journal of Supercritical Fluids, 133, 573–590. 33. Xu, D., Xiong, Y., Zhang, S., & Su, Y. (2019). The influence of preparation method of char supported metallic Ni catalysts on the catalytic performance for reforming of biomass tar. International Journal of Energy Research, 43(13), 6922–6933. 34. Richardson, D. B. (2013). Electric vehicles and the electric grid: A review of modeling approaches, impacts, and renewable energy integration. Renewable and Sustainable Energy Reviews, 19, 247–254. 35. Luo, Z., Hu, Y., Xu, H., Gao, D., & Li, W. (2020). Cost-economic analysis of hydrogen for China’s fuel cell transportation field. Energies, 13(24), 6522. 36. Luta, D. N., & Raji, A. K. (2018). Decision-making between a grid extension and a rural renewable off-grid system with hydrogen generation. International Journal of Hydrogen Energy, 43(20), 9535–9548. 37. Mohseni, S., & Brent, A. C. (2020). Economic viability assessment of sustainable hydrogen production, storage, and utilisation technologies integrated into on-and off-grid micro-grids: A performance comparison of different meta-heuristics. International Journal of Hydrogen Energy, 45(59), 34412–34436. 38. Nazir, H., Muthuswamy, N., Louis, C., Jose, S., Prakash, J., Buan, M. E., Flox, C., Chavan, S., Shi, X., Kauranen, P., & Kallio, T. (2020). Is the H2 economy realizable in the foreseeable future? Part II: H2 storage, transportation, and distribution. International Journal of Hydrogen Energy, 45(41), 20693–20708. 39. Witkowski, A., Rusin, A., Majkut, M., & Stolecka, K. (2017). Comprehensive analysis of hydrogen compression and pipeline transportation from thermodynamics and safety aspects. Energy, 141, 2508–2518. 40. GH2, The Green Hydrogen Standard, 2022. 41. TÜV SÜD. (2011). TÜV SÜD standard: Generation of green hydrogen. TÜV SÜD Industrie Service GmbH.

Chapter 6

The Role of Green Hydrogen in Achieving Low and Net-Zero Carbon Emissions: Climate Change and Global Warming Mohammad Shaterabadi

6.1

, Saeid Sadeghi, and Mehdi Ahmadi Jirdehi

Overview and General Background Outlook

The primary culprit behind climate change and global warming is widely believed to be greenhouse gas (GHG) emissions [1]. This action can be attributed to the world’s heavy reliance on fossil fuels for over 80% of its energy needs, resulting in a significant increase in GHG emissions released into the atmosphere over the last few decades [2]. One of the most significant obstacles in achieving a harmonious balance between environmental considerations and economic and technological requirements is the finite and nonrenewable aspect of fossil fuels, compounded by the rapidly increasing demand for energy in today’s society [3]. The Paris Agreement has established a crucial objective of limiting the increase in average global temperature to below 2 degrees Celsius (°C), with an ideal target of 1.5 °C. This goal is vital in mitigating and reducing the harmful effects of global warming. Therefore, achieving a climate-neutral world by the mid-century and a zero net carbon economy by 2050 can become more feasible if the aforementioned target is met [4]. A host of complex challenges has long impeded the attainment of a net carbon society. In order to advance toward this goal, research and analysis are being conducted to identify and overcome these obstacles [5]. The shift toward sustainable and renewable energy sources is critical in establishing a viable and livable environment for future generations [6]. Generally, low-net-zero emission systems (LNZE) are energy systems that emit minimal or no greenhouse gases [7]. Ongoing research is being conducted on related topics, such as the technologies, markets, and policies involved in systems utilizing renewable energy resources. One such resource is green hydrogen (GH2), which has M. Shaterabadi · S. Sadeghi · M. A. Jirdehi (✉) Department of Electrical Engineering, Kermanshah University of Technology, Kermanshah, Iran e-mail: [email protected] © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 V. Vahidinasab et al. (eds.), Green Hydrogen in Power Systems, Green Energy and Technology, https://doi.org/10.1007/978-3-031-52429-5_6

141

142

M. Shaterabadi et al.

Fig. 6.1 The assumption pathway of hydrogen from past to future Table 6.1 Produced hydrogen and its usage based on the investigation [15] Sources Natural gas Coal Oil Water Utilizing of hydrogen Power Industry Heat Transportation

Percentage of total produced hydrogen (%) 48 30 18 4 Percentage of total usage (%) 18.3 9.3 12.5 46.9

shown promise as a dependable and sustainable energy source [8]. Despite this potential, the research on GH2 is still in its early stages [9]. Hydrogen is a promising fuel option due to its nonpolluting attributes, which align intending to establish a carbon-free world [10]. Its widespread availability, reversible production cycle, and ability to reduce greenhouse gases are among the advantageous qualities that make hydrogen a noteworthy and viable fuel alternative [11]. The hydrogen energy system is a reliable and sustainable energy source, as it does not rely on primary energy sectors. Hydrogen can be produced in various forms, such as gray, blue, green, pink, yellow, or turquoise, depending on the production method [12]. However, it is essential to note that only green hydrogen (GH2) generation can effectively contribute to attaining a low net-zero carbon society by 2050 [13]. Figure 6.1 supports this assertion, showing that green hydrogen is critical for achieving a sustainable energy transition and global low net-zero emission economies [14]. The tabulated data in Table 6.1 highlights the variable quantities of hydrogen generated from different inlet fuels. Notably, natural gas yields the highest percentage of hydrogen, albeit with associated harmful emissions and negative contributions to global warming. The data shows that the transportation industry is a significant consumer of hydrogen produced. Future projections have outlined the path toward achieving ultimately environmentally friendly hydrogen.

6

The Role of Green Hydrogen in Achieving Low and Net-Zero Carbon Emissions:. . .

6.2

143

Low and Net-Zero Emission Clarification

In simplified terms, the concepts of low and net-zero emissions pertain to the amount of carbon dioxide and other greenhouse gases released into the atmosphere by all sources, including those from land, water, or air. The presence of greenhouse gases, such as methane and carbon dioxide, is the primary contributor to the increase in global temperatures. Countries and organizations worldwide are endeavoring to reduce these emissions to the fullest extent possible. This objective may be accomplished through various means, including enhancing energy efficiency, adopting renewable energy sources, and investing in sustainable technologies. In order to ensure that global warming does not exceed 1.5 or 2 °C, it is imperative to achieve zero-CO2 emissions and eliminate all global greenhouse gases by either 2050 or the latter half of the century. This action can be accomplished by implementing zero greenhouse gas emissions systems, effectively reducing them by eliminating them naturally, or utilizing innovative technologies to balance the remaining greenhouse gas emissions. For instance, Azevedo et al. [16] examined models of integrated global evaluation and national-level studies in more detail about industries, technology, time, and place. The concept of “zero greenhouse gas emissions systems” has been defined in various ways, and these definitions hold significant implications for policy and planning. The inclusion or exclusion of particular sectors of the economy can impact the timeline for attaining “zero emissions.” It is essential to recognize that achieving zero emissions necessitate substantial changes in energy systems. Specifically, there would be a need to shift demand, behavior, and performance and foster the development of new technologies and fuel sources. Policies targeting emission reduction must address various issues, including diverse impacts and varying areas of emphasis. Although there may be regional and national variances, fundamental characteristics can be observed in zero greenhouse gas emission systems, which are explicated in this section. Moreover, specific features referenced herein have also been identified in prior research on reducing greenhouse gases at a lower percentage. Nevertheless, the disparities between 80% and 100% greenhouse gas percentages in zero-emissions systems may be more significant. Azevedo et al. investigated that each carbon dioxide (CO2) emission must be balanced by absorbing the same amount of CO2 in zero-emissions systems. A practical approach to achieving zero emissions with fossil fuels is using carbon dioxide capture and storage (CCS) technology. With recent advancements, CCS has become a highly efficient solution for reducing emissions. Moreover, energy systems that achieve zero emissions require significantly less fossil fuels. However, the specific amount of fossil fuels needed depends on various factors such as the cost of biotechnology, hydrogen, and electricity fuels, the cost and scope of CCS implementation, and the cost and breadth of carbon dioxide removal (CDR) within the energy system. The demand for coal worldwide remains insufficient in systems that emit no pollutants unless significant progress is made in

144

M. Shaterabadi et al.

CCS. In such systems, the demand for fossil fuels is expected for gasoline and gas, owing to their high energy density and utility as raw materials. However, the precise quantity of liquid and gas fuels required in zero-emission systems is uncertain and varies depending on the type of usage such as electrical or other low-carbon fuel. This underscores the importance of energy policies and reduced greenhouse gas emissions, particularly in hydrogen technology. The findings suggest that different policies, such as carbon budgets and taxes, can achieve zero carbon dioxide emissions at varying costs. Due to its versatility across various industries, hydrogen technology can play a vital role in the zero carbon dioxide emissions system. Policymakers must cautiously develop energy-efficient policies and evaluate their impact on consumers and industries.

6.2.1

Barriers and Required Infrastructure for Reaching Low-Net-Zero Emission

Achieving zero emissions by 2050 is ambitious and requires significant investments in technology and infrastructure. However, the high costs associated with these investments can challenge progress toward this objective. In order to achieve lownet-zero emissions, countries must prioritize investment in renewable energy sources such as wind and solar power, as well as energy storage systems [17]. This procedure requires substantial financial resources, land, and other essential resources. While renewable energy sources are becoming increasingly efficient and cost-effective, significant infrastructure investments are still necessary to maximize their potential due to their limited efficiency and cost-effectiveness [18]. The forecast for renewable energy sources is encouraging. The shift toward low-net-carbon emissions entails modifying behavior, including reducing energy consumption, embracing energyefficient products, and implementing renewable energy instead of nonrenewable sources. However, this transformation can present challenges as specific individuals may exhibit resistance to change and may not possess the resources to invest in the necessary infrastructure [19]. The attainment of the intended goals can be daunting, considering the considerable political and economic hurdles that must be overcome.

6.2.2

How Penalties and Taxes on Available Technology Can Lead to Low-Net-Zero Emission

Carbon dioxide and other greenhouse gas emission reduction can be accelerated by implementing taxes and penalties. As a result of human activities, these emissions occur and can be curbed by imposing taxes on carbon emissions [20]. These actions would increase the cost of fossil fuels, inducing people to shift toward renewable

6

The Role of Green Hydrogen in Achieving Low and Net-Zero Carbon Emissions:. . .

145

energy sources and consequently reducing their carbon footprint. Governments can also levy financial penalties on companies that produce high levels of greenhouse gases [21]. One potential solution to reducing emissions is incentivizing companies to become more cost-effective in their emissions release. The government can provide subsidies to companies that invest in renewable energy sources and energy storage systems, making long-term planning for low-net-zero emissions more economically feasible. This affordable technology can enable companies to transition to this state quickly [22]. Furthermore, the government can offer subsidies to individuals who invest in energy-efficient products such as appliances and cars, encouraging them to focus less on energy-intensive products in the future [23]. Numerous experts assert that implementing penalties and taxes can substantially contribute to transitioning from a high-emitting economy to a low-emitting one. Governments can support establishing vital infrastructure by investing in energy-efficient products to attain zero emissions. Furthermore, they can foster an atmosphere that encourages individuals and businesses to invest in renewable energy sources by increasing pollution costs and promoting sustainable alternatives [24].

6.3

Road to Carbon Neutrality

In order to achieve carbon neutrality, a transition from traditional fossil fuels toward sustainable and renewable energy sources is necessary. This approach requires a comprehensive strategy that involves capturing, storing, and utilizing carbon and implementing economic incentives. Such measures are essential to promote a more environmentally friendly and sustainable future.

6.3.1

Carbon Capture, Storage, and Utilization Versus Pollution Emitting

The technology of carbon capture, storage, and utilization (CCUS) has emerged as a promising strategy for mitigating the greenhouse effect and reducing CO2 emissions. CCUS entails capturing and storing carbon dioxide (CO2) before it is released into the atmosphere, as illustrated in Fig. 6.2. While this approach holds potential for the future, it should not be viewed as a substitute for reducing pollution from other sources. The most effective way to reduce CO2 emissions is by decreasing the consumption of fossil fuels for energy generation. This step can be accomplished using renewable energy sources such as solar, wind, and geothermal. Moreover, implementing energy efficiency measures, such as enhancing insulation and using more energy-efficient appliances, can help decrease the energy required to power homes and businesses [25].

146

M. Shaterabadi et al.

Fig. 6.2 The CCUS technology and various absorption processes and utilization benefits

6.3.2

Economic and Environmental Aspects of Carbon Neutrality

Carbon neutrality requires a significant investment of economic resources from business entities and governmental bodies. In order to accomplish this, it is essential to invest in innovative technologies such as CCUS and renewable energy sources. Furthermore, policies such as carbon taxes, emissions trading, and subsidies may be necessary to encourage individuals and companies to decrease their carbon footprints. However, maintaining a balance between the environmental benefits of carbon neutrality and the associated economic costs is of utmost importance [26]. Reducing CO2 emissions can prevent further environmental changes and positively impact humans and other species. Research indicates that transitioning from fossil fuels to renewable energy sources can reduce air and water pollution, improving public health. In order to safeguard the environment and ensure a sustainable future for all, achieving carbon neutrality has become a top priority. Therefore, governments and investors must invest in green technologies and policies to reduce emissions and transition to a more sustainable energy system [27].

6.4

Green Hydrogen Explanation and Definition

The production of GH2, or green hydrogen, is widely recognized as a highly effective method for reducing carbon dioxide emissions and generating clean energy. Various renewable energy sources such as wind, hydroelectric, and solar

6

The Role of Green Hydrogen in Achieving Low and Net-Zero Carbon Emissions:. . .

147

power can be utilized to produce electricity to create this sustainable energy source. GH2 boasts a multitude of applications, including powering buildings, generating electricity, and fueling vehicles. GH2 is environmentally friendly and renewable from renewable sources such as wind turbines or solar panels. The production of green hydrogen is expected to experience significant growth in the future but further research, development, and increased production are required to mitigate its cost. Produced through renewable technologies such as water electrolysis, green hydrogen represents a potential global economic competitor and can meaningfully contribute to renewable energy production [28]. In contrast, gray hydrogen and hydrogen produced from fossil fuels have the highest greenhouse gas emissions. Blue hydrogen, which utilizes carbon absorption and storage technologies, can lead to reduced greenhouse gas emissions [29]. Ultimately, the production of green hydrogen is considered a secure and clean solution for hydrogen production and other applications [30]. In the most recent explanations, Fig. 6.3 illustrates the diverse forms of hydrogen generated through disparate techniques.

Fig. 6.3 Diverse forms of hydrogen produced in different manners and their usage

148

6.4.1

M. Shaterabadi et al.

How Can We Pass Through Different Types of H2 to GH2

In various instances, GH2 can act as a viable alternative to other forms of hydrogen. However, several nations face the challenge of insufficient access to reliable renewable energy, making producing GH2 a daunting task. Furthermore, sophisticated electrolysis technology is necessary to efficiently separate water molecules into oxygen and hydrogen atoms. To transition from other variants of hydrogen to GH2, it is imperative to invest in renewable energy sources such as wind and solar power. This approach would enable nations to generate more renewable energy and utilize it to manufacture GH2. Furthermore, investing in advanced electrolysis technology can improve productivity, enhance efficiency, and reduce the production costs of GH2 [31].

6.4.2

Status Quo, Challenges, and Outlook to Achieving and Developing GH2: Infrastructure, Limits, and Policies

One of the primary obstacles to advancing the development of GH2 as an energy source is the absence of sufficient infrastructure to support its production, storage, transportation, and utilization. Consequently, the potential of GH2 is somewhat limited. To fully realize the benefits of GH2, a substantial infrastructure must be established. Furthermore, existing policies and regulations in many countries impose limitations on the utilization of GH2, impeding its expansion and progress. It is imperative to update these policies and regulations to enable GH2 to achieve its full potential and overcome these obstacles. Finally, the production cost of GH2 remains relatively high compared to other energy sources, posing a challenge for countries to invest in it without a guaranteed return on investment [24, 32].

6.4.3

The Intermittency Effect of Renewable Units for Transition to GH2

It is critical to consider that renewable energy sources are subject to variability, which challenges the widespread adoption of GH2. Nevertheless, this obstacle can be surmounted through investments in energy storage technologies. Due to fluctuations in weather conditions, renewable energy production may be inconsistent, which can lead to difficulties in maintaining steady GH2 output. Energy storage technology can stabilize renewable energy units and preserve their output, facilitating consistent GH2 production [33]. Furthermore, governments must allocate resources toward grid

6

The Role of Green Hydrogen in Achieving Low and Net-Zero Carbon Emissions:. . .

149

modernization and smart grid technologies to ensure the efficient and reliable use of renewable energy sources. Investing in these technologies can produce more efficient and dependable GH2 for nations [34].

6.5

How Can GH2 Help to Achieve Low and Net-Zero Emissions?

A viable solution to address the issue of carbon emissions and counteract climate change is adopting green hydrogen (GH2) as a clean energy alternative. GH2 is generated through electrolysis, which involves breaking down water molecules into hydrogen and oxygen through electrical power sourced from sustainable resources such as solar and wind energy. This hydrogen gas can be applied across various domains, such as electricity production, transportation, and industrial activities, leading to a considerable reduction in greenhouse gas emissions. The production of carbon-neutral hydrogen from renewable energy sources can constitute a significant milestone in the sphere of the circular economy. This effort becomes particularly pertinent in light of the mounting global carbon footprint and the imperative to sever the link between economic growth and carbon emissions. The production of green hydrogen can assume a pivotal role in the renewable energy landscape and help curtail greenhouse gas emissions in the future.

6.5.1

Utilizing GH2 in the Residential and Mobility Sectors: GH2 Route in Energy Sectors

One viable approach to reducing emissions in both residential and mobility sectors is implementing GH2 as an alternative energy source. Specifically, utilizing GH2 in residential buildings presents a promising solution to curbing pollution and greenhouse gas emissions [35]. Substituting natural gas, oil, and other fossil fuels with GH2 can significantly reduce residential emissions. Furthermore, GH2 can serve as a potential power source for vehicles, offering the opportunity to make substantial progress in mitigating public health risks and diminishing emissions [36].

6.5.2

Future Scopes and Effects of GH2 on Pollution Emission

The production cost of GH2 has been declining, making it a formidable contender against fossil fuels in the race to reduce global carbon emissions. GH2 has the potential to decrease carbon emissions in the future significantly and can serve as

150

M. Shaterabadi et al.

a renewable energy storage solution and a grid battery. By enlisting more countries and businesses to join the movement, we can maximize the impact of global carbon reduction efforts. GH2’s benefits extend beyond environmental concerns, as it can aid in electrifying factories and other industrial sites, effectively eliminating emissions. Furthermore, it can facilitate the transition to renewable energy sources, further reducing emissions from the electricity sector.

6.6

Renewable-Based Units’ Role in Smoothing the Way to Reach Low-Net-Zero Emission: Clean and Adequate GH2 Generation

In the context of combating climate change and global warming, the utilization of green hydrogen (GH2) has become increasingly significant in achieving low-carbon emissions and net-zero carbon emissions. GH2 is a form of energy produced from renewable sources such as solar, wind, geothermal, and biomass, making it a clean and carbon-free alternative. Due to its versatility, GH2 can be utilized in numerous applications such as heating, transportation, and industry, making it a crucial element in transitioning to a low-carbon economy. Renewable-based units have become essential in producing clean and sufficient GH2 to achieve low-net-zero emissions [37]. These units, such as wind farms and solar panels, can also enhance the stability of the power grid and reduce the cost of GH2 production. Proper installation and management of these units are vital in ensuring a clean supply of GH2. Anticipating future demands and deploying renewable-based energy units in optimal locations is also necessary. Developing smart grids and storage solutions, such as hydrogen fuel cells, further aid in achieving low net greenhouse gas emissions while providing future energy storage opportunities [38].

6.7

Numerical Analysis of GH2

For this purpose, various scenarios have been considered. Here is a simplified numerical analysis to demonstrate the potential impact of hydrogen (H2) on reducing greenhouse gas (GHG) emissions in the context of transportation and power generation. Please note that these numbers are for illustrative purposes. • Scenario: Hydrogen fuel cells vs. internal combustion engines (ICE). Assumption Comparing hydrogen fuel cell vehicle (FCVs) emissions to traditional internal combustion engine vehicles (ICEVs). Emissions from Hydrogen Fuel Cells Assume an FCV powered by a hydrogen fuel cell has an efficiency of 60% and produces only water vapor as emissions. Emissions = 0 g CO2/km.

6

The Role of Green Hydrogen in Achieving Low and Net-Zero Carbon Emissions:. . .

151

Emissions from Internal Combustion Engine Assume an ICEV has a fuel efficiency of 10 km/L and emits 2.3 kg of CO2 per liter of gasoline. Emissions = (10 km/L) × (2.3 kg CO2/L) = 23 g CO2/km. Result By switching from ICEVs to hydrogen fuel cell vehicles, emissions could be reduced from 23 g CO2/km to 0 g CO2/km, indicating a significant reduction in direct emissions. • Scenario: Hydrogen-enhanced power generation vs. coal power generation. Assumption Replacing a portion of coal in a power plant with hydrogen to reduce CO2 emissions. Emissions from Coal Power Generation Assume a coal-fired power plant emits 900 g CO2/kWh. Emissions = 900 g CO2/kWh. Emissions from Hydrogen-Enhanced Power Generation If 20% of the coal is replaced by hydrogen with an efficiency of 45%, the emissions reduction can be calculated as follows: Emissions = (0.20 × 45%) × (900 g CO2/kWh) = 162 g CO2/ kWh. Result By integrating hydrogen into power generation, emissions can be reduced from 900 g CO2/kWh to 162 g CO2/kWh, showcasing a potential reduction in CO2 emissions.

6.8

Summary and Conclusion

Achieving low and net-zero carbon emissions is a critical objective that requires implementing innovative solutions. Green hydrogen, which is produced using renewable energy sources such as wind and solar, has emerged as a promising option for achieving this goal. The utilization of green hydrogen in the transportation, industrial, and residential sectors has the potential to significantly reduce emissions and facilitate the transition toward a more sustainable future. Therefore, the efficient use of green hydrogen is essential in the global effort to combat climate change and reduce carbon emissions. In conclusion, addressing the current issue of carbon emissions necessitates a comprehensive approach that involves policy reform, community initiatives, and targeted educational campaigns. Furthermore, any solution must be tailored to meet the specific needs and context of the community in which it is intended to be implemented. Ultimately, the success of any strategy depends on stakeholders’ ability to collaborate and develop a comprehensive and sustainable solution that addresses everyone’s concerns.

152

M. Shaterabadi et al.

References 1. Firoozi, H., Khajeh, H., Laaksonen, H., & Shaterabadi, M. (2022). Active building as an electricity network service provider. In Active building energy systems: Operation and control (pp. 273–293). Springer. 2. Jirdehi, M. A., & Shaterabadi, M. (2021). A low-carbon strategy using INVELOX turbines in the presence of real-time energy price uncertainty. Greenhouse Gases: Science and Technology, 11(3), 461–482. 3. Shaterabadi, M., Mehrjerdi, H., Amiri, N., Jirdehi, M. A., & Iqbal, A. (2023a). Green roof positive impact on changing a plus-ZEB to drastic plus-ZEB: The multi-objective energy planning and audit in real condition outlook. Energy and Buildings, 287, 112983. 4. Jirdehi, M. A., & Shaterabadi, M. (2020). Incentive programs caused by the carbon capture utilization and storage technology profit’s effect: Optimal configuration and energy planning of hybrid microgrid involving INVELOX turbine. Energy Technology, 8(10), 2000398. 5. Jirdehi, M. A., Shaterabadi, M., Tabar, V. S., & Jordehi, A. R. (2022). Impact of diverse penetration levels of thermal units on a hybrid microgrid energy management considering the time of use and function priority. Applied Thermal Engineering, 217, 119164. 6. Shaterabadi, M., & Jirdehi, M. A. (2020). Multi-objective stochastic programming energy management for integrated INVELOX turbines in microgrids: A new type of turbines. Renewable Energy, 145, 2754–2769. 7. Shaterabadi, M., Jirdehi, M. A., Amiri, N., & Omidi, S. (2020). Enhancement the economical and environmental aspects of plus-zero energy buildings integrated with INVELOX turbines. Renewable Energy, 153, 1355–1367. 8. van Renssen, S. (2020). The hydrogen solution? Nature Climate Change, 10(9), 799–801. 9. Sorgulu, F., & Dincer, I. (2022). Analysis and techno-economic assessment of renewable hydrogen production and blending into natural gas for better sustainability. International Journal of Hydrogen Energy, 47(46), 19977–19988. 10. Reigstad, G. A., Roussanaly, S., Straus, J., Anantharaman, R., de Kler, R., Akhurst, M., Sunny, N., Goldthorpe, W., Avignon, L., & Pearce, J. (2022). Moving toward the low-carbon hydrogen economy: Experiences and key learnings from national case studies. Advances in Applied Energy, 8, 100108. 11. Shaterabadi, M., Khajeh, H., Firoozi, H., & Laaksonen, H. (2022b). Energy management systems of grid-connected active buildings. In Active building energy systems: Operation and control (pp. 251–271). Springer. 12. Raman, R., Nair, V. K., Prakash, V., Patwardhan, A., & Nedungadi, P. (2022). Green-hydrogen research: What have we achieved, and where are we going? Bibliometrics analysis. Energy Reports, 8, 9242–9260. 13. Yadav, Y. (2022). Green hydrogen-fuel for the future. School of Petroleum Management. 14. Shaterabadi, M., Mehrjerdi, H., & Jirdehi, M. A. (2023b). How INVELOX can affect the perspective of renewable energy exploitation: Demand response and multilateral structure planning outlook. Sustainable Cities and Society, 91, 104421. 15. Li, X., Raorane, C. J., Xia, C., Wu, Y., Tran, T. K. N., & Khademi, T. (2023). Latest approaches on green hydrogen as a potential source of renewable energy towards sustainable energy: Spotlighting of recent innovations, challenges, and future insights. Fuel, 334, 126684. 16. Azevedo, I., Bataille, C., Bistline, J., Clarke, L., & Davis, S. (2021). Net-zero emissions energy systems: What we know and do not know. Energy and Climate Change, 2, 100049. 17. Shaterabadi, M., & Jirdehi, M. A. (2021). Smart scheduling of transmission line switching: Optimization of multi-objective microgrid’s day-ahead energy scheduling with considering high penetration of green energies and INVELOX. Electrical Engineering, 103, 1753–1767. 18. Amiri, N., Shaterabadi, M., Reza Kashyzadeh, K., & Chizari, M. (2021). A comprehensive review on design, monitoring, and failure in fixed offshore platforms. Journal of Marine Science and Engineering, 9(12), 1349.

6

The Role of Green Hydrogen in Achieving Low and Net-Zero Carbon Emissions:. . .

153

19. Shaterabadi, M., Jirdehi, M. A., & Amiri, N. (2022a). Impact of fuel price increasing policies on fuel smuggling and acceptance of green cars: A case study of Iran. The Journal of Engineering, 2022, 1239. 20. Quarton, C. J., & Samsatli, S. (2021). How to incentivise hydrogen energy technologies for net zero: Whole-system value chain optimisation of policy scenarios. Sustainable Production and Consumption, 27, 1215–1238. 21. Razi, F., & Dincer, I. (2022). Challenges, opportunities and future directions in hydrogen sector development in Canada. International Journal of Hydrogen Energy, 47, 9083. 22. Mneimneh, F., Ghazzawi, H., Abu Hejjeh, M., Manganelli, M., & Ramakrishna, S. (2023). Roadmap to achieving sustainable development via green hydrogen. Energies, 16(3), 1368. 23. Nickels, T. (2022). Hydrogen-a zero-emission solution. MTZ Worldwide, 83(12), 80–80. 24. Rabiee, A., Keane, A., & Soroudi, A. (2021b). Technical barriers for harnessing the green hydrogen: A power system perspective. Renewable Energy, 163, 1580–1587. 25. Nandhini, R., Sivaprakash, B., Rajamohan, N., & Vo, D.-V. N. (2022). Carbon-free hydrogen and bioenergy production through integrated carbon capture and storage technology for achieving sustainable and circular economy – A review. Fuel, 342, 126984. 26. Fazeli, R., Beck, F. J., & Stocks, M. (2022). Recognizing the role of uncertainties in the transition to renewable hydrogen. International Journal of Hydrogen Energy, 47(65), 27896–27910. 27. Maynard, I., & Abdulla, A. (2023). Assessing benefits and costs of expanded green hydrogen production to facilitate fossil fuel exit in a net-zero transition. Renewable Energy Focus, 44, 85–97. 28. Muradov, N. (2017). Low to near-zero CO2 production of hydrogen from fossil fuels: Status and perspectives. International Journal of Hydrogen Energy, 42(20), 14058–14088. 29. Rabiee, A., Keane, A., & Soroudi, A. (2021a). Green hydrogen: A new flexibility source for security constrained scheduling of power systems with renewable energies. International Journal of Hydrogen Energy, 46(37), 19270–19284. 30. Zhang, Y., Davis, D., & Brear, M. J. (2022). The role of hydrogen in decarbonizing a coupled energy system. Journal of Cleaner Production, 346, 131082. 31. Newborough, M., & Cooley, G. (2021). Green hydrogen: The only oxygen and water balanced fuel. Fuel Cells Bulletin, 3, 16–19. 32. Sadik-Zada, E. R. (2021). Political economy of green hydrogen rollout: A global perspective. Sustainability, 13(23), 13464. 33. Tiwari, A. (2021). The emerging global trends in hydrogen energy research for achieving the net zero goals. Advanced Materials Letters, 12(10), 1–5. 34. Younas, M., Shafique, S., Hafeez, A., Javed, F., & Rehman, F. (2022). An overview of hydrogen production: Current status, potential, and challenges. Fuel, 316, 123317. 35. Oliveira, A. M., Beswick, R. R., & Yan, Y. (2021). A green hydrogen economy for a renewable energy society. Current Opinion in Chemical Engineering, 33, 100701. 36. Dash, S. K., Chakraborty, S., Roccotelli, M., & Sahu, U. K. (2022). Hydrogen fuel for future mobility: Challenges and future aspects. Sustainability, 14(14), 8285. 37. Devbalan, I., & Yadav, A. (2022). Green hydrogen from green electricity. In 2022 advances in science and engineering technology international conferences (ASET) (pp. 1–3). IEEE. 38. Dong, Z. Y., Yang, J., Yu, L., Daiyan, R., & Amal, R. (2022). A green hydrogen credit framework for international green hydrogen trading towards a carbon neutral future. International Journal of Hydrogen Energy, 47(2), 728–734.

Chapter 7

Bioreactor Design Selection for Biohydrogen Production Using Immobilized Cell Culture System Nur Kamilah Abd Jalil , Umi Aisah Asli , Haslenda Hashim , Mimi Haryani Hassim, Norafneza Norazahar, and Aziatulniza Sadikin

7.1

Introduction

Generally, biohydrogen fermentation can be accomplished using the batch or continuous mode. In the early stages of optimization of biohydrogen fermentation, a batch fermentation is more suitable. However, continuous fermentation is preferred for commercial purposes. Fermentation begins when microorganisms or bacteria produce enzymes to break complex (two molecules) sugars into single-molecule sugars, then convert single-molecule sugars into commercial chemicals, desired products, and by-products. A continuous stirred tank reactor (CSTR) is preferable due to its simplicity of construction, effective mixing, and easy operation. In addition, the CSTR controls the microbial growth rate by maintaining a specific hydraulic retention time (HRT) [1]. The other types of bioreactors include the anaerobic sequencing batch reactor (ASBR) [2], membrane bioreactor (MBR) [3], carrier-induced granular sludge bed reactor (CIGSB) [4], agitated granular sludge

N. K. A. Jalil · M. H. Hassim · A. Sadikin Faculty of Chemical & Energy Engineering, Universiti Teknologi Malaysia, Johor Bahru, Johor, Malaysia U. A. Asli (✉) Chemical Reaction Engineering Group (CREG), Faculty of Chemical & Energy Engineering, Universiti Teknologi Malaysia, Johor Bahru, Johor, Malaysia e-mail: [email protected] H. Hashim Green Energy and Environmental Planning (GREEN), Faculty of Chemical & Energy Engineering, Universiti Teknologi Malaysia, Johor Bahru, Johor, Malaysia N. Norazahar Centre of Hydrogen Energy (CHE), Universiti Teknologi Malaysia, Johor Bahru, Johor, Malaysia © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 V. Vahidinasab et al. (eds.), Green Hydrogen in Power Systems, Green Energy and Technology, https://doi.org/10.1007/978-3-031-52429-5_7

155

156

N. K. A. Jalil et al.

bed reactor (AGSBR) [5], fixed bed [6] or packed bioreactor (PBR) [7], fluidized bed bioreactor (FBR) [7], and upflow anaerobic sludge blanket bioreactor (UASB) [8]. Immobilized cell culture is often used as an alternative to free cell culture in biohydrogen production in the lab to enhance microorganisms’ activity [5]. This is due to its advantages in repeatability [9], overcoming washout problems, requiring less space [10], tolerance to environmental distraction [11], and reducing the lag phase of bacterial cultivation [12]. Previous studies have discovered that higher biohydrogen could be yielded through continuous bioreactors with immobilized culture. The immobilized culture system provides maximum physical retention of microbial biomass [13] with minimal mass transfer effects without significantly increasing the energy demand for fluidization [14]. The bioreactor system for immobilized culture may utilize natural flocs or granules [15], self-immobilized microbes [16], microbial immobilization concerning inert materials [12], microbial-based biofilms, or retentive membranes [17]. According to Show et al. [18], the bioreactors’ choice depends on the feedstock’s nature, which could be converted into organic acids, alcohols, and biogas with the help of microorganisms. However, the immobilized culture has restricted usage on specific bioreactor designs due to some limiting factors that could affect the operation or production yield [19], which cannot be avoided such as clogging [20] and metabolic limitation [21]. Hence, this chapter discusses different designs of bioreactors that are suitable for immobilized culture in generating biohydrogen based on work done by several researchers. This chapter also reviews the important factors to consider, such as substrate type and microorganism, pH, temperature (T), hydraulic or solid retention time (HRT/SRT), and the bioreactor configuration.

7.2

Immobilization of Microbial Culture in Biohydrogen Production

Immobilization is one method of creating a stable environment for rapid growth and metabolism. Generally, the surface attachment method has gained more attention in biohydrogen production [22]. The immobilization of culture via surface attachment method can be achieved via entrapment, adsorption, encapsulation, and containment with synthetic polymers.

7.2.1

Cell Entrapment

Microorganisms are trapped in porous environments (Fig. 7.1) to facilitate the diffusion of cell substrates and the removal of cell by-products. Natural and synthetic polymers such as agar, alginate, lignocellulosic materials, polyvinyl alcohol, polyacrylamides, and ethylene-vinyl acetate are used as support materials [19] because

7

Bioreactor Design Selection for Biohydrogen Production Using. . .

Fig. 7.1 Cell entrapment method

157

Microbial cell

Porous matrix

Fig. 7.2 Immobilization using the adsorption method

Microbial cell

Support material

they are not toxic to microorganisms. Synthetic polymers are preferable because their mechanical strength and durability are better than natural polymers. For example, Singh et al. [11] immobilized Clostridium sp. LS2 cell by entrapment method into polyethylene glycol (PEG) prepolymer. The medium culture was heated for 3 min at 80 °C before use. The PEG and N,N-methylene bisacrylamide crosslinkers were dissolved in water and mixed with inoculum. The polymerization was initiated as potassium persulfate (K2S2O8) was added, and the mixture could stand for 20 min for bead formation. Then, polymerized substances were cut into 3 mm beads. The highest hydrogen production obtained using beads prepared onto 10% (w/v) PEG was 0.38 L H2/g COD. In another study, Wu and Chang [4] used polymeric materials consisting of polymethyl methacrylate (PMMA), collagen, and activated carbon to entrap biomass for hydrogen production. The favorable conditions using PMMA-immobilized cells were 35 °C, pH 6, and 20 g COD/sucrose, with a hydrogen production rate of 238 mL/h/L. It was investigated that the yield obtained was considerably high at 2.68 mol bioH2 mol/sucrose but can still be improved by avoiding the accumulation of gas products that may inhibit or be unfavorable to the H2-producing kinetic. It was suggested that the pores of immobilized cells could be appropriately adjusted to increase mechanical strength and stability to allow long-term operation [4].

7.2.2

Adsorption

This method allows the surface of the support matrix to attach to the microbial cells for better mass transfer efficiency and higher biomass retention capacity, improves substrate utilization at short hydraulic retention time, and produces more stable biohydrogen production, as shown in Fig. 7.2. The substrate conversion efficiency

158

N. K. A. Jalil et al.

can be improved as the nutrient directly interacts with immobilized cells. The cells were spread into porous carrying materials or embedded in the gel-forming polymer in this method. Patel et al. [13] used lignocellulosic materials of banana leaves, coconut coir, groundnut shells, or pea shells as support materials packed into polyvinylchloride (PVC) tubes. The tube was tied with a 10 cm2 nylon net, and inoculum was individually added with different microbial mixed cultures at 10 μg cell protein/ mL of medium. The bioH2 yield achieved about 1.54–1.65 mol/mol glucose. In another study, Chang [6] used a loofah sponge for the immobilized matrix to generate up to 190 mL of biohydrogen from activated sewage sludge. Among several immobilization carriers, activated carbon has been recognized as an effective support material used in fermentation. Different sponges, such as plastic scouring, plastic nylon, and black porous, were also used to immobilize mixed culture for biohydrogen fermentation [23]. The immobilized coculture on the loofah was less efficient than activated carbon (AC) [16].

7.2.3

Encapsulation

Encapsulation involves retaining cells in a permeable membrane, allowing nutrients to diffuse, as illustrated in Fig. 7.3. The cultured cells may be encapsulated within a polymer matrix during synthesis or post-synthesis. By this method, the problem of cell leakage, contamination, and inhibitory material can be prevented or at least reduced. A previous study reported that encapsulation increased biohydrogen yield by at least 2.7 folds compared to suspended cells [19]. Another study on cell entrapment was applied by Zhao et al. [10], where the immobilized Clostridium sp. T2 was encapsulated in a calcium alginate bead. In their study, an inoculum was mixed with autoclaved sodium alginate in anaerobic conditions before being pumped through a sterilized syringe into a cold presterilized CaCl2 solution to form beads (2.5–3.0 mm diameter). The maximum hydrogen production rate obtained from the study was 2.76 mmol bioH2/L/h, more than a 40% increase compared to the carrier-free process.

Fig. 7.3 Method of immobilization by encapsulation

Microbial cell Porous matrix

7

Bioreactor Design Selection for Biohydrogen Production Using. . .

Fig. 7.4 Method of containment within synthetic polymers

159

Microbial cell

Polymer gel

7.2.4

Containment Within Synthetic Polymers

In this method, the bacteria or cells are trapped inside a support material such as synthetic polymers, including polyacrylamide, polyurethane, polyvinyl alcohol, polyethylene glycol, and polycarbamoyl sulfonate, which could be used as support material. In one interesting study, Barros et al. [23] successfully obtained a maximum bioH2 yield of 2.25 mol bioH2/mol of glucose by using anaerobic sludge as inoculum, particles of the ground tire, and polyethylene terephthalate (PET) as support material for immobilization. This method is illustrated in Fig. 7.4. As a concluding remark, the most crucial factor to be considered for a carrier of the immobilizer is its suitability for biohydrogen-producing microorganisms to be dominant in the process [24]. As for that, the carrier must be non-toxic to the microorganisms and have a large surface area for the cell to adhere. The carrier should have good mechanical, chemical, and thermal stabilities and resistance to any inhibitors. It is possible to recycle the waste materials and scale up the process using these materials.

7.3

Types of Bioreactors

Similarly to other fermentations, biohydrogen fermentation is also possible in batches or continuously. Batch mode fermentations are more suitable for the initial phase of immobilization development studies. However, in most cases, industrial processes require continuous or semicontinuous operation (fed or sequential batch). The typical bioreactors that are suitable for immobilized culture include continuous stirred tank reactor (CSTR), upflow anaerobic sludge bioreactor (UASB), fluidized bed reactor (FBR), and fixed/packed bed reactor (PBR).

N. K. A. Jalil et al.

160

7.3.1

Bioreactor for Batch System

Many studies on hydrogen production fermentation were conducted in batch systems due to its simple operative and easy-to-control processes. A batch system was commonly implemented for the laboratory scale to ease monitoring. It mainly aims to identify the conditions to enhance hydrogen production by substrate [20] or bacteria culture selections [25] and adjustment of operation parameters [26]. Usually, the operative parameters controlled are less compared to the continuous system. The significant parameters to be controlled in a batch fermentation are pH, T, substrate volume, inoculum percentage, and fermentation time. Conceptually, the bioreactor in a batch system is alternately fed with fresh media to increase broth media volume with time. On the other hand, in a fed-batch system, part of the broth is drained at the end of the cycle and fed in a recycle batch. It was called a repeated fed-batch, where the residual broth from the previous batch serves as inoculum for the next fed-batch. This method significantly saves labor costs and time [27]. A stirred tank reactor (STR) is a type of bioreactor that was popularly used in batch, semi-batch, and continuous systems. For batch and semi-batch systems, the culturing is usually conducted at a thermophilic condition with a pH range of 5.5–6.5 at a minimum HRT of 24 h and a maximum of 21 days [28] to achieve better production. The usage of this simple design did not receive popular demand due to poor settling characteristics and its sensitivity to inconsistent operating conditions [20]. The use of STR over the years identified some problems and drawbacks, which had researchers to explore other designs or types of bioreactors. There are some disadvantages of STR, such as biomass washout caused by poor settling characteristics and short solid or hydraulic retention times [19], and occasionally fluctuating high sensitivity conditions such as pH, T, and HRT [19]. Figure 7.5 displays a typical schematic diagram of STR that is commonly used in a laboratory scale: (a) batch, (b) semi-batch, and (c) continuous systems.

Influent In luent

Influent In luent

Stirrer/agitator

Effluent luent Rotation pump (a) Fig. 7.5 Typical CSTR diagram

(b)

7

Bioreactor Design Selection for Biohydrogen Production Using. . .

Biogas collection

161

Sample collection

Connect to computer Gas collection tube Biogas

pH and

temperature probe

Stirring magnetic bar

Food waste Water bath

Fig. 7.6 Benchtop bioreactor [31]

In a study by Sekoai et al. [30], a batch fermentation was conducted to examine the biohydrogen production from organic solid waste using immobilized anaerobic sludge at pH 7.9 and 30.3 °C for 90 h. This study proved that immobilized bacteria in calcium alginate effectively increased the biohydrogen fraction by 4.21% in a batch system of a semi-pilot scale run with an agitation speed of 100 rpm. This study successfully obtained a biohydrogen (bioH2) yield of about 215.39 mL bioH2/ g TVS. Besides that, a simple experimental setup by Xiao et al. [31] is illustrated in Fig. 7.6 for batch fermentation generating bioH2 [31]. Culture bottles were used as fermenters using a magnetic stirrer for mixing. Escherichia cloacae and Enterobacter aerogenes acted as inoculums. The fermentation was carried out for 48 h, at pH 7 and 37 °C, with biohydrogen obtained at approximately 155.2 mL/g of VS in food waste. On the other hand, another batch setup by Kao et al. was carried out to study the effects of free and immobilized cells on bioH2 production [12]. Clostridium butyricum, the common biohydrogen producer bacteria, was first cocultured with Rhodopseudomonas palustris, which was then immobilized in alginate beads to ferment sucrose at 37 °C and pH 6.5 for up to 118 h of fermentation. The bioH2 production of 604 mL was achieved using the immobilized cell coculture. This amount was significantly reportedly higher than the production gained using free cell coculture.

7.3.2

Bioreactor for Continuous System

Most bioreactors can support a continuous system as it has an inflow of medium into the bioreactor and an outflow. The most influential significant factor is the hydraulic retention time. A specific dilution rate must be maintained to avoid significant

162

N. K. A. Jalil et al.

Table 7.1 Comparison of different selected continuous bioreactors [33] Type of bioreactor CSTR UASB

Advantages Simple process, easy to operate Good retention biomass

FBR

Good retention of biomass No clogging Good mass transfer due to efficient mixing

PBR

No need for mechanical mixing Good retention of biomass

Granular bioreactor or immobilized cell-seeded bioreactor

Excellent biomass retention (allows very high loading and short HRT) Rapid sludge granulation (short start-up time) Maximum space available for biomass (no or low amount of carrier)

Disadvantages Low biomass retention Slow development of granules (long start-up periods) Instability of H2 production Volume occupied by carrier (less volume available for biomass) Strong shear forces can detach biomass Energy needed for biomass fluidization Clogging Lower mass transfer than FBR Gas hold-up Volume occupied by carrier (less volume available for biomass) Poor mass transfer Channeling of flow and formation of gas pockets (if no mixing)

problems of biomass washout [9], cells, and microbes [32]. Various carbohydratebased feedstock or substrates that aid the inoculum (sludge or hydrogen-producing bacteria) are summarized in this review. A summary by Rahma [33] on the benefits and drawbacks of commonly used bioreactors is enlisted in Table 7.1. (a) CSTR As previously mentioned, this bioreactor can be operated continuously with an organic loading rate of up to 20 g substrate/L/h [34]. As the substrate concentration was kept constant, the organic loading rate values decreased exponentially by increasing HRT in the reactors [35]. The performance was highly dependent on the HRT and substrate concentration (g/COD), which both factors could influence the mass transfer in the bioreactor. Generally, short HRT can cause biomass washout, while broth agitation could affect stirred tanks’ mass transfer coefficients [10]. Since mixed operating patterns could disrupt fermentation, the immobilized technique has been applied to enhance biomass retention. Figure 7.7 displays the diagram of an immobilized CSTR bioreactor. Besides the commonly used bacterial-based inoculum, Han et al. [36] also used granular activated carbon as a support medium for cell immobilization with the mixture of H2-producing sludge. The effect of organic loading rate (OLR) was

7

Bioreactor Design Selection for Biohydrogen Production Using. . .

Feed pump

163

Agitator

Biogas meter

Reactor

Water lock Waste water box Fig. 7.7 Immobilized CSTR used for biohydrogen production [34]

adjusted by immobilizing the sludge with granular activated carbon to a range of 8–24 kg/m3 at a mesophilic temperature of 36 °C. The CSTR system produced the highest bioH2 production at OLR of 24 kg/m3 with VSS 17.74 g/L [36]. The result indicated that OLR is also an important parameter to be considered to enhance the bioH2 production rate (10.74 mL/h/L). Meanwhile, Zhao et al. [10] reported their study on immobilized Clostridium sp. in CSTR. The concentration of substrate used was 10 g/L at 37 °C, pH 7, and 200 rpm (agitation speed) for 10 h of retention time. This study utilized xylose as a substrate with immobilized Clostridium sp. T2 using two different techniques: one entrapped in sodium alginate and the other adhered to mycelia pellets. Results indicated that the entrapment techniques in CSTR are more efficient in improving biohydrogen production by a rate of more than 40.8% higher than carrier-free cells [10]. In another study on CSTR, Hu et al. [31] used cheese whey and glucose as substrates. This work emphasized the utilization of activated sewage sludge as hydrogen (H2)-producing bacteria immobilized in calcium alginate beads. Fermentation was conducted at 35 °C, pH 5–8 for 2 days at 150 rpm. The production rate of bioH2 was significantly dependent on the substrate concentration in which the maximum bioH2 production was obtained at 35 g COD/L whey with 27 mL bioH2/g COD. (b) UASB UASB is an extensively applied anaerobic treatment system with high efficiency and a short hydraulic retention time (HRT). This reactor can maintain high concentrations of large granules with high bioactivity for efficient operation. UASB was usually used for biogas production from wastewater effluent in laboratories or pilot-scale studies [22]. However, it has also effectively treated organic wastes and biohydrogen production. The main structure of the reactor includes a dense sludge bed located at the bottom [37]. The active biomass is retained in the reactor. It is suspended by recirculating part of the effluent to the bottom of the reactor while keeping the upward velocity of the mixed liquor in the range of 0.5–3 m/h [38]. A complete UASB system for lab scale use is illustrated in Fig. 7.8. On the upper part

N. K. A. Jalil et al.

164 Fig. 7.8 Schematic representation of a UASB reactor [8]

Biogas Biogas

Biogas bubble

Granular

Inlet

Fig. 7.9 Schematic of complete UASB reactor system

Gas/liquid separator

Thermal insulating

Gas meter

Blanket Zone Bed zone Temperature controller

Sampling ports Pump Substrate tank

of the reactor, just above the sludge bed, is a blanket zone where some biomass particles are suspended (Fig. 7.9). This zone acts as a separation zone between the water flowing and suspended biomass [39]. According to Graaff et al. [8], UASB reactors enable long sludge retention times (SRT) at relatively short hydraulic retention times (HRT) because an internal gas/sludge/liquid separation system accomplishes biomass retention. In addition, one of the main advantages of UASB is the low sludge production and the ability to maintain a high concentration of biomass inside the digester [40]. According to Singh et al. [42], immobilized culture is also recommended for UASB using POME (20 g COD/L) as a substrate and immobilized seed sludge in PEG. They also reported that immobilized cells could produce better bioH2 production at lower HRT (2 h) than suspended cells in UASB, at 0.589 and 0.348 L bioH2/ L, respectively. Hu [28] ran UASB with chloroform-treated granules into calcium alginate using a similar pH and T as in the previous study. The review summarized that maximum bioH2 production occurred at HRT 13–5.3 h for 3 days.

7

Bioreactor Design Selection for Biohydrogen Production Using. . .

165

outlet

Gas collector

Fluidized bacterial granural bed

Effluent connecting pipe to

Gas outflow

gas disengager

Water jacket Activated carbon Water jacket

Influent inlet

Bed of glass bead Pump

Fig. 7.10 Schematic diagram of modified FBR

(c) Fluidized Bed Reactor (FBR) FBR is a combined packed bed and stirred tank reactor, where the cells are attached to solids as biofilm and granules. It has excellent heat mass transfer characteristics and is flexible to work at low and high HRT. Therefore, this reactor became more efficient due to less chance of biomass and cell washout compared to UASB [7]. According to Abu Rahma, FBR has good biomass retention due to its efficient mixing and prevention of clogging. However, bioH2 production is still unstable since the carrier mostly occupies the volume of biomass. Hence, activated carbon (AC) in the shape of cylindrical pellets for bacterial granulation, as displayed in Fig. 7.10, with modified FBR at HRT 10 h and thermophilic condition (65 °C), was designed for better performance [17]. The modified FBR enabled the production of bioH2 at approximately 7.57 L bioH2/L/min with a bioH2 content of 69% at HRT 1 h [17]. However, Balachandar et al. [5] and Abu Rahma [33] detected a drawback in this system where the reactor requires more energy to achieve fluidization as the shear solid force required to detach the biomass. (d) Fixed/Packed Bed Reactor (PBR) This reactor is completely packed with carrier materials that hold biomass retention, where the retention mass transfer is considered good [33] for bioH2 production. Usually, mass transfer resistance will lower the production and substrate conversion. This is due to the low turbulence of the hydraulic mixing regime [6], which simultaneously affects the pH gradient distribution along the reactor column and causes the heterogeneous distribution of microbial activity [18]. Therefore, this

166

N. K. A. Jalil et al.

Fig. 7.11 Immobilized bioreactor with ceramic ball. (Adapted from [41])

Gas measurement

Outlet tank

Immobilized reactor

Inlet tank

Peristaltic pump

Magnetic stirrer

reactor used recirculation flow to overcome the bottlenecks. However, the reactor has the potential to clog if run continuously. In principle, high turbulence mixing favors mass transfer and lowers the gas hold-up [5], but the mass transfer is still recorded to be lower than FBR. Among other continuous bioreactors, PBR is also called an immobilized reactor. Keskin et al. [41] reported the use of ceramic balls as a cell immobilization material in comparison with CSTR, which indicated improvement in hydrogen production by immobilization. As displayed in Fig. 7.11, the biohydrogen could be efficiently produced using ceramic balls in an immobilized system. The system was proven to work at thermophilic conditions, with a pH of 5.5 and HRT of 3 h, producing 2.7 L bioH2/L/day. The work showed that an immobilized reactor is more robust than CSTR, with minimal leaching or washout problems. Moreover, PBR helps to hold the growing bacteria for a longer time [33] and maintain high cell concentration [27] to improve the efficiency and stability of H2 production. FBR can be considered as the economically feasible way for biohydrogen production as it requires minimal biomass or substrate pretreatment [1].

7.3.3

Bioreactor Performance

The optimum conditions for immobilized microbial activity are critical for improving biohydrogen production in different bioreactors. For the batch system, fermentation time is the limiting factor to obtain maximum hydrogen production depending on the H2-producing bacteria used. Meanwhile, HRT and OLR are important parameters besides the T and pH to obtain optimal production using immobilized bioreactors in continuous processes. Selected bioreactors for biohydrogen production using immobilized culture are tabulated in Table 7.2. Among the different types of bioreactors, UASB has been commonly used to resolve problems faced in CSTR. Fermentation processes using FBR or PBR are frequently applied with immobilized techniques, where their efficiency and stability are well established. Furthermore,

7

Bioreactor Design Selection for Biohydrogen Production Using. . .

167

Table 7.2 Bioreactor types with various substrates and biohydrogen production performance No.

Substrate

Bioreactor (mode)

Immobilized culture

H2 production performance

References

1

Glucose

PBR (C)

Sludge

26 g H2/kg COD

[9]

2

Glucose

FBR (C)

Sludge

10.4 g H2/kg COD

[9]

3

Food waste

Sludge reactor (S)

Aspergillus awamori and Aspergillus oryzae

353.9 mL H2/h/L

[48]

4

Organic solid waste

Benchtop bioreactor (B)

Anaerobic sludge

215.39 mL H2/g

[30]

5

Xylose

CSTR (C)

Clostridium sp. T2

3.15 mmol H2/L/h

[10]

6

Sucrose

CSTR (C)

Seed sludge

2.7 L H2/L/day

[41]

7

Biomass

Modified AFBR (C)

Dung and sewage sludge

7.57 L H2/L/min

[33]

8

Synthetic food waste

Benchtop bioreactor (B)

Escherichia cloacae and Enterobacter aerogenes

155.2 mL/g

[31]

9

Anaerobic digested sludge

CSTR (B)

Recirculated sludge

N.A. (off-gas contained 30% H2)

[49]

10

POME

UASB (C)

Anaerobic sludge

0.589 L H2/L

[25]

11

Glucose, sewage sludge

UASB (C)

Anaerobic granular sludge

11.6 L H2/L/day

[50]

12

Sucrose

Benchtop bioreactor (B)

Clostridium butyricum and Rhodopseudomonas palustris

N.A. (cumulative H2 is 728 mL)

[12]

13

Various organic acids

CSTR (C)

Purple non-sulfur bacteria

N.A. (45% H2 conversion)

[27]

14

Wheat flour hydrolysate

Benchtop bioreactor (B)

Aspergillus awamori

1.9 mol H2/mol glucose

[48]

15

Glucose

Benchtop bioreactor (C)

Bacillus cereus and Enterobacter cloacae

2.87 mol bioH2/ mol glucose

[51]

16

Wheat straw

CSTR (S)

Thermophilic bacteria

1.3 mol bioH2/mg sugar

[52]

17

Cheese whey

Benchtop bioreactor (B)

Bacillus cereus and Enterobacter cloacae

139 mL/L/h

[53]

18

Agricultural waste

Minireactor (B)

Activated sludge

2.6 mmol H2/g dry substrate

[54]

19

OFMSW

CSTR (S)

Anaerobic digester sludge

360 mL H2/g VS

[28]

20

Pineapple waste

Benchtop STR (B)

Clostridium sporogenes and Enterobacter aerogenes

35.9 mmol H2/h/ Lsubstrate

[55]

B batch, S semicontinuous, C continuous

N. K. A. Jalil et al.

168

these bioreactors are flexible for modification and integration to achieve better production and optimum biohydrogen yield. The entrapment method is the most effective immobilized method for selective bioreactors. It is important to note that the design of bioreactors can impact capital costs, operating costs, and production efficiency. Different bioreactor designs differ in cost because the cost is highly subjective to many factors such as the material used, dimensions and sizes, operating features, maintenance requirements, and the specific needs of the process. A stirred-tank bioreactor is common in industrial settings because of its scalability and familiarity with the process. Due to the need for robust mechanical components and agitation systems, initial capital investment may be higher. A stirred-tank bioreactor requires energy to agitate and regulate temperatures, which may increase costs. Similarly, a fluidized-bed reactor can be more challenging and costly due to the need for larger equipment and more sophisticated controls. Maintenance and replacement costs may arise with mechanical components, especially when immobilized structures are used in the bioreactor.

7.4

Status Quo, Challenges, and Outlook

Numerous researchers are still exploring and improving biohydrogen production owing to its potential as an alternative form of sustainable energy. So far, various carbohydrates and waste biomass have been proven suitable for biohydrogen fermentation. However, the fermentative process for biohydrogen is still challenging to adopt by industries. Current fermentative biohydrogen technology is limited to laboratory and pilot scales [43]. The main challenge is to achieve highly consistent and stable hydrogen yields from biomass or organic substrates. Several strategies have been investigated to improve the efficiency of immobilization of cell culture in fermentative biohydrogen production. Currently, there are various studies to find a more effective immobilization matrix such as catalytic nanoparticles [44], COOHfunctionalized multiwalled carbon nanotubes [45], and polyaniline nanoparticles immobilized on sludge beads [46]. A synergistic cell-enzyme immobilization is another interesting strategy to enhance fermentation yields and biohydrogen production [47]. Implementing a hybrid reactor seems able to boost biohydrogen productivity by optimizing the advantages of two different systems [48]. With ongoing research and technological advancements, bioreactor design upgrading looks promising. Nevertheless, the industrial biohydrogen production requires innovation in biotechnology, fermentation processes, and reactor design. Acknowledgments The author acknowledges the funding from Universiti Teknologi Malaysia (UTM) provided under Grant of Vot numbers Q.J130000.2546.17H90 and Q. J130000.2544.09H24.

7

Bioreactor Design Selection for Biohydrogen Production Using. . .

169

References 1. Kumar, G., & Lin, C.-Y. (2014, February). Biogenic hydrogen conversion of de-oiled jatropha waste via anaerobic sequencing batch reactor operation: Process performance, microbial insights, and CO2 reduction efficiency. Scientific World Journal, 2014, 946503. https://doi. org/10.1155/2014/946503 2. Kim, D.-H., Kim, S.-H., Kim, K.-Y., & Shin, H.-S. (2010). Experience of a pilot-scale hydrogen-producing anaerobic sequencing batch reactor (ASBR) treating food waste. International Journal of Hydrogen Energy, 35(4), 1590–1594. https://doi.org/10.1016/j.ijhydene.2009. 12.041 3. Li, C., & Fang, H. H. P. (2007). Fermentative hydrogen production from wastewater and solid wastes by mixed cultures. Critical Reviews in Environmental Science and Technology, 37(1), 1–39. https://doi.org/10.1080/10643380600729071 4. Wu, K. J., & Chang, J. S. (2007). Batch and continuous fermentative production of hydrogen with anaerobic sludge entrapped in a composite polymeric matrix. Process Biochemistry, 42(2), 279–284. https://doi.org/10.1016/j.procbio.2006.07.021 5. Balachandar, G., Khanna, N., & Das, D. (2013). Chapter 6 – Biohydrogen production from organic wastes by dark fermentation. In Biohydrogen (1st ed.). Elsevier. https://doi.org/10. 1016/B978-0-444-59555-3.00006-4 6. Chang, J. (2002). Biohydrogen production with fixed-bed bioreactors. International Journal of Hydrogen Energy, 27(11–12), 1167–1174. https://doi.org/10.1016/s0360-3199(02)00130-1 7. Wu, K.-J., Chang, C.-F., & Chang, J.-S. (2007). Simultaneous production of biohydrogen and bioethanol with fluidized-bed and packed-bed bioreactors containing immobilized anaerobic sludge. Process Biochemistry, 42(7), 1165–1171. https://doi.org/10.1016/j.procbio.2007. 05.012 8. De Graaff, M. S., Temmink, H., Zeeman, G., & Buisman, C. J. N. (2010). Anaerobic treatment of concentrated black water in a UASB reactor at a short HRT. Water (Basel), 2(1), 101–119. https://doi.org/10.3390/w2010101 9. Arimi, M. M., Knodel, J., Kiprop, A., Namango, S. S., Zhang, Y., & Geißen, S.-U. (2015). Strategies for improvement of biohydrogen production from organic-rich wastewater: A review. Biomass and Bioenergy, 75, 101–118. https://doi.org/10.1016/j.biombioe.2015.02.011 10. Zhao, L., et al. (2012). Enhanced bio-hydrogen production by immobilized clostridium sp. T2 on a new biological carrier. International Journal of Hydrogen Energy, 37(1), 162–166. https:// doi.org/10.1016/j.ijhydene.2011.09.103 11. Singh, L., Wahid, Z. A., Siddiqui, M. F., Ahmad, A., Rahim, M. H. A., & Sakinah, M. (2013). Biohydrogen production from palm oil mill effluent using immobilized Clostridium butyricum EB6 in polyethylene glycol. Process Biochemistry, 48(2), 294–298. https://doi.org/10.1016/j. procbio.2012.12.007 12. Kao, P.-M., Hsu, B.-M., Huang, K.-H., Tao, C.-W., Chang, C.-M., & Ji, W.-T. (2014). Biohydrogen production by immobilized co-culture of Clostridium butyricum and Rhodopseudomonas palustris. Energy Procedia, 61, 834–837. https://doi.org/10.1016/j. egypro.2014.11.976 13. Patel, S. K. S., Purohit, H. J., & Kalia, V. C. (2010). Dark fermentative hydrogen production by defined mixed microbial cultures immobilized on ligno-cellulosic waste materials. International Journal of Hydrogen Energy, 35(19), 10674–10681. https://doi.org/10.1016/j.ijhydene.2010. 03.025 14. Tenca, A., Schievano, A., Perazzolo, F., Adani, F., & Oberti, R. (2011). Biohydrogen from thermophilic co-fermentation of swine manure with fruit and vegetable waste: Maximizing stable production without pH control. Bioresource Technology, 102(18), 8582–8588. https:// doi.org/10.1016/j.biortech.2011.03.102

170

N. K. A. Jalil et al.

15. Chinellato, G., Cavinato, C., Bolzonella, D., Heaven, S., & Banks, C. J. (2013). Biohydrogen production from food waste in batch and semi-continuous conditions: Evaluation of a two-phase approach with digestate recirculation for pH control. International Journal of Hydrogen Energy, 38(11), 4351–4360. https://doi.org/10.1016/j.ijhydene.2013.01.078 16. Mohan, S. V., Mohanakrishna, G., Reddy, S. S., Raju, B. D., Rao, K. S. R., & Sarma, P. N. (2008). Self-immobilization of acidogenic mixed consortia on mesoporous material (SBA-15) and activated carbon to enhance fermentative hydrogen production. International Journal of Hydrogen Energy, 33(21), 6133–6142. https://doi.org/10.1016/j.ijhydene.2008.07.096 17. Engliman, N. S., Jahim, J. M., Abdul, P. M., Ling, T. P., Tan, J. P., & Ong, C. B. (2020, March). Effectiveness of fouling mechanism for bacterial immobilization in polyvinylidene fluoride membranes for biohydrogen fermentation. Food and Bioproducts Processing, 120, 48–57. https://doi.org/10.1016/J.FBP.2019.12.004 18. Show, K.-Y., Lee, D.-J., & Chang, J.-S. (2011). Bioreactor and process design for biohydrogen production. Bioresource Technology, 102(18), 8524–8533. https://doi.org/10.1016/j.biortech. 2011.04.055 19. Sekoai, P. T., et al. (2017). Microbial cell immobilization in biohydrogen production: A short overview. Critical Reviews in Biotechnology, 38(2), 157–171. https://doi.org/10.1080/ 07388551.2017.1312274 20. Liao, Q., Yang, Y. X., Zhu, X., Chen, R., & Fu, Q. (2017, January). Pore-scale lattice Boltzmann simulation of flow and mass transfer in bioreactor with an immobilized granule for biohydrogen production. Science Bulletin (Beijing), 62(1), 22–30. https://doi.org/10.1016/J. SCIB.2016.11.004 21. Hallenbeck, P. C., & Ghosh, D. (2009, May). Advances in fermentative biohydrogen production: The way forward? Trends in Biotechnology, 27(5), 287–297. https://doi.org/10.1016/J. TIBTECH.2009.02.004 22. Show, K.-Y., et al. (2010). Critical assessment of anaerobic processes for continuous biohydrogen production from organic wastewater. International Journal of Hydrogen Energy, 35(24), 13350–13355. https://doi.org/10.1016/j.ijhydene.2009.11.110 23. Barros, A. R., Adorno, M. A. T., Sakamoto, I. K., Maintinguer, S. I., Varesche, M. B. A., & Silva, E. L. (2011). Performance evaluation and phylogenetic characterization of anaerobic fluidized bed reactors using ground tire and pet as support materials for biohydrogen production. Bioresource Technology, 102(4), 3840–3847. https://doi.org/10.1016/j.biortech.2010. 12.014 24. Jones, J., Kindembe, D., Branton, H., Lawal, N., Montero, E.L., Mack, J., Shi, S., Patton, R. &Montague, G. (2023). Improved control strategies for the environment within cell culture bioreactors. Food and Bioproducts Processing, 138, 209–220. https://doi.org/10.1016/j.fbp. 2023.02.004 25. Singh, L., Siddiqui, M. F., Ahmad, A., Rahim, M. H. A., Sakinah, M., & Wahid, Z. A. (2013). Biohydrogen production from palm oil mill effluent using immobilized mixed culture. Journal of Industrial and Engineering Chemistry, 19(2), 659–664. https://doi.org/10.1016/j.jiec.2012. 10.001 26. Argun, H., & Kargi, F. (2011). Bio-hydrogen production by different operational modes of dark and photo-fermentation: An overview. International Journal of Hydrogen Energy, 36(13), 7443–7459. https://doi.org/10.1016/j.ijhydene.2011.03.116 27. Basak, N., Jana, A. K., Das, D., & Saikia, D. (2014, April). Photofermentative molecular biohydrogen production by purple-non-sulfur (PNS) bacteria in various modes: The present progress and future perspective. International Journal of Hydrogen Energy, 39(13), 6853–6871. https://doi.org/10.1016/j.ijhydene.2014.02.093 28. Hu, B. (2013). Biological hydrogen production via self immobilized bacteria. Journal of Chemical Information and Modeling, 53(9), 1689–1699. https://doi.org/10.1017/ CBO9781107415324.004

7

Bioreactor Design Selection for Biohydrogen Production Using. . .

171

29. Ri, P. C., et al. (2019, July). Effect of hydraulic retention time on the hydrogen production in a horizontal and vertical continuous stirred-tank reactor. International Journal of Hydrogen Energy, 44(33), 17742–17749. https://doi.org/10.1016/J.IJHYDENE.2019.05.136 30. Sekoai, P., Yoro, K., & Daramola, M. (2016). Batch fermentative biohydrogen production process using immobilized anaerobic sludge from organic solid waste. Environments, 3(4), 38. https://doi.org/10.3390/environments3040038 31. Hu, B., Liu, Y., Chi, Z., & Chen, S. (2007). Biological hydrogen production via bacteria immobilized in calcium alginate gel beads. Biological Engineering, 1(1), 25–37. https://doi. org/10.13031/2013.24191 32. Rajesh Banu, J., et al. (2021). A critical review on limitations and enhancement strategies associated with biohydrogen production. International Journal of Hydrogen Energy, 46(31), 16565–16590. https://doi.org/10.1016/j.ijhydene.2021.01.075 33. Abu Rahma, A. (2013). Biohydrogen production by modified anaerobic fluidized bed reactor (AFBR) using mixed bacterial cultures in thermophilic condition. Islamic University-Gaza. [Online]. Available: https://library.iugaza.edu.ps/thesis/111749.pdf. Accessed: 1 Oct 2023 34. Han, W., Wang, Z., Chen, H., Yao, X., & Li, Y. (2011). Simultaneous biohydrogen and bioethanol production from anaerobic fermentation with immobilized sludge, 2011, 1–6. https://doi.org/10.1155/2011/343791 35. Keskin, T., Aksöyek, E., & Azbar, N. (2011). Comparative analysis of thermophilic immobilized biohydrogen production using packed materials of ceramic ring and pumice stone. International Journal of Hydrogen Energy, 36(23), 15160–15167. https://doi.org/10. 1016/j.ijhydene.2011.08.078 36. Han, W., Liu, D. N., Shi, Y. W., Tang, J. H., Li, Y. F., & Ren, N. Q. (2015). Biohydrogen production from food waste hydrolysate using continuous mixed immobilized sludge reactors. Bioresource Technology, 180, 54–58. https://doi.org/10.1016/j.biortech.2014.12.067 37. Mao, C., Feng, Y., Wang, X., & Ren, G. (2015). Review on research achievements of biogas from anaerobic digestion. Renewable and Sustainable Energy Reviews, 45, 540–555. https:// doi.org/10.1016/j.rser.2015.02.032 38. K. Stamatelatou, G. Antonopoulou, and P. Michailides, Biomethane and biohydrogen production via anaerobic digestion/fermentation. 2014. doi:https://doi.org/10.1533/9780857097385. 2.476. 39. Gómez, X., Fernández, C., Fierro, J., Sánchez, M. E., Escapa, A., & Morán, A. (2011). Hydrogen production: Two stage processes for waste degradation. Bioresource Technology, 102, 8621–8627. https://doi.org/10.1016/j.biortech.2011.03.055 40. Ferguson, R. M. W., Coulon, F., & Villa, R. (2016). Organic loading rate: A promising microbial management tool in anaerobic digestion. Water Research, 100, 348–356. https:// doi.org/10.1016/j.watres.2016.05.009 41. Keskin, T., Giusti, L., & Azbar, N. (2012). Continuous biohydrogen production in immobilized biofilm system versus suspended cell culture. International Journal of Hydrogen Energy, 37(2), 1418–1424. https://doi.org/10.1016/j.ijhydene.2011.10.013 42. Singh, L., Siddiqui, M. F., Ahmad, A., Rahim, M. H. Ab., Sakinah, M., & Wahid, Z. A. (2013). Application of polyethylene glycol immobilized Clostridium sp. LS2 for continuous hydrogen production from palm oil mill effluent in upflow anaerobic sludge blanket reactor. Biochemical Engineering Journal, 70, 158–165. https://doi.org/10.1016/j.bej.2012.10.010 43. Zhang, Z., et al. (2023, June). Enhanced biohydrogen yield and light conversion efficiency during photo-fermentation using immobilized photo-catalytic nano-particles. Bioresource Technology, 377, 128931. https://doi.org/10.1016/J.BIORTECH.2023.128931 44. Wannapokin, A., Huang, H. T., Chang, P. H., Chien, Y. W., & Hung, C. H. (2022, Decmber). Improving production of biohydrogen from COOH-functionalized multiwalled carbon nanotubes through co-immobilization with clostridium pasteurianum. International Journal of Hydrogen Energy, 47(96), 40704–40713. https://doi.org/10.1016/J.IJHYDENE.2022.09.095

172

N. K. A. Jalil et al.

45. Hellal, M. S., Abou-Taleb, E. M., Rashad, A. M., & Hassan, G. K. (2022, July). Boosting biohydrogen production from dairy wastewater via sludge immobilized beads incorporated with polyaniline nanoparticles. Biomass and Bioenergy, 162, 106499. https://doi.org/10.1016/J. BIOMBIOE.2022.106499 46. Woo, W. X., et al. (2022, Decmber). An overview on cell and enzyme immobilization for enhanced biohydrogen production from lignocellulosic biomass. International Journal of Hydrogen Energy, 47(96), 40714–40730. https://doi.org/10.1016/J.IJHYDENE.2022.08.164 47. Zhao, C., et al. (2021, March). A developed hybrid fixed-bed bioreactor with Fe-modified zeolite to enhance and sustain biohydrogen production. Science of the Total Environment, 758, 143658. https://doi.org/10.1016/J.SCITOTENV.2020.143658 48. Han, W., et al. (2015). Fermentative hydrogen production using wheat flour hydrolysate by mixed culture. International Journal of Hydrogen Energy, 40(13), 4474–4480. https://doi.org/ 10.1016/j.ijhydene.2015.02.016 49. Gottardo, M., Cavinato, C., Bolzonella, D., & Pavan, P. (2013). Dark fermentation optimization by anaerobic digested sludge recirculation: Effects on hydrogen production. Chemical Engineering Transactions, 32, 997–1002. https://doi.org/10.3303/CET1332167 50. Hu, B. (2013). Biological hydrogen production via self immobilized bacteria. Journal of Chemical Information and Modeling, 53(9), 1689–1699. https://doi.org/10.1017/ CBO9781107415324.004 51. Patel, S. K. S., Kumar, P., Mehariya, S., Purohit, H. J., Lee, J. K., & Kalia, V. C. (2014). Enhancement in hydrogen production by co-cultures of Bacillus and Enterobacter. International Journal of Hydrogen Energy, 39(27), 14663–14668. https://doi.org/10.1016/j.ijhydene. 2014.07.084 52. Chandolias, K., Pardaev, S., & Taherzadeh, M. J. (2016). Biohydrogen and carboxylic acids production from wheat straw hydrolysate. Bioresource Technology, 216, 1093–1097. https:// doi.org/10.1016/j.biortech.2016.05.119 53. Patel, A. K., Vaisnav, N., Mathur, A., Gupta, R., & Tuli, D. K. (2016). Whey waste as potential feedstock for biohydrogen production. Renewable Energy, 98, 221–225. https://doi.org/10. 1016/j.renene.2016.02.039 54. Robledo-Narváez, P. N., et al. (2013, October). The influence of total solids content and initial pH on batch biohydrogen production by solid substrate fermentation of agroindustrial wastes. Journal of Environmental Management, 128, 126–137. https://doi.org/10.1016/j.jenvman. 2013.04.042 55. Abd Jalil, N. K., Asli, U. A., Hashim, H., Abd Jalil, A., Ahmad, A., & Khamis, A. K. (2018, August). Biohydrogen production from pineapple biomass residue using immobilized co-culture of clostridium sporogenes and Enterobacter aerogenes. Journal of Energy and Safety Technology (JEST), 1(1). https://doi.org/10.11113/jest.v1n1.8

Chapter 8

Biomass-Based Polygeneration Systems with Hydrogen Production: A Concise Review and Case Study Zahra Hajimohammadi Tabriz, Mousa Mohammadpourfard, Gülden Gökçen Akkurt, and Başar Çağlar

8.1 8.1.1

Introduction Biomass-Driven Polygeneration Systems

Fossil fuel dependency of energy infrastructure creates many challenges. The lack of fossil fuels, their non-uniform distribution in the world, the economic limitations of their access, and the increase in the emission of atmospheric pollutants have reduced the attractiveness of fossil fuels in today’s world. Conventional power plants have 30–35% efficiency [1], the production and distribution of energy through the combustion of fossil fuels cause a large amount of chemical energy to be wasted, and the emission of greenhouse gases, especially carbon dioxide, has dramatically increased due to the use of these fuels. Environmental problems lead to climate change. To deal with this issue, the European Union has pledged to decrease carbon dioxide emissions in 2050 by 80–95% compared to 1990 [2]. Reaching this goal requires a transition toward clean energy and more efficient systems. Renewable energies are clean energy sources that can be replaced or reproduced shortly after consumption. Using these resources solves energy needs at the local level, reduces energy supply costs in remote areas, and helps governments improve life comfort [3]. The popularity of renewable energy sources is increasing. Z. Hajimohammadi Tabriz Faculty of Chemical and Petroleum Engineering, University of Tabriz, Tabriz, Iran M. Mohammadpourfard (✉) Faculty of Chemical and Petroleum Engineering, University of Tabriz, Tabriz, Iran Department of Energy Systems Engineering, Izmir Institute of Technology, Izmir, Türkiye e-mail: [email protected] G. Gökçen Akkurt · B. Çağlar Department of Energy Systems Engineering, Izmir Institute of Technology, Izmir, Türkiye © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 V. Vahidinasab et al. (eds.), Green Hydrogen in Power Systems, Green Energy and Technology, https://doi.org/10.1007/978-3-031-52429-5_8

173

174

Z. Hajimohammadi Tabriz et al.

Renewable energies, mainly biomass, are expected to meet nearly half the world’s energy needs by 2030 [4]. The main challenge of renewable energies is their intermittent nature; however, biomass is the most popular option, exempting this disadvantage [5]. The European Union defines biomass as the decomposable part of materials, leftovers, and refuse from agriculture, forestry, and associated industries, along with the decomposable portion of industrial and municipal waste [6]. One of the main advantages of biomass is that it allows the use of waste as feed for fuel and chemical production, thereby preventing environmental damage caused by improper disposal and providing valorization. Furthermore, biomass’s high availability and carbon neutrality have made it the third most commonly used energy source globally, following oil and coal [7]. As mentioned, the low performance of plants is an essential problem that should be addressed in the transition toward solving energy problems. Polygeneration systems offer a solution by simultaneously producing two or more products, thereby increasing efficiency. The integration of different subsystems in polygeneration systems leads to a reduction in energy and material waste in the system [5]. Furthermore, these systems can significantly reduce CO2 emissions, with estimates suggesting a potential 950 Mton/year reduction by 2030 [8]. Integrating these two solutions leads to biomass-driven polygeneration systems. A schematic of biomass-driven polygeneration systems is presented in Fig. 8.1. These systems combine the advantages of polygeneration with the advantages of using biomass to generate various products such as energy vectors, biofuels, various chemical products, etc. One of the most important products of biomass-driven polygeneration systems is hydrogen, a clean fuel that is more valuable if produced this way. In conclusion, renewable energy sources have become increasingly important in our efforts to combat climate change and reduce our dependence on fossil fuels. Among them, biomass stands out as a versatile and accessible renewable energy source, with the benefit of being carbon neutral. Its potential as an essential player in facilitating the shift toward a more environmentally friendly energy system cannot be overstated, especially when used in a polygeneration system. Furthermore, biomass can also be used to produce hydrogen, which has the potential to revolutionize the way we power our industries, homes, and vehicles. The following section will explore the possibilities of employing biomass as a feedstock for generating hydrogen and how this could help us achieve a cleaner and more efficient energy future.

8.1.2

Hydrogen

Hydrogen has been recognized as a promising energy carrier in transitioning to a sustainable future. As an energy carrier, it can help fill the gap between the intermittent nature of renewable energy sources and the world’s growing energy

8

Biomass-Based Polygeneration Systems with Hydrogen Production

175

Fig. 8.1 A schematic of the biomass-driven polygeneration system. (Adapted from [5], copyright (2023), with permission from Elsevier)

demand [9]. Hydrogen offers several advantages as an eco-friendly and sustainable fuel [10]. One of its significant advantages is its ability to store surplus energy from renewable sources, making it an excellent energy storage option. In addition, hydrogen has a high gravimetric energy density (120 MJ/kg), which is more than two-fold higher than that of other fossil-based hydrocarbon fuels. Its CO2-free ignition process makes it an attractive option for reducing carbon emissions. Hydrogen is also non-toxic and has a highly effective octane number and burning speed, which make it an efficient and safe fuel option [11]. The critical thing about clean fuel is that it is important when its entire life cycle is clean; In other words, it must be produced sustainably and consumed sustainably [9]. An appropriate classification of different feedstocks and methods for generating hydrogen is proposed by Taipabu et al. [12]. As seen in Fig. 8.2, renewable feedstocks include biomass, electricity, and water, and fossil-based feedstocks include natural gas, coal, and oil. Hydrogen production methods are very diverse, but 75% of the total hydrogen generation is made by steam reforming of natural gas in petrochemical refineries for internal consumption (e.g., hydrocracking, hydrodesulfurization, hydrodenitrogenation, etc.). The steam reforming process is energy intensive and leads to excessive CO2 emission (i.e., responsible for the emission of 3% of the total share of carbon dioxide emissions [12]), suggesting that it needs to be replaced by alternative green processes.

176

Z. Hajimohammadi Tabriz et al.

Fig. 8.2 Hydrogen production: feedstocks and methods. (Adapted from [12], copyright (2023), with permission from Elsevier)

Electrolysis of water is a powerful method for this purpose, but it is costly due to its high energy consumption (50 kWh/kg H2) and high water consumption. The related drawbacks create challenges to employing this technology for large-scale hydrogen production [12]. The other alternative is to use biomass feedstock as a hydrogen source. Hydrogen production from biomass feedstock, which is the main focus of this chapter, is a promising and environmentally benign solution that has attracted the attention of researchers. In addition, hydrogen, as one of the polygeneration plant’s products, brings all the advantages that have been counted so far. Biomass-based hydrogen production is attractive due to its economical and eco-friendly nature. It reduces greenhouse gas emissions and offers a sustainable solution for waste management. However, the energy efficiency of biomass-based hydrogen production methods needs to be improved for the economic feasibilities of processes. In this respect, polygeneration methods can be used, allowing simultaneous valorization of system outputs and integrating system components more

8

Biomass-Based Polygeneration Systems with Hydrogen Production

177

efficiently. Thermodynamic evaluation at the system level is necessary to fully understand the potential of biomass-based hydrogen production via polygeneration systems. This will allow us to assess the feasibility and efficiency of these systems.

8.1.3

Thermodynamic Evaluation of Polygeneration Systems

Thermodynamic evaluation of the polygeneration plants is usually done from energy, exergy, economic, and environmental perspectives. To conduct comprehensive evaluations, researchers welcome approaches based on the integration of these analyses under the titles of exergoeconomic and exergoenvironmental assessments. Energy and exergy analyses are based on mass conservation, energy conservation, and exergy balance equations. Assuming the steady-state conditions and disregarding any alterations in potential or kinetic energy, the following equations are presented [13]: m_ i = Q_ j þ

m_ e he

ð8:2Þ

e

i

T Q_ j 1 - 0 Tj

ð8:1Þ

_ þ m_ i hi = W

j

j

m_ e e

i

_ þ m_ i exi = W

þ

_D m_ e exe þ Ex

ð8:3Þ

e

i

Also, the exergy of each state is defined as [13]:

exch = k

ex = exph þ exch

ð8:4Þ

exph = h - h ° - T ° ðs - s ° Þ

ð8:5Þ

1 yk M k

yk ex ° k þ RT 0

yk Lnðyk Þ

ð8:6Þ

k

k

In the above equations, the subscripts i, e, 0, and D refer to input, exit, reference environment conditions, and destruction, respectively. Exergoeconomic analysis is based on the cost balance equation [13]: C_ in,k þ C_ q,k þ Z_ k = in

C_ out,k þ C_ w,k

ð8:7Þ

out

_ C_ = c × Ex

ð8:8Þ

178

Z. Hajimohammadi Tabriz et al.

where C_ and Z_ k are the cost rate and the capital investment cost, respectively. Also, c refers to the unit cost of exergy. Exergoenvironmental analysis has been done in different ways by researchers. The third sub-section will discuss more details about one of the common approaches.

8.1.4

The Objective of This Work

This chapter is presented in four sections to investigate biomass-based polygeneration systems with hydrogen production. Figure 8.3 presents a schematic illustrating the work process. The chapter begins with an introduction about the generalities of this topic, which includes the introduction and importance of biomass-based polygeneration systems, hydrogen production, and the thermodynamic evaluation of these systems. The second section briefly reviews relevant studies on biomass-based hydrogen production in polygeneration systems. Then, the third section presents a case study that evaluates the feasibility of integrating a sewage sludge-driven polygeneration system in the largest wastewater treatment plant in Izmir province, Çiğli WWTP, from 4E (energy, exergy, exergoeconomic, and exergoenvironmental) perspectives. The fourth section discusses the status quo, challenges, and outlook related to these systems. Finally, the last section summarizes the chapter and presents the main results of this study.

Fig. 8.3 Schematic structure of this chapter

8

Biomass-Based Polygeneration Systems with Hydrogen Production

8.2

179

Hydrogen Production in Biomass-Based Polygeneration Systems

In this section, the studies conducted in recent years on biomass-based polygeneration systems and their ability to produce hydrogen as one of their products have been reviewed. This review includes systems whose only energy source is biomass, and no other auxiliary energy sources have been used for them. Biomass is used in two ways for hydrogen production in these systems: as an energy source for starting the hydrogen production unit and as a feedstock for hydrogen production. Systems in which biomass serves as an energy source for initiating the hydrogen production unit usually use the integration of water electrolysis units to produce hydrogen. Water electrolysis is a clean hydrogen production method suitable for small-scale applications. This process splits water into its components. The essential types of electrolyzer are proton exchange membrane electrolyzer (PEME), solid oxide electrolyzer (SOE), and alkaline electrolyzer (AE). Each of these has its advantages and disadvantages. However, today PEME is the most popular and common type because of its higher hydrogen production rate, more compact design, and higher energy efficiency. It reacts quickly to fluctuations, and this advantage makes it easy to integrate it with a renewable source that is intermittent in nature [14]. However, it should be noted that PEMEs have some drawbacks that limit their utilization, despite their benefits. As mentioned in Sect. 8.1.2, these technologies have high water and electricity consumption and lead to high costs. In addition, the presence of water vapor in the hydrogen generated by PEME requires the removal of moisture, which is a notable drawback [15]. As mentioned earlier, biomass can also be used as a renewable feedstock for hydrogen production. This way of hydrogen production is a solution that has attracted more attention recently, but only a few studies have been done to improve it. The abundant and rapid growth of existing biomass ensures its sustainability. In addition, the optimal use of biomass helps to reduce waste, eliminate harm, and reduce atmospheric emissions [12]. In the following sub-sections, the research background of each method of hydrogen production in biomass-based polygeneration systems will be discussed separately. By exploring these different approaches, researchers can develop more efficient and sustainable methods for producing hydrogen from biomass.

8.2.1

Biomass as a Fuel for Driving Hydrogen Production Unit

The two most frequently studied methods for converting biomass are anaerobic digestion and gasification. However, there are still few studies on the biomass anaerobic digestion-based plant with hydrogen production.

180

Z. Hajimohammadi Tabriz et al.

Safari et al. [16] designed a system based on the anaerobic digestion of sewage sludge and evaluated it from the energy and exergy points of view. They indicated that increasing the methane mole fraction in the biogas increases the hydrogen production rate obtained from the PEME unit. Adebayo et al. [17] developed a biomass-driven plant based on a solid oxide fuel cell. A PEME is integrated with this system for hydrogen production, producing 19.94 m3/day of hydrogen. The parametric investigations have shown that enhancing the number of cells in the fuel cell has increased the hydrogen production rate. Biomass gasification-based plants with hydrogen production seem to be the most popular option. Most of the research in this field has investigated gasification as a conversion method and the integration of PEME for hydrogen production. Moharamian et al. [18] proposed a gasification-based plant that uses hydrogen injection into the combustion chamber. The results showed that this injection reduces fuel consumption and, thus, the unit cost of products by 27%. Farajollahi et al. [19] designed a municipal solid waste-driven cogeneration system integrated with a PEME to produce hydrogen and power. This system can produce 608.8 m3/h of hydrogen at a feed rate of 1.155 kg/s biomass. Yuksel et al. [20] presented a gasification-based plant integrated with a PEME for producing hydrogen. The Linde Hampson hydrogen liquefaction cycle converts the hydrogen from this unit into liquid. It has been stated that the hydrogen generation rate is 0.077 kg/s, and the results of parametric studies have shown that increasing the electrolyzer area increases the hydrogen production rate. Yilmaz et al. [21] studied a gasificationbased plant to produce power, cooling, heating, hydrogen, drying, and clean water. This system produced 0.072 kg/s of hydrogen; its energy and exergy efficiencies are 63.84% and 59.26%, respectively. The system developed by Ishaq et al. [22] is new in that its subsystems use low-grade waste heat. Hydrogen production in this system is 0.13 kg/day. Comparative studies hold significant value as they allow decision-makers to comprehensively understand by comparing two or more options. In this field of research, in addition to the cases mentioned above, studies have been conducted to compare the types of biomass feedstock. Xu et al. [23] proposed a biomass-driven tri-generation plant for power, hydrogen, and freshwater production. Hydrogen is produced through solid oxide electrolyzer cells. In this study, the impact of different biomasses used on plant performance has been investigated. For the base case condition, municipal solid waste was considered the feedstock, and then the evaluations were repeated for paper, paddy husk, and wood. The results have shown that municipal solid waste is the best choice because it has the highest exergetic performance, the highest rate of freshwater and hydrogen production, and the lowest cost of products. In addition, studies have also been performed to examine the impact of hydrogen production unit integration on the system’s overall performance. Habibollahzade et al. [24] investigated the system in three different scenarios: model (a) includes a fuel cell and a gasifier; in model (b), a Stirling engine is added to the system, and model (c) is completed by integrating a PEME. Their findings showed that model (b) is the best model, and model (c) is preferable to model (a) because it has higher

8

Biomass-Based Polygeneration Systems with Hydrogen Production

181

Fig. 8.4 Schematic of the plant proposed by Gholamian et al. (Adapted from [25], copyright (2023), with permission from Elsevier)

H2 production and exergy efficiency. According to the results, this system’s hydrogen production rate is 56.5 kg/s. Also, the parametric studies have shown that the increase in the current density of fuel cells increases the power generation and, thus, the hydrogen production rate, while increasing the utilization factor does the opposite. Gholamian et al. [25] assessed a system (see Fig. 8.4) in three similar models. In this study, model (b) is the best, and model (c) is superior to model (a). The hydrogen production rate in this system is 8.393 kg/s, and the exergy efficiency at the best solution point is reported as 32.3%. The integration of thermochemical cycles such as vanadium-chlorine (V-Cl), which is the most popular, and copper–chlorine (Cu-Cl) cycles has also been considered for hydrogen production in biomass-driven polygeneration systems [26, 27]. In these studies, it has been shown that these units integration with biomass provides a lower cost per unit of exergy compared to solar sources. In addition, researchers have also conducted comparative studies to compare the use of the V-Cl cycle and PEME. Cao et al. [28] proposed a system based on molten carbonate fuel cells and investigated hydrogen production in two scenarios of integration of PEME and V-Cl cycle. The results indicated that the V-Cl cycle system performs better than the system with PEME. The hydrogen production rate when the V-Cl cycle is integrated with the system is higher than PEME integration (see Fig. 8.5). Also, costs are lower when using the V-Cl cycle because it utilizes waste heat instead of electricity.

182

Z. Hajimohammadi Tabriz et al.

Fig. 8.5 The comparison of hydrogen production with two methods. (Adapted from [28], copyright (2023), with permission from Elsevier)

Biomass gasification-based and biomass anaerobic digestion-based systems were reviewed above. Each biomass conversion method has its advantages and disadvantages. In a comparative study Cao et al. [15] compared anaerobic digestion and gasification methods for biomass conversion in a polygeneration system. The hydrogen production unit considered for this system is an alkaline electrolyzer, which has lower costs than other electrolyzers. The results of this study show that using anaerobic digestion compared to gasification has the advantages of a higher production rate, higher exergy efficiency (by 6.4%), and less carbon dioxide emission (by 30.0%). On the other hand, gasification integration is economically better than anaerobic digestion (because of the lower production cost of 21.6%). As a result, choosing one of these two methods will depend on the decision-maker’s criteria. Table 8.1 gives a summary of the discussed systems.

8.2.2

Biomass as a Feedstock for Hydrogen Production

The demand for hydrogen is increasing daily. Hydrogen production using biomass, as previously discussed, seems to be a sustainable solution. In the previous subsection, hydrogen production in biomass-based polygeneration systems was presented, and the systems in which biomass is used as a fuel for driving hydrogen production

8

Biomass-Based Polygeneration Systems with Hydrogen Production

183

Table 8.1 Summary of research on biomass-driven systems with H2 production Biomass conversion Digestion

H2 production PEME

Maize silage

Digestion

PEME

Wood Municipal solid waste Various waste materials

Gasification Gasification

PEME PEME

Gasification

PEME

Demolition wood

Gasification

PEME

Rice husk

Gasification

PEME

Municipal solid waste, wood, paddy husk, paper Municipal solid waste

Gasification

SOE

Gasification

PEME

Power and H2

Municipal solid waste Algal biomass Sawdust wood

Gasification

PEME

Power and H2

Gasification

V-Cl

Power and H2

Gasification

Cu-Cl

Municipal solid waste

Gasification

PEME/VCl

Electricity, heating, cooling, drying, H2, and hot water Power and H2

Municipal solid waste

Digestion/ Gasification

AE

Biomass Sewage sludge

Outputs Power, H2, freshwater, and hot water Electricity, H2, cooling, and domestic hot water Power and H2 Power and H2 Power, heatingcooling, H2, and hot water Power, cooling, drying, H2, heating, and hot water Power, heating, and H2 Power, freshwater, and H2

Power and H2

Conclusion η = 63.6%, ψ = 40%

References [16]

η = 69.86%, ψ = 47.4%

[17]

– ψ = 29.44%

[18] [19]

η = 61.57%, ψ = 58.15%

[20]

η = 63.84%, ψ = 59.26%

[21]

η = 58.03%, ψ = 32.78%

[22]

For MSW (best performance): ψ = 16.72%

[23]

ηopt = 31.13%, 67.38%, 66.41%, ψopt = 28.51%, 39.41%, 38.03%, in 3 scenarios ψopt,b = 33.22%, ψopt, c = 32.3% η = 53.2%, ψ = 52.6% η = 56.71%, ψ = 53.59%

[24]

Hydrogen production rate: PEME = 293 kg/day, and V-Cl cycle = 585 kg/day ψopt,a = 39.20%, ψopt, b = 41.70%

[28]

[25] [26] [27]

[15]

184

Z. Hajimohammadi Tabriz et al.

Fig. 8.6 Effect of gasifier temperature and type of agent on the hydrogen concentration of the syngas for wood biomass. (Adapted from [30], copyright (2023), with permission from Elsevier)

units were reviewed. In this subsection, hydrogen production from syngas will be discussed; in other words, a review of the systems in which biomass is used as a feedstock for hydrogen production. Syngas is a rich source of hydrogen. This gas, obtained from the gasification reaction of organic materials, is mainly composed of carbon monoxide, hydrogen, and a small amount of hydrocarbon compounds. This technology has expanded and can contribute to socioeconomic development and energy security [29]. The quality of the syngas from the gasification process depends on various factors: the type of biomass, the type of agent, agent-to-biomass ratio, and gasification temperature. Shayan et al. [30] investigated the effect of important factors on the quality of syngas and its hydrogen content. In that study, two types of biomasses, wood and paper, and four types of agents, air, oxygen, air-enriched with oxygen, and steam, have been considered. The findings of this study demonstrate that using steam as an agent exhibits the highest efficiency for hydrogen production (see Fig. 8.6). They reported that oxygen utilized as an agent results in greater hydrogen production from paper than from wood. Furthermore, parametric studies reveal that elevating the gasification temperature initially amplifies hydrogen production but subsequently diminishes it. This phenomenon is primarily attributed to the influence of shift reactions, which are endothermic and, therefore, favor higher temperatures. However, as the temperature increases, other reactions are impacted, ultimately decreasing hydrogen production. Safarian et al. [7] performed a similar comparative investigation for another type of biomass, Timber and wood waste, and considered two agents, air and a mixture of air and steam. In their proposed system, the syngas from the gasifier enters a water-gas shift reactor, and the hydrogen in the gas is separated during the PSA process. Their results agree with the previous study’s results, and using steam as an agent is recommended.

8

Biomass-Based Polygeneration Systems with Hydrogen Production

185

Despite the potential of this approach, biomass utilization for hydrogen generation has received less attention than water electrolysis in research studies on polygeneration systems. However, as discussed earlier, utilizing biomass as a feedstock can offer environmental benefits by reducing waste and promoting sustainability. The following paragraph highlights recent studies conducted in this area. Sotoodeh et al. [31] proposed a woody biomass-based multigeneration system for power, cooling, heating, and hydrogen generation. Biomass produces syngas under the gasification reaction, and the hydrogen in this syngas is separated in a PSA process. This process recovers 85% of the hydrogen content of syngas. This system’s overall energy and exergy efficiencies are 52.3% and 41.3%, respectively. Also, the amount of hydrogen produced is reported to be 0.0718 kg/s. Ishaq et al. [32] developed a corn straw-based polygeneration system. Under the gasification process, biomass generates syngas, which subsequently undergoes a water-gas shift reaction (WGSR) to increase its hydrogen content. The calculation revealed that the hydrogen production rate before and after the WGSR is 129.5 and 171 mol/s, respectively. Ishaq et al. [33] presented a dry olive pits-driven trigeneration system that produces hydrogen through a PEME and a WGSR of syngas with energy and exergy performances of 53.7% and 45.5%, respectively. Also, the rate of hydrogen production by this system is 10.74 mol/s for the biomass flow rate of 0.4 kg/s. The schematic of this system is presented in Fig. 8.7.

Fig. 8.7 Schematic of the plant proposed by Ishaq et al. (Adapted from [33], copyright (2023), with permission from Elsevier)

186

Z. Hajimohammadi Tabriz et al.

In conclusion, this section reviewed the studies conducted on biomass-based polygeneration systems and their ability to produce hydrogen as a product. The process of biomass conversion and hydrogen production were discussed in detail. Biomass can be used as an energy source for starting the hydrogen production unit or as a feedstock for hydrogen generation. Water electrolysis is a clean method for hydrogen production, with PEME being the most popular type due to its higher hydrogen production rate, more compact design, and higher energy efficiency. Biomass is recognized as a renewable feedstock for generating hydrogen, which ensures sustainability and helps to reduce waste and atmospheric emissions. Overall, these studies provide valuable insights into developing more efficient and sustainable methods for producing hydrogen from biomass.

8.3

Case Study

Considering the importance of producing hydrogen from biomass-driven polygeneration systems, and also the lack of studies in the field of thermodynamic evaluations of hydrogen production from syngas through the water-gas shift reaction process, a new bio-waste driven multigeneration system for power, hydrogen, freshwater, and heating production is designed and analyzed from 4E perspective. In this system hydrogen is produced through a water-gas shift reaction unit. Furthermore, the proposed system has been studied for Çiğli, an urban area in the Izmir province of Turkey, with an average temperature and relative humidity of 18.11 °C and 60.05% [34]. The feedstock of this system is the sewage sludge from the Çiğli WWTP, which has an average capacity of 600,000 m3 of wastewater per day. The authors’ recent study [35] introduced this system, which is being examined in this chapter due to its significant subjectivity and ability to bridge the research gap mentioned earlier.

8.3.1

System Description

The schematic of the proposed system is presented in Fig. 8.8. The feed of this system is sewage sludge, and different subsystems are integrated for power, heating, hydrogen, and freshwater production. Biomass feedstock is converted in the anaerobic digestion unit and gasifier. The biogas that come by the digester is used as the Brayton cycle fuel, and the syngas obtained from the gasifier enters the water-gas shift reaction unit to produce hydrogen. The energy potential of hot gases from different subsystems is used in atmospheric water harvesting units, organic Rankine cycles, and heating production.

Fig. 8.8 Schematic of the proposed system. (Reprinted from Tabriz et al. [35], copyright (2023), with permission from Elsevier)

8 Biomass-Based Polygeneration Systems with Hydrogen Production 187

188

Z. Hajimohammadi Tabriz et al.

8.3.2

Materials and Methods

The proposed system has been molded using engineering equation solver (EES) software and is evaluated from a thermodynamic point of view. 4E analyses is implemented on the system considering the following main assumptions to simplify the equations [13, 36]: . . . .

The system works in a steady-state condition. Gases have the ideal gas behavior. Changes in kinetic and potential energy and exergy are insignificant. The reference condition is air with a temperature of 25 °C, pressure of 101.3 kPa, and relative humidity of 40%. Also, the chemical composition of air is assumed to be 77.48% N2, 20.59% O2, 1.90% H2O, and 0.03% CO2.

The necessary equations for system modeling were presented in Sect. 8.1.3. In the following, some other relations will be discussed. Because a 4E analysis has been implemented on this system, the equations from the four perspectives of energy, exergy, exergoeconomic, and exergoenvironmental will be presented separately.

8.3.2.1

Energy Analysis

Conservation of mass and conservation of energy laws are applied to the energy analysis of various subsystems. In the following, the most important equations related to each one, which is the basis of the energy analysis of that subsystem, are presented. The general reaction of the anaerobic digestion, complete combustion of biogas, steam gasification, and water-gas shift reaction are expressed as [33, 37–39]: C n H a Ob þ n -

a b n a b n a b H2O → CH 4 þ þ - þ CO2 ð8:9Þ 4 2 2 8 4 2 8 4

λ½xCH 4 CH 4 þ xCO2 CO2 ] þ ½0:7748N 2 þ 0:2059O2 þ 0:0003CO2 þ 0:019H 2 O] → 1 þ λ * ½Y N 2 N 2 þ Y O2 O2 þ Y CO2 CO2 þ Y H 2 O H 2 O]

ð8:10Þ

CH a Ob þ wH 2 O þ mH 2 O → a1 H 2 þ a2 CO þ a3 CO2 þ a4 H 2 O þ a5 CH 4 ð8:11Þ CO þ H 2 O → CO2 þ H 2

ð8:12Þ

To investigate the overall system performance, energy efficiency, exergy efficiency, total power generation rate, and heating production rate are defined as: ηen =

_ net þ Q_ Heating þ m_ H 2 LHV H 2 þ m_ water hwater W m_ Biomass LHV Biomass þ Q_ Gas þ m_ 15 h15 þ m_ 17 h17

ð8:13Þ

8

Biomass-Based Polygeneration Systems with Hydrogen Production

189

W_ net = W_ GT þ W_ ST , SRC þ W_ ST ,ORC1 þ W_ ST ,ORC2 - W_ AC þ W_ FC þ W_ pmp, SRC þ W_ pmp,ORC1 þ W_ pmp,ORC2 þ W_ pmp, ARC þ W_ Fan

ð8:14Þ Q_ Heating = Q_ Heater þ Q_ Cond,SRC = m_ 61 ðh61 - h60 Þ þ m_ 67 ðh67 - h66 Þ

8.3.2.2

ð8:15Þ

Exergy Analysis

Unlike energy, conservation law does not hold for exergy, and irreversibility causes exergy loss. The main equations of exergy balance are given in Sect. 8.1.3, here the overall equations, including total exergy destruction rate, the ratio of components’ exergy destruction, and overall exergy efficiency, are given: _ D,Tot = Ex

_ D,k Ex

ð8:16Þ

k

Y D,k =

ψ ex =

_ D,k Ex _ D,Tot Ex

ð8:17Þ

_ net þ ðm_ 61 ðex61 - ex60 Þ þ m_ 67 ðex67 - ex66 ÞÞ þ m_ H 2 exH 2 þ m_ water exwater W m_ Biomass exBiomass þ Q_ Gas 1 - T 0 =T Gasf þ m_ 15 ex15 þ m_ 17 ex17 ð8:18Þ

8.3.2.3

Exergoeconomic Analysis

Exergoeconomic Analysis provides a suitable approach for system evaluation by combining economic and exergy concepts. The total cost rate for this system and the unit cost of the system’s products are defined below [40–45]: C_ TOTAL = C_ f þ C_ env þ C_ Q þ C_ D,TOTAL þ

Z_ k

ð8:19Þ

k

cproduct =

C_ product _ product Ex

ð8:20Þ

where _ f ; cf = 2 =GJ C_ f = cf ∙ Ex

ð8:21Þ

190

Z. Hajimohammadi Tabriz et al.

C_ env = cCO2 ∙ m_ CO2 ,40 ; cCO2 = 0:024 =kg

ð8:22Þ

C_ Q = cq ∙ Q_ Gas ; cq = 0:04 =kWh

ð8:23Þ

_ D,k C_ D,k ; C_ D,k = cF,k ∙ Ex

C_ D,TOTAL =

ð8:24Þ

k

8.3.2.4

Exergoenvironmental Analysis

To implement this analysis, the parameters of the exergoenvironment factor ( fei), the environmental damage effectiveness factor (θei), the exergoenvironmental impact coefficient (Cei), exergoenvironmental impact improvement (θeii), exergy stability factor ( fes), and unit emission of carbon dioxide (EMI CO2 ) are defined as follows [46–49]: f ei =

_ D,Tot Ex _ in Ex

θei = f ei ∙ C ei ; Cei =

ð8:26Þ ð8:27Þ

_ D,Tot Ex _ _ out þ 1 ExD,Tot þ Ex

ð8:28Þ

m_ CO2 ,emitted × 3600 _ _ W net þ QHeating þ m_ H 2 LHV H 2 þ m_ water hwater

ð8:29Þ

f es =

8.3.3

1 ψ ex

1 θei

θeii =

EMICO2 =

ð8:25Þ

Results and Discussion

In this section, the results of the evaluations performed on the system are presented and discussed in detail. The primary input data for modeling is listed in Table 8.2, and the composition of the sewage sludge is taken from ref. [50]. The main results obtained from evaluating the proposed system in four groups of results related to energy analysis, exergy analysis, exergoeconomic analysis, and exergoenvironmental analysis are given in Table 8.3. The system’s overall energy and exergy efficiencies are 35.48% and 40.18%, respectively. These numbers are reasonable compared to the efficiencies reported for other biomass-driven plants since they typically have high energy and exergy losses due to chemical reactions involved in the processes [5]. The total power generated by this system is about three

8

Biomass-Based Polygeneration Systems with Hydrogen Production

191

Table 8.2 Input data for system modeling AD [13, 16] m_ Sewage sludge LHVSludge LHVDigestate TDigester PDigester Dest.Digester Sludge formula BC [13] rP, Comp ηs, Comp T9 T10 rP, GT ηs, GT Gasification [31] T2 T15 T16 PGas STBM Digestate formula WGSRU T17 T18

7.52 kg/s 18,000 kJ/kg 14,500 kJ/kg 35 °C 101.3 kPa 70% C13.08H18.95O5.17 10 85% 576.85 °C 1246.85 °C 8.3 85% 35 °C 400 °C 800 °C 101.3 kPa 1 CH0.89O0.28 200 °C 25 °C

ORC1 [51] Working fluid m_ T46 P46 P47 ηs, Pump ηs, ST ORC2 [51] Working fluid m_ P51 P52 ηs, Pump ηs, ST ARC [52] ηs, Pump εSHX εCEHX T32 T29 T20 AWH [36] RH34 RH35

Cyclohexane 13 kg/s 300 °C 3000 kPa 100 kPa 90% 85% Isobutane 5 kg/s 3000 kPa 100 kPa 90% 85% 90% 100% 95% -10 °C 40 °C 40 °C 40% 100%

times the required power of Çiğli WWTP, which means that integrating this plant into the treatment plant not only enables the removal of harmful waste but also provides surplus electricity in addition to the need of the plant. Moreover, the results of the exergoeconomic analysis show that the highest contribution to the overall system cost belongs to heat supply to the gasifier, suggesting that the utilization waste heat for the gasifier is important to improve the system economy. The share of exergy destruction rate of different subsystems is shown in Fig. 8.9. As can be seen, the highest exergy destruction is caused by the biomass conversion subsystem with 70%, followed by the water-gas shift reaction unit with 10%. This is mainly due to the high irreversibility of chemical reactions, which increases the exergy destruction rate. Also, the gasifier has the highest share (57.75%) of exergy destruction among all the equipment. Parametric studies are performed to examine the impact of biomass flow rate and gasification temperature, as the main variables, on the most important overall performance criteria of the system, including overall energy efficiency, overall exergy efficiency, unit emission of carbon dioxide, total cost rate of the system, and unit cost of total products. As shown in the diagrams of Fig. 8.10, an increase in

_ H2 = 3180 kg=h m

Q_ Heating = 47440 kW WGR = 18.42 l/h

4E analysis results ηOverall = 35.48% _ net = 17750 kW W

Table 8.3 Results of 4E analysis [35]

YD, max = Gas

= 0.5775

ψ Overall = 40.18% _ D,Tot = 146170 kW Ex

cp, TOTAL = 13.05 $/GJ

C_ D,TOTAL = 30:31 M=Year C_ TOTAL = 102:2 M=Year

C_ env = 7:647 M=Year C_ Q = 55:36 M=Year

Z_ total = 1:949 M=Year C_ f = 6:978 M=Year

fei = 0.594

EMICO2 = 0:2327 t=MWh

fes = 0.5383

θeii = 0.6764

Cei = 2.489

θei = 1.478

192 Z. Hajimohammadi Tabriz et al.

8

Biomass-Based Polygeneration Systems with Hydrogen Production

BCS BC

0% 2%

0%

10% 5%

17%

AWH

Digester

7%

SRC ORC

0%

Gasifier

6%

Heat Ex.

MIX HEAT

193

70%

83%

WGSRU

_ D,sub:sys ). (Reprinted Fig. 8.9 The percentage share of exergy destruction rate of subsystems (Ex from Tabriz et al. [35], copyright (2023), with permission from Elsevier)

Fig. 8.10 The impact of biomass flow rate (a) and gasification temperature (b) on the overall energy and exergy efficiencies and the unit emission of CO2. (Reprinted from Tabriz et al. [35], copyright (2023), with permission from Elsevier)

biomass feed rate improves the energy performance and the unit emission of CO2 while decreasing exergy efficiency due to the increase in the system’s total exergy destruction rate. Unlike the biomass feed rate, gasifier temperature has a positive effect on all parameters, i.e., the energy and exergy efficiencies increase by 16.58 and 17.3%, respectively, and CO2 emission decreases by about 135 kg/MWh. This is mainly due to the increase in hydrogen production rate with gasification temperature. The impact of the gasifier’s temperature on the system’s total cost rate and unit cost of total products are shown in Fig. 8.11. As seen from the Figure, the total cost rate and unit cost of total products increases with gasifier temperature. This is related to the increase in the exergy destruction rate, which also leads to higher unit costs of electricity and hydrogen. These results indicate that although increasing the temperature of the gasifier increases the amount of hydrogen production, it also increases the unit cost of hydrogen production due to the higher heating cost of the gasifier. The effect of gasifier’s temperature and biomass feedstock rate on the hydrogen production rate was also studied. The results show that hydrogen production increases with both the biomass supply rate and the gasifier temperature (Fig. 8.12). The latter is explained by the fact that hydrogen-producing reactions are favored at high temperatures due to their endothermic nature.

194

Z. Hajimohammadi Tabriz et al.

Fig. 8.11 The impact of gasification temperature on the system’s total cost rate and unit cost of total products (a), and unit cost of electricity and hydrogen (b). (Reprinted from Tabriz et al. [35], copyright (2023), with permission from Elsevier)

Fig. 8.12 The impact of feed rate (a) and gasifier temperature (b) on the hydrogen production rate

Hydrogen production is significantly affected by the characteristics of syngas, which mainly depends on gasification temperature, biomass moisture content, and steam-to-biomass ratio. Figures 8.13, 8.14, and 8.15 show the impact of these parameters on the composition and heating value of syngas. As explained above, the gasifier temperature has a positive effect on the hydrogen production rate. On the other hand, a higher humidity inside the gasifier and a higher steam-to-biomass ratio cause lower hydrogen production. This is mainly due to the resulting temperature drop in the gasifier.

8.4

Status Quo, Challenges, and Outlook

The design, construction, and advancement of hydrogen energy systems necessitate significant attention and innovative approaches. It is particularly crucial to identify methods for producing green hydrogen and evaluate their technical, economic and environmental performances. Recent studies have explored various material and energy resources for green hydrogen production systems. Each approach presents

8

Biomass-Based Polygeneration Systems with Hydrogen Production

195

Fig. 8.13 The effect of gasification temperature on the components’ fractions and LHV of the syngas. (Reprinted from Tabriz et al. [35], copyright (2023), with permission from Elsevier)

Fig. 8.14 The effect of biomass moisture content on the components’ fractions and LHV of the syngas. (Reprinted from Tabriz et al. [35], copyright (2023), with permission from Elsevier)

distinct advantages and disadvantages, and their selection hinges on diverse criteria such as efficiency, accessibility, affordability, and environmental impact. Hydrogen production heavily relies on fossil fuels, but their limited availability and contribution to atmospheric pollution necessitate a transition to renewable energy sources. However, hydrogen production systems based on renewable energy often exhibit low efficiency and high costs. Thus, research in this field is still ongoing.

196

Z. Hajimohammadi Tabriz et al.

Fig. 8.15 The effect of steam-to-biomass ratio on the components’ fractions and LHV of the syngas. (Reprinted from Tabriz et al. [35], copyright (2023), with permission from Elsevier)

Hydrogen production within biomass-based polygeneration systems presents a promising approach for achieving green hydrogen production. This process not only addresses energy demands with enhanced efficiency but also encompasses crucial economic and environmental considerations. By harnessing biomass as a carbonneutral energy source, research endeavors focusing on hydrogen production in biomass-driven polygeneration systems can contribute significantly to the transition toward clean energy systems, thereby reducing fossil fuel-sources gas emissions. The conversion of biomass into hydrogen and the utilization of biomass as an energy source present several ongoing challenges. One prominent method for biomass conversion, gasification, is widely adopted but comes with certain drawbacks. High exergy destruction in gasifiers, caused by the irreversible nature of chemical reactions, and the production of tar pose significant issues. These factors reduce efficiency and increase costs, thus limiting its application. Consequently, conducting comprehensive research is imperative to address these shortcomings and improve overall efficiency in this field. Given the high availability of biomass, it holds great potential for energy production and the generation of clean fuels, such as hydrogen. However, unlocking this potential requires financial investment and governmental support for industrial experts and academic researchers. By fostering advancements in biomass, especially bio-waste, utilization, not only the energy production can be enhanced, but also waste management approaches can be improved, promoting the principles of a circular economy. This comprehensive approach holds significant promise for addressing energy needs and bolstering sustainability efforts.

8

Biomass-Based Polygeneration Systems with Hydrogen Production

197

Furthermore, the future of studies on hydrogen production from biomass is anticipated to witness a greater involvement of artificial intelligence methods, including machine learning techniques and multi-objective optimizations. These advanced methods leverage data collection from diverse reports to efficiently forecast the impact of various operational parameters on systems’ performance criteria with the least time and cost consumption.

8.5

Conclusions

The increasing demand for hydrogen as a clean energy carrier necessitates its production through environmentally safe processes. Presently, hydrogen production from fossil fuels has adverse environmental impacts. Polygeneration systems offer a promising solution to mitigate these drawbacks due to their high efficiency and low pollution. Furthermore, coupling polygeneration systems with renewable energy sources amplifies their benefits. In addition, biomass, especially wasted biomass (biowaste) is a smart solution that uses a free resource to produce a valuable product while preventing the damage of releasing these wastes into the environment. Two approaches to hydrogen production in biomass-based polygeneration systems are possible: using biomass as a fuel to power a hydrogen production unit and using biomass as a feedstock for hydrogen generation. The first approach mainly involves integrating PEMEs into the system and supplying its necessary power from biomass, while the second approach focuses on separating hydrogen from the syngas. However, PEMEs can be used on small scales and have some limitations and disadvantages, such as high water and electricity consumption, which result in high costs. Therefore, research and development efforts are needed to improve the performance and sustainability of this method. Despite these challenges, separating hydrogen from the syngas has significant potential for meeting the growing demand for clean energy. The progress of research and commercialization of technologies shows that in the long term, hydrogen production from syngas is the technology that will have the brightest future in response to society’s increasing demands. Syngas is produced during the gasification process. Gasification is an important process that is more environmentally friendly than combustion. Steam gasification is the most feasible option as it produces hydrogen-rich syngas without costly oxidizing agents like oxygen. However, predicting the ideal working conditions for gasification can be difficult due to the complex interactions between various parameters. Fortunately, advancements in artificial intelligence and machine learning algorithms may offer promising solutions to these challenges.

198

Z. Hajimohammadi Tabriz et al.

References 1. Khani, L., & Mohammadpourfard, M. (2023). A new stable solar system for electricity, cooling, heating, and potable water production in sunny coastal areas. In V. Vahidinasab & B. Mohammadi-Ivatloo (Eds.), Energy systems transition: Digitalization, decarbonization, decentralization and democratization (pp. 53–86). Springer International Publishing. 2. Segurado, R., Pereira, S., Correia, D., & Costa, M. (2019). Techno-economic analysis of a trigeneration system based on biomass gasification. Renewable and Sustainable Energy Reviews, 103, 501–514. https://doi.org/10.1016/j.rser.2019.01.008 3. Mousavi, S. A., Mehrpooya, M., & Delpisheh, M. (2022). Development and life cycle assessment of a novel solar-based cogeneration configuration comprised of diffusion-absorption refrigeration and organic Rankine cycle in remote areas. Process Safety and Environmental Protection, 159, 1019–1038. https://doi.org/10.1016/j.psep.2022.01.067 4. Tezer, Ö., Karabağ, N., Öngen, A., Çolpan, C. Ö., & Ayol, A. (2022). Biomass gasification for sustainable energy production: A review. International Journal of Hydrogen Energy, 47(34), 15419–15433. https://doi.org/10.1016/j.ijhydene.2022.02.158 5. Tabriz, Z. H., Khani, L., Mohammadpourfard, M., & Akkurt, G. G. (2022). Biomass driven polygeneration systems: A review of recent progress and future prospects. Process Safety and Environmental Protection, 169, 363. https://doi.org/10.1016/j.psep.2022.11.029 6. https://www.eea.europa.eu/help/glossary/eea-glossary/biomass. Accessed. 7. Safarian, S., Unnthorsson, R., & Richter, C. (2022). Hydrogen production via biomass gasification: Simulation and performance analysis under different gasifying agents. Biofuels, 13(6), 717–726. https://doi.org/10.1080/17597269.2021.1894781 8. Cho, H., Smith, A. D., & Mago, P. (2014). Combined cooling, heating and power: A review of performance improvement and optimization. Applied Energy, 136, 168–185. 9. Dincer, I., & Acar, C. (2018). Smart energy solutions with hydrogen options. International Journal of Hydrogen Energy, 43(18), 8579–8599. https://doi.org/10.1016/j.ijhydene.2018. 03.120 10. Lykas, P., Georgousis, N., Bellos, E., & Tzivanidis, C. (2023). A comprehensive review of solar-driven multigeneration systems with hydrogen production. International Journal of Hydrogen Energy, 48(2), 437–477. https://doi.org/10.1016/j.ijhydene.2022.09.226 11. Moravvej, Z., Soroush, E., Makarem, M. A., & Rahimpour, M. R. (2021). Chapter 7: Thermochemical routes for hydrogen production from biomass. In M. R. Rahimpour, R. Kamali, M. Amin Makarem, & M. K. D. Manshadi (Eds.), Advances in bioenergy and microfluidic applications (pp. 193–208). Elsevier. 12. Taipabu, M. I., Viswanathan, K., Wu, W., Hattu, N., & Atabani, A. E. (2022). A critical review of the hydrogen production from biomass-based feedstocks: Challenge, solution, and future prospect. Process Safety and Environmental Protection, 164, 384–407. https://doi.org/10.1016/ j.psep.2022.06.006 13. Auracher, H. (1996). Thermal design and optimization: Adrian Bejan, George Tsatsaronis and Michael Moran (Vol. 19, p. 482). Wiley. 14. Ishaq, H., Dincer, I., & Crawford, C. (2022). A review on hydrogen production and utilization: Challenges and opportunities. International Journal of Hydrogen Energy, 47(62), 26238–26264. https://doi.org/10.1016/j.ijhydene.2021.11.149 15. Cao, Y., Dhahad, H. A., Hussen, H. M., Anqi, A. E., Farouk, N., & Issakhov, A. (2022). Development and tri-objective optimization of a novel biomass to power and hydrogen plant: A comparison of fueling with biomass gasification or biomass digestion. Energy, 238, 122010. https://doi.org/10.1016/j.energy.2021.122010 16. Safari, F., & Dincer, I. (2019). Development and analysis of a novel biomass-based integrated system for multigeneration with hydrogen production. International Journal of Hydrogen Energy, 44(7), 3511–3526. https://doi.org/10.1016/j.ijhydene.2018.12.101

8

Biomass-Based Polygeneration Systems with Hydrogen Production

199

17. Adebayo, V., Abid, M., Adedeji, M., & Hussain Ratlamwala, T. A. (2022). Energy, exergy and exergo-environmental impact assessment of a solid oxide fuel cell coupled with absorption chiller & cascaded closed loop ORC for multigeneration. International Journal of Hydrogen Energy, 47(5), 3248–3265. https://doi.org/10.1016/j.ijhydene.2021.02.222 18. Moharamian, A., Soltani, S., Rosen, M. A., & Mahmoudi, S. M. S. (2018). Exergoeconomic and thermodynamic analyses of an externally fired combined cycle with hydrogen production and injection to the combustion chamber. International Journal of Hydrogen Energy, 43(2), 781–792. https://doi.org/10.1016/j.ijhydene.2017.11.136 19. Farajollahi, A., Hejazirad, S. A., & Rostami, M. (2022). Thermodynamic modeling of a power and hydrogen generation system driven by municipal solid waste gasification. Environment, Development and Sustainability, 24(4), 5887–5916. https://doi.org/10.1007/s10668-02101690-9 20. Yuksel, Y. E., Ozturk, M., & Dincer, I. (2019). Energy and exergy analyses of an integrated system using waste material gasification for hydrogen production and liquefaction. Energy Conversion and Management, 185, 718–729. https://doi.org/10.1016/j.enconman.2019.02.033 21. Yilmaz, F., Ozturk, M., & Selbas, R. (2019). Development and techno-economic assessment of a new biomass-assisted integrated plant for multigeneration. Energy Conversion and Management, 202, 112154. https://doi.org/10.1016/j.enconman.2019.112154 22. Ishaq, H., Islam, S., Dincer, I., & Yilbas, B. S. (2020). Development and performance investigation of a biomass gasification based integrated system with thermoelectric generators. Journal of Cleaner Production, 256, 120625. https://doi.org/10.1016/j.jclepro.2020.120625 23. Xu, Y.-P., et al. (2022). Optimization of a biomass-driven Rankine cycle integrated with multieffect desalination, and solid oxide electrolyzer for power, hydrogen, and freshwater production. Desalination, 525, 115486. https://doi.org/10.1016/j.desal.2021.115486 24. Habibollahzade, A., Gholamian, E., Houshfar, E., & Behzadi, A. (2018). Multi-objective optimization of biomass-based solid oxide fuel cell integrated with Stirling engine and electrolyzer. Energy Conversion and Management, 171, 1116–1133. https://doi.org/10.1016/ j.enconman.2018.06.061 25. Gholamian, E., Hanafizadeh, P., Habibollahzade, A., & Ahmadi, P. (2018). Evolutionary based multi-criteria optimization of an integrated energy system with SOFC, gas turbine, and hydrogen production via electrolysis. International Journal of Hydrogen Energy, 43(33), 16201–16214. https://doi.org/10.1016/j.ijhydene.2018.06.130 26. Safari, F., & Dincer, I. (2022). Assessment and multi-objective optimization of a vanadiumchlorine thermochemical cycle integrated with algal biomass gasification for hydrogen and power production. Energy Conversion and Management, 253, 115132. https://doi.org/10.1016/ j.enconman.2021.115132 27. Onder, G., Yilmaz, F., & Ozturk, M. (2020). Thermodynamic performance analysis of a copper–chlorine thermochemical cycle and biomass based combined plant for multigeneration. International Journal of Energy Research, 44(9), 7548–7567. https://doi.org/10.1002/er.5482 28. Cao, Y., Bani Hani, E. H., Mansir, I. B., & Safarzadeh, M. (2022). Design analysis and tri-objective optimization of a novel integrated energy system based on two methods for hydrogen production: By using power or waste heat. International Journal of Hydrogen Energy, 47(62), 26574–26588. https://doi.org/10.1016/j.ijhydene.2021.09.257 29. Lanjekar, P. R., Panwar, N. L., & Agrawal, C. (2023). A comprehensive review on hydrogen production through thermochemical conversion of biomass for energy security. Bioresource Technology Reports, 21, 101293. https://doi.org/10.1016/j.biteb.2022.101293 30. Shayan, E., Zare, V., & Mirzaee, I. (2018). Hydrogen production from biomass gasification; a theoretical comparison of using different gasification agents. Energy Conversion and Management, 159, 30–41. https://doi.org/10.1016/j.enconman.2017.12.096 31. Sotoodeh, A. F., Ahmadi, F., Ghaffarpour, Z., Ebadollahi, M., Nasrollahi, H., & Amidpour, M. (2022). Performance analyses of a waste-to-energy multigeneration system incorporated with thermoelectric generators. Sustainable Energy Technologies and Assessments, 49, 101649. https://doi.org/10.1016/j.seta.2021.101649

200

Z. Hajimohammadi Tabriz et al.

32. Ishaq, H., & Dincer, I. (2021). An efficient energy utilization of biomass energy-based system for renewable hydrogen production and storage. Journal of Energy Resources Technology, 144(1). https://doi.org/10.1115/1.4050875 33. Ishaq, H., & Dincer, I. (2020). A new energy system based on biomass gasification for hydrogen and power production. Energy Reports, 6, 771–781. https://doi.org/10.1016/j.egyr.2020.02.019 34. Çiğli Climate. https://climate.onebuilding.org/WMO_Region_6_Europe/TUR_Turkey/index. html#IDIZ_Izmir. Accessed. 35. Hajimohammadi Tabriz, Z., Mohammadpourfard, M., Gökçen Akkurt, G., & Zeinali Heris, S. (2023). Energy, exergy, exergoeconomic, and exergoenvironmental (4E) analysis of a new bio-waste driven multigeneration system for power, heating, hydrogen, and freshwater production: Modeling and a case study in Izmir. Energy Conversion and Management, 288, 117130. https://doi.org/10.1016/j.enconman.2023.117130 36. Salek, F., Moghaddam, A. N., & Naserian, M. M. (2018). Thermodynamic analysis and improvement of a novel solar driven atmospheric water generator. Energy Conversion and Management, 161, 104–111. https://doi.org/10.1016/j.enconman.2018.01.066 37. Buswell, A., & Hatfield, W. (1936). Anaerobic fermentations, state of Illinois. In Urbana, illinois: Department of registration and education, division of the state wate r survey. 38. Zhang, M., Chen, H., Zoghi, M., & Habibi, H. (2022). Comparison between biogas and pure methane as the fuel of a polygeneration system including a regenerative gas turbine cycle and partial cooling supercritical CO2 Brayton cycle: 4E analysis and tri-objective optimization. Energy, 257, 124695. https://doi.org/10.1016/j.energy.2022.124695 39. Loha, C., Chatterjee, P. K., & Chattopadhyay, H. (2011). Performance of fluidized bed steam gasification of biomass – Modeling and experiment. Energy Conversion and Management, 52(3), 1583–1588. https://doi.org/10.1016/j.enconman.2010.11.003 40. Zare, A. A. D., Yari, M., Mohammadkhani, F., Nami, H., & Desideri, U. (2022). Thermodynamic and exergoeconomic analysis of a multigeneration gas-to-X system based on fuel-rich combustion to produce power, hydrogen, steam and heat. Sustainable Cities and Society, 86, 104139. https://doi.org/10.1016/j.scs.2022.104139 41. Taheri, M. H., Khani, L., Mohammadpourfard, M., Aminfar, H., & Akkurt, G. G. (2022). Multiobjective optimization of a novel supercritical CO2 cycle-based combined cycle for solar power tower plants integrated with SOFC and LNG cold energy and regasification. International Journal of Energy Research, 46(9), 12082–12107. 42. Balafkandeh, S., Zare, V., & Gholamian, E. (2019). Multi-objective optimization of a tri-generation system based on biomass gasification/digestion combined with S-CO2 cycle and absorption chiller. Energy Conversion and Management, 200, 112057. https://doi.org/10. 1016/j.enconman.2019.112057 43. Liu, Z., & Ehyaei, M. A. (2022). Thermoeconomic and exergoenvironmental assessments of a combined micro-gas turbine and superheated Kalina cycles for cogeneration of heat and electrical power using biomass. International journal of Environmental Science and Technology, 19(11), 11233–11248. https://doi.org/10.1007/s13762-022-04329-y 44. Ahmadi, P., Dincer, I., & Rosen, M. A. (2014). Thermoeconomic multi-objective optimization of a novel biomass-based integrated energy system. Energy, 68, 958–970. https://doi.org/10. 1016/j.energy.2014.01.085 45. Akrami, E., Chitsaz, A., Nami, H., & Mahmoudi, S. M. S. (2017). Energetic and exergoeconomic assessment of a multigeneration energy system based on indirect use of geothermal energy. Energy, 124, 625–639. https://doi.org/10.1016/j.energy.2017.02.006 46. Ehyaei, M. A., Baloochzadeh, S., Ahmadi, A., & Abanades, S. (2021). Energy, exergy, economic, exergoenvironmental, and environmental analyses of a multigeneration system to produce electricity, cooling, potable water, hydrogen and sodium-hypochlorite. Desalination, 501, 114902. https://doi.org/10.1016/j.desal.2020.114902 47. Talebizadehsardari, P., et al. (2020). Energy, exergy, economic, exergoeconomic, and exergoenvironmental (5E) analyses of a triple cycle with carbon capture. Journal of CO2 Utilization, 41, 101258. https://doi.org/10.1016/j.jcou.2020.101258

8

Biomass-Based Polygeneration Systems with Hydrogen Production

201

48. Mohammadpourfard, M., Hajimohammadi Tabriz, Z., & Khani, L. (2022). Design and assessment of a new dual solid oxide fuel cell – Gas turbine hybrid system. Journal of Energy Management and Technology. https://doi.org/10.22109/jemt.2022.329698.1372 49. Khani, L., Mohammadpour, M., Mohammadpourfard, M., Heris, S. Z., & Akkurt, G. G. (2022). Thermodynamic design, evaluation, and optimization of a novel quadruple generation system combined of a fuel cell, an absorption refrigeration cycle, and an electrolyzer. International Journal of Energy Research, 46(6), 7261–7276. https://doi.org/10.1002/er.7634 50. Cartmell, E., et al. (2006). BiosolidsA fuel or a waste? An integrated appraisal of five co-combustion scenarios with policy analysis. Environmental Science & Technology, 40(3), 649–658. https://doi.org/10.1021/es052181g 51. Almahdi, M., Dincer, I., & Rosen, M. A. (2016). A new solar based multigeneration system with hot and cold thermal storages and hydrogen production. Renewable Energy, 91, 302–314. https://doi.org/10.1016/j.renene.2016.01.069 52. Adewusi, S. A., & Zubair, S. M. (2004). Second law based thermodynamic analysis of ammonia–water absorption systems. Energy Conversion and Management, 45(15), 2355–2369. https://doi.org/10.1016/j.enconman.2003.11.020

Chapter 9

Integration of Solar PV and GH2 in the Future Power Systems Hassan Majidi-Gharehnaz , Hossein Biabani , Ali Aminlou Mohammad Mohsen Hayati , and Mehdi Abapour

9.1

,

Introduction

Future power systems are evolving toward a more stable, reliable, clean, and efficient energy system [1]. One of the basic needs for this evolution is the integration of distributed renewable energy sources, such as wind and solar power [2]. The use of green hydrogen (GH2) in future power systems is of great importance due to its decarbonization and energy storage capabilities [3]. In the GH2 mechanism, hydrogen is produced by using renewable energy sources and does not emit carbon dioxide or other harmful greenhouse gases during its production or use. Additionally, GH2 can be used as an energy storage medium, allowing excess energy from renewable sources to be stored and used later when needed [4]. Overall, the use of GH2 in future power systems can contribute to a more sustainable, reliable, and efficient energy system that can meet the growing demand for electricity while reducing greenhouse gas emissions and addressing climate change [5]. Solar energy is a widely available energy source across most parts of the world. However, the amount of energy that can be harvested from it varies throughout the day depending on the amount of sunlight available. On clear days, the energy production starts increasing in the morning and peaks at noon before decreasing in the evening. Due to the variability of solar energy production, relying solely on it can be unreliable. Therefore, it is essential to supplement it with flexible resources to ensure a steady supply of electricity [6]. One promising solution is to integrate green hydrogen technology with renewable energy sources [7]. This allows excess renewable energy to be stored as green hydrogen, which is an environmentally friendly form of energy storage [8]. During times when renewable energy production exceeds

H. Majidi-Gharehnaz (✉) · H. Biabani · A. Aminlou · M. M. Hayati · M. Abapour Faculty of Electrical and Computer Engineering, Energy Systems Research Institute (ESRI), Smart Energy Systems Lab, University of Tabriz, Tabriz, Iran e-mail: [email protected]; [email protected] © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 V. Vahidinasab et al. (eds.), Green Hydrogen in Power Systems, Green Energy and Technology, https://doi.org/10.1007/978-3-031-52429-5_9

203

204

H. Majidi-Gharehnaz et al.

demand, this stored energy can be injected into the gas network or converted back into electricity via a fuel cell when renewable energy production is insufficient [9]. The studies presented in this chapter aim to investigate the operational and planning aspects of microgrids, as exemplified by the microgrid depicted in Fig. 9.1 Within this illustration, the microgrid incorporates both DC and AC loads, as well as a solar power generation source, all integrated with GH2 technology. This microgrid is capable of energy exchange with both electricity and gas networks. The seamless integration of this microgrid with the gas network is achieved through the utilization of GH2 technologies.

9.2

Status Quo, Challenges, and Outlook

The current state of GH2 integration with solar resources marks an essential milestone in the transition towards sustainable and decarbonized energy systems. While the concept of GH2 has gained considerable attention and recognition for its potential to revolutionize the energy landscape, its widespread implementation is still in its early stages. Various pilot projects, research studies, and experimental deployments have showcased the viability of combining solar energy generation with GH2 production and storage. These endeavors have shed light on the technical, economic, and environmental aspects of such integrated systems, contributing to a growing body of knowledge [10]. In the existing landscape, GH2 integration with solar resources has demonstrated its potential to address the intermittency of solar energy production and enhance the stability of power grids. The successful operation of microgrids and energy systems that combine solar panels, electrolyzers, fuel cells, and hydrogen storage represents a promising step toward achieving a cleaner and more reliable energy supply. These integrated systems provide valuable insights into optimizing energy production, storage, and distribution to maximize efficiency and minimize costs [11]. Despite the progress made, the integration of GH2 with renewable resources also faces several challenges that must be addressed to realize its full potential. Some of the key challenges include [12]: . Technological Development: Further advancements in electrolysis technology and fuel cell efficiency are required to optimize the conversion processes and reduce energy losses during hydrogen production and utilization. . Economic Viability: The current costs associated with GH2 technology, including the production, storage, and distribution of hydrogen, need to be reduced to make the integrated systems economically competitive with conventional energy sources. . Infrastructure and Storage: Establishing the necessary infrastructure for hydrogen storage, transportation, and distribution on a large scale presents significant challenges. The development of cost-effective and efficient storage solutions is crucial for the success of GH2 integration.

Compressor

Electrolyzer

DC Bus

Fig. 9.1 Schematic of solar and GH2 integrated microgrid

Gas Network

Gas Link

Electrical Link

Solar Panel

DC Load

Hydrogen Tank Storage

Fuel Cell

AC Bus

Electrical Network

AC Load

9 Integration of Solar PV and GH2 in the Future Power Systems 205

206

H. Majidi-Gharehnaz et al.

. Regulatory Frameworks: Clear and supportive regulatory frameworks are essential to incentivize the deployment of integrated solar and GH2 systems. Policies that promote renewable energy adoption, facilitate grid integration and ensure fair market access are needed. . System Integration: Integrating GH2 with existing energy grids requires careful planning and coordination to ensure seamless operation, minimize grid imbalances, and optimize the use of renewable resources. . Public Awareness and Acceptance: Public awareness and acceptance of GH2 as a viable energy solution, along with education about its benefits and safety, are vital for gaining public support and overcoming potential resistance. Looking ahead, the integration of GH2 with solar resources holds immense promise for advancing the transition to sustainable and carbon-neutral energy systems. As technological innovations continue to drive down the costs of renewable energy sources and hydrogen production, the economic feasibility of integrated systems is expected to improve. Ongoing research and development efforts aimed at enhancing the efficiency and reliability of GH2 technology will contribute to overcoming the current challenges. The integration of GH2 with solar resources aligns with global efforts to reduce greenhouse gas emissions, combat climate change, and ensure energy security. The potential for energy storage and grid stabilization offered by GH2 can play a pivotal role in enabling higher penetration of variable renewable energy sources, such as solar and wind, into the energy mix [13]. In the coming years, it is anticipated that pilot projects will evolve into largerscale demonstrations and commercial implementations of GH2-integrated energy systems. Governments, industries, and research institutions will collaborate to accelerate technological advancements, drive down costs, and establish supportive policy frameworks. The success of these endeavors will contribute to a more resilient, sustainable, and low-carbon energy future, unlocking new opportunities for economic growth and environmental preservation.

9.3

Related Works

The research conducted on green hydrogen as an integrated component of energy systems in recent years can generally be divided into two main categories. The first category includes studies that explore the operation of energy systems integrated with green hydrogen, either as an energy storage system or as a means of connecting the electricity system to the gas network in the presence of renewable energy sources. The second category includes studies that investigate the planning and design of power systems integrated with renewable energy sources to enhance reliability. In studies conducted in the first category, which explored the operation of energy systems integrated with green hydrogen and renewable energy sources, [14]

9

Integration of Solar PV and GH2 in the Future Power Systems

207

proposed a Markov decision process model to determine the optimal control strategy for a system that combines offshore wind turbines, an electrolyzer, a storage facility, and a fuel cell. Similarly, [15] proposes a novel structure for multi-microgrids (MMG) based on power, heat, and hydrogen, and applies a robust optimization method to deal with uncertainties in electricity prices and demands. Cai et al. [16] combine stochastic programming and robust optimization to optimize the operation of power systems with wind generation, hydrogen storage systems, and demand response programs under uncertainty. Meanwhile, [17] explores the use of renewable energy sources and peer-to-peer energy trading to reduce electricity costs and risk for residential buildings. Heris et al. [18] evaluate the impact of power-tohydrogen conversion and hydrogen storage technology in multi-carrier energy systems and show that hydrogen storage can reduce wind power curtailment and improve the operation of the system. Liang et al. [19] present an optimization strategy for a wind-solar renewable energy hydrogen production system used to power hydrogen buses, while [20] reviews the potential of hydrogen energy storage for achieving a 100% renewable electricity system and compares hydrogen storage with battery storage. Ruiming [21] presents an algorithm for optimizing the day-ahead power dispatching of an integrated energy system that combines wind, solar, and hydrogen energy sources. Overall, these papers explore various aspects of using hydrogen as a means of storing and converting excess electricity from wind and solar farms, and optimizing its use in different energy systems to reduce cost and carbon emissions. The second part of the studies explores various related works that focus on the planning of renewable energy sources integrated with hydrogen technology for sustainable and reliable energy production. Chen et al. [22] propose a hybrid hydrogen/electricity refueling station that utilizes surplus wind power to charge electric trucks and produce “green hydrogen” for hydrogen trucks. Cao et al. [23] present a model for optimizing the capacity of an island microgrid that uses wind, solar, and green hydrogen as energy sources. Cao et al. [24] compare solar hydrogen production via water electrolysis between Morocco and Southern Europe. Meanwhile, [25] examines the potential of power-to-gas (P2G) technology to recover excess electricity from renewable energy sources (RES) in Italy. Hatti et al. [26] present a hybrid power system that combines photovoltaic (PV) and proton exchange membrane fuel cell (PEMFC) generators to supply a load via a DC-Bus. Moradpoor et al. [27] examine the feasibility of producing green hydrogen from water electrolysis using renewable electricity for oil refining in Finland. Kaviani et al. [28] aim to optimize the design of a hybrid system that uses wind, solar, and hydrogen energy to meet the power demand in a reliable and cost-effective way. Meanwhile [29] proposes a microgrid system that uses photovoltaic power generation, battery, and hydrogen energy as energy storage devices. Farhani et al. [30] present a novel design of a hydrogen production station that uses solar energy to power a proton exchange membrane (PEM) electrolyzer, concluding that it is a promising solution for green hydrogen production and can contribute to the development of a hydrogen economy, while [31] presents a method to design a standalone hybrid green power system that can supply electricity and hydrogen to a local demand with on-line charging of electric vehicles.

208

H. Majidi-Gharehnaz et al.

9.4

Operation of Solar and GH2 Integrated Energy System

As mentioned earlier, solar farms produce energy during the day, and at night, there is no production. However, excess energy produced during the day can be cleanly stored in the form of hydrogen gas using GH2 technology. This hydrogen can be injected into the gas or electricity grid for later use. In this study, we investigate a microgrid consisting of solar farms and GH2 technology to minimize its maintenance cost. Figure 9.1 shows the schematic of this microgrid, which consists of two AC and DC buses connected to gas and electricity networks. The main components of this network are solar panels, inverters, electrolyzers, fuel cells, and hydrogen storage tanks. In the following sections, we will discuss the modeling of these components in the network.

9.4.1

Solar Panels

The modeling of solar panel power generation under specific levels of radiation is as follows [31]: PPV = ir ðt, δPV Þ × PPV,rated × ηPV,conv

ð9:1Þ

ir ðt, δPV Þ = ir V ðt Þ × cosðδPV Þ þ r H ðt Þ × sinðδPV Þ

ð9:2Þ

where, PPV is the output power of the arrays, ir(t, δPV) is the amount of sunlight received at a certain time and epsilon angle, PPV, rated is the nominal power of the solar panels, eta is the conversion efficiency of the panels, irV is the vertical solar radiation, and irH is the horizontal solar radiation.

9.4.2

Inverter

Inverters convert DC electricity to AC, or vice versa. The mathematical relationships governing their operation are as follows: PPv,inv2DC ðt Þ = PPV ðt Þ × ηinv,DCDC Pinv2AC ðt Þ = ðPPV ,inv2DC ðt Þ þ PFC ðt Þ - PEL ðt Þ - PDCload ðt ÞÞ × ηinv,ACDC Pinv2DC ðt Þ = ðPEN ðt Þ - PACload ðt ÞÞ × ηinv,ACDC

ð9:3Þ ð9:4Þ ð9:5Þ

Where, PPv, inv2DC is the power sent by the solar panels to the DC bus, ηinv, DCDC is the efficiency of the DC-DC inverter, ηinv, ACDC is the efficiency of the AC-DC inverter, Pinv2AC is the power sent to the AC bus, PFC is the fuel cell power, PEL is the

9

Integration of Solar PV and GH2 in the Future Power Systems

209

electrolysis power, PDCload is the DC load power, PACload is the AC load power, Pinv2DC is the power sent to the DC bus, and PEN is the power from/to the power grid.

9.4.3

Electrolyzers

The electrolyzer converts electrical power into hydrogen and stores it in a hydrogen tank. The mathematical relationship for electrolyzer is as follows: PEL2HST ðt Þ = PEL ðt Þ × ηEL

ð9:6Þ

where PEL2HST is the power sent by the electrolyzer to HST and, ηEL is the efficiency of the electrolyzer.

9.4.4

Fuel Cells

The Fuel cells convert hydrogen power into electricity and send it in electrical grid. The mathematical relationship for Fuel cells is as follows: PFC ðt Þ = PHST2FC ðt Þ × ηFC × ηHST

ð9:7Þ

where PHST2FC is the power sent by HST to the fuel cell, ηFC is the efficiency of the fuel cell and, ηHST is the efficiency of the HST.

9.4.5

Hydrogen Storage Tanks

Hydrogen storage tanks store excess hydrogen and inject it into the network when needed. The mathematical relationship for Fuel cells is as follows: EHST ðt Þ = E HST ðt - 1Þ þ PEL2HST ðtÞ × Δt - PHST2FC ðtÞ × Δt × ηHST - E sold ðt Þ × ηHST

ð9:8Þ mHST ðt Þ =

E HST ðt Þ HHV H 2

ð9:9Þ

Where, EHST(t) is the energy stored in HST at the time of t, mHST(t) is the hydrogen mass stored in HST at the time of t, Esold(t) is sold energy to gas network, and, HHV H 2 is the hydrogen Higher Heating Value and is equal to 39.7kWh/kg [31].

210

9.4.6

H. Majidi-Gharehnaz et al.

Objective Function

The objective function to optimize the microgrid is to minimize operating costs. The microgrid purchases electricity from the power grid during certain hours and sells electricity and hydrogen during other hours to supply its load. The operating cost is calculated as the cost of buying electricity minus the profits gained from the sale of electricity and hydrogen. The mathematical relationship for fuel cells is as follows: TOC =

OT

ðPEN ðt Þ × EPðt Þ - mGN ðt Þ × HPðt ÞÞ

ð9:10Þ

t=1

Where, TOC is the total microgrid operation cost, OT is operation time and usually it is equal to 24 h, EP is the electrical price, mGN is hydrogen sold to the gas network, and HP is the hydrogen price.

9.4.7

Constraints

The most important constraints of this problem are the equality of production and consumption power at every time and the requirement to not exceed the capacity of the equipment. The mathematical relationship for these constraints is as follows: PPV ðt Þ × ηinv,DCDC þ PEN ðt Þ × ηinv,ACDC þ PFC = PEL þ PDCload þ PACload × ηinv,ACDC

ð9:11Þ 0 ≤ mHST ðt Þ ≤ CHST

ð9:12Þ

0 ≤ PFC ≤ MPFC

ð9:13Þ

0 ≤ PEL ≤ MPEL

ð9:14Þ

where CHST is hydrogen storage capacity, MPFC is the maximum power output of the fuel cell, and MPEL is maximum power output of the electrolyzer.

9.4.8

Case Study

The first case study examines the operation of the microgrid network shown in Fig. 9.1. to minimize its operation cost. The study uses data from Germany in 2015 [32], including solar farm production, electricity consumption, and market prices. Specifically, the data from the third day of January is used, and peak consumption and solar production power values are adjusted to 200 kW and 600 kW, respectively, using a constant numerical multiplier. It should be noted that only AC load is

9

Integration of Solar PV and GH2 in the Future Power Systems

211

included in this analysis. Figures 9.2, 9.3, and 9.4 display the profiles of consumed power, power production of the solar farm and the price curve of the electricity market, respectively. The study assumes an electrolyzer efficiency of 0.75 with a power capacity of 400 kW, and a fuel cell efficiency of 0.50 with a capacity of 400 kW. The efficiencies of the HST and inverter are assumed to be 0.95 and 0.9, respectively, while the capacity of the HST is set at 75 kilos [31]. The price of green hydrogen is set at one dollar. The optimization of this problem was carried out using the PYOMO library in Python. The code used for this optimization is available in [33]. The optimal expense of operating the microgrid for a single day is -38.13 US dollars, representing a profit

Fig. 9.2 The microgrid consumed power AC power

Fig. 9.3 Power production of the solar farm

212

H. Majidi-Gharehnaz et al.

Fig. 9.4 Price of electricity

Fig. 9.5 Microgrid purchased and sold power

of 38.13 US dollars. Figure 9.5 illustrates the power purchased and sold. The hydrogen storage levels are shown in Fig. 9.6, while Fig. 9.7 displays the power output of the fuel cell and electrolyzer. Finally, Fig. 9.8 shows the sales of green hydrogen. As evident from the figures, the microgrid purchases electricity from the grid during the early hours when it is relatively cheap, to supply its load and store excess energy. During hours 17, 18, and 19, when the energy cost is high, the microgrid sells some of its excess energy in the form of electricity to the power grid and transfers the remaining amount to the gas grid.

9

Integration of Solar PV and GH2 in the Future Power Systems

213

Fig. 9.6 Hydrogen storage levels

Fig. 9.7 Power output of the fuel cell and electrolyzer

9.5

Planning of Solar and Gh2 Integrated Energy System

Power system planning involves developing plans and procedures to assure the reliable, efficient, and cost-effective supply of energy to fulfill consumers’ present and future needs. In planning the integration of green hydrogen with solar resources, the goal is to optimize the placement angle of solar panels, the production capacity of solar resources, the capacity of the electrolyzer, the capacity of the fuel cell, and the

214

H. Majidi-Gharehnaz et al.

Fig. 9.8 Green hydrogen sold

capacity of the electric storage. This optimization is carried out to ensure that the resulting microgrid can reliably supply a certain percentage of its load. Any shortfall in supply from the microgrid can be met through the upstream network of the microgrid. The main factor affecting the reliability of the microgrid is the potential loss of supply sources for the microgrid load.

9.5.1

Reliability Indexes

The paper [34] discusses the probability that a device works properly to supply load demand under operational conditions within a certain time. The literature lists several metrics for power system reliability, including loss of load expectation (LOLE), expected energy not supplied (EENS), loss of energy expectation (LOEE), loss of power supply probability (LPSP), and equivalent loss factor (ELF), which are listed below: LOLE =

N

E½LOLðt Þ]

ð9:15Þ

1

where LOL loss of load, and N is number of load loss and calculate as flows: T s × Ps

E½LOL] =

ð9:16Þ

s2S

where S is the set of all possible states of system failures, Ts is the duration of loss of load at state s, and the probability of state s occurrence.

9

Integration of Solar PV and GH2 in the Future Power Systems

LOEE =

N

E ½LOE ðt Þ]

215

ð9:17Þ

1

where LOE is loss of energy and calculated as flows: Qs × Ps

E½LOE ] =

ð9:18Þ

s2S

where Qs is the duration of loss of load at state s. LPSP =

LOEE N t=1

ð9:19Þ

D ðt Þ

where D(t) load demand. ELF =

1 N

N t=1

Qðt Þ Dðt Þ

ð9:20Þ

It is important to note that ELF provides a greater level of detail on both the number of power outages and the amount of load demand that has not been met [28].

9.5.2

Objective Function

To address multiple HGPS design analysis methodologies, four objective functions are constructed as one objective function utilizing the weighting factor technique with equal weighting to treat all objectives equally [35]. The objective function of this problem is to minimize the total cost of the system, which is composed of three components: investment costs, energy purchase costs from external sources, and penalty costs for breaching the reliability requirements. The objective function is formulated as follows: minfC total g =

NPC i þ NPC loss þ Penalty

ð9:21Þ

i

where Ctotal is the total cost, NPCi the net present cost of i-th equipment, NPCloss is Net present cost of unmet load, and Penalty is penalty term of optimization.

216

H. Majidi-Gharehnaz et al.

NPC i = N i × ðCC i þ RC i × K þ MC i × PWAÞ

ð9:22Þ

where CCi is capital cost, RCi is replacement cost, MCi is operation and maintenance cost, and K and PWA are two economic factors that enable the conversion of replacement and operational costs into a single present cost. The factors are defined as follows: K=

Yi

1 ð1 þ ir ÞL × n

ð9:23Þ

ð1 þ ir ÞR - 1 ir ð1 þ ir ÞR

ð9:24Þ

ðir nom - f Þ ð1 þ f Þ

ð9:25Þ

n=1

PWAi = ir =

where Yi is the number of replacements, L is useful time, ir and irnom is real and nominal interest rate respectively, R is project useful time (here is 20 years), and f is inflation rate. NPC Loss = LOEE × C loss × PWA

ð9:26Þ

where Closs is the average cost of unmet load. The penalty part is also considered a coefficient proportional to the amount of the violation.

9.5.3

Constraints

This problem has three main conditions that must be met. Firstly, the ELF must be less than a certain value. Secondly, the equipment capacity cannot be negative. Finally, the panel angle must be chosen between 0 and 90°. The mathematical expressions for these conditions are given below: ELF ≤ ELF max

ð9:27Þ

0 ≤ Capi

ð9:28Þ

0 ≤ θPV ≤

π 2

ð9:29Þ

9

Integration of Solar PV and GH2 in the Future Power Systems

9.5.4

217

Solving Method

The design of the integrated solar farm microgrid with GH2 requires the solution of a non-linear objective function, which can be solved using non-linear optimization methods. Some of the methods that can be used for non-linear optimization include Gradient-Based Methods such as the steepest descent method, conjugate gradient method, quasi-Newton methods, and Newton’s method, which are efficient for small or medium-sized problems, but may converge slowly or fail for large or complex problems. Nonlinear Programming methods, such as sequential quadratic programming, interior point methods, and augmented Lagrangian methods, are often used for large or complex problems but require careful problem formulation and significant computational resources. Constraint Satisfaction Methods, such as linear programming, quadratic programming, and mixed-integer programming, are used to solve non-linear optimization problems that involve constraints and are often used in engineering, finance, and operations research. Direct Search Methods, such as pattern search, Nelder-Mead simplex, and Powell’s method, rely on iterative search procedures and are often used for problems with non-differentiable or discontinuous objective functions. Evolutionary Algorithms, such as genetic algorithms, particle swarm optimization, and simulated annealing, can be effective for non-linear optimization problems with a large number of variables, non-convex objective functions, and noisy or stochastic environments. Given the complexities inherent in this problem, we recommend using meta-heuristic and evolutionary methods to solve the nonlinear objective function that has been modeled here. The population-based meta-heuristic Gray Wolf Optimization (GWO) algorithm was inspired by wild gray wolf social behavior and hunting techniques [36]. The GWO algorithm is based on the behavior of gray wolves, which are known for their social hierarchy and cooperation in hunting prey. The algorithm initializes a population of candidate solutions, called “wolves,” and updates their positions using three main steps: hunting, searching, and attacking. During the hunting phase, the algorithm identifies the best solution among the wolves, known as the “alpha” wolf. During the searching phase, the peripheral wolves search for prey in the vicinity of the alpha, beta, and delta wolves, using a random search strategy. During the attacking phase, the alpha, beta, and delta wolves update their positions based on the positions of the other wolves in the population. This process continues until a stopping criterion is met. The GWO algorithm is effective for a range of optimization problems, including continuous, discrete, and mixed-integer optimization problems. However, the performance of the algorithm can be sensitive to the specific problem being solved and the parameters used in the algorithm [36].

218

9.5.5

H. Majidi-Gharehnaz et al.

Case Study

In this case study, we will design a micro-grid based on Fig. 9.1, but without a connection to the gas network. Two scenarios will be considered, in which the micro-grid relies on the upstream network for 1% or 10% of its load per year and their results will be compared. In this design, the aim is to optimize the capacity of solar power sources, the angle of panel placement, and the capacity of the electrolyzer, fuel cell, and HST. The cost information for all equipment is presented in Table 9.1 For the annual network load, the German load of 2015 [32] has been utilized. To make this load appropriate for the current problem, a fixed number has been multiplied by the annual peak, and the load has been moved to 200 kW. The radiation and radiation angle data of the city of San Antonio are used to calculate the annual solar radiation [37]. The upstream electricity price is set at 0.4, and the real interest rate is assumed to be 6% per year. The optimization is performed using the GWO method in Python software. The Python codes related to this optimization are available at [38]. By conducting optimization for two scenarios in which the microgrid supplies 99% and 90% of its own load, respectively, the convergence curves of the GWO algorithm are depicted in Figs. 9.9 and Fig. 9.10, respectively. As depicted in the figures, the algorithm has converged successfully for both scenarios. The hydrogen storage capacity in the HST for these two scenarios is presented in Figs. 9.11 and Fig. 9.12, respectively. Tables 9.2 and 9.3 display the optimization results and costs of the optimization period, respectively. Taking into consideration that the designed network relies solely on solar resources, significant planning costs are involved. However, if wind turbines are integrated into the network, the presence of wind generators capable of producing energy at night would lead to a reduction in planning costs. The investigation of two scenarios has revealed that utilizing the aforementioned network capacities can lower planning costs. Nevertheless, it is important to note that in certain cases, depending on the analyzed scenario, relying on the upstream network may prove to be more expensive or even impossible to establish a connection.

9.6

Conclusion

Green hydrogen, in combination with renewable energy sources, can enhance the flexibility of the network across various conditions. Among the renewable energy sources, solar energy production fields demonstrate significant potential for integration with green hydrogen (GH2). During daylight hours, solar energy is converted into electricity using solar panels. A portion of this electricity is utilized within the grid, while the surplus is directed to an electrolyzer for conversion into hydrogen. The hydrogen is subsequently stored in a hydrogen storage tank (HST). The stored

Equipment PV array Electrolyzer HST Fuel cell Inverter

Capital cost (US$/unit) 2000 2000 1300 4000 800

Replacement cost (US$/unit) 1700 1500 1200 3500 750

Table 9.1 The cost and reliability information of equipment Maintenance cost (US$/unit-year) 20 25 15 300 8

R (year) 20 20 20 5 15

η (%) – 75 95 50 90

Unit (kW) 1 1 1 1 1

Availability 96 100 100 100 99.89

9 Integration of Solar PV and GH2 in the Future Power Systems 219

220

H. Majidi-Gharehnaz et al.

Fig. 9.9 Gray wolf optimization algorithm convergence curve for microgrid supplies 99% of its own load scenario

Fig. 9.10 Gray wolf optimization algorithm convergence curve for microgrid supplies 90% of its own load scenario

hydrogen can be reintegrated into the electricity cycle or injected into the gas network during periods of low electrical energy demand, utilizing fuel cells for energy conversion. Studies on GH2 integration with solar resources can be broadly categorized into two areas: operation and planning. In the operational domain, the objective is to optimize the operational strategy of GH2-integrated networks, which may encompass sector coupling. In the planning realm, the aim is to design networks with GH2 that meet the network’s requirements. In the first case study, the operation of a microgrid connected to the electricity and gas networks, incorporating GH2 and a wind farm, was conducted with the aim of cost reduction. This study demonstrated

9

Integration of Solar PV and GH2 in the Future Power Systems

221

Fig. 9.11 Hydrogen stored in HST for microgrid supplies 99% of its own load scenario

Fig. 9.12 Hydrogen stored in HST for microgrid supplies 90% of its own load scenario Table 9.2 Optimization results for both scenarios Variable Production capacity of solar resources (kW) Panels angel (°) HST capacity (kg) Fuel cell capacity (kW) Electrolyzer capacity (kW)

Microgrid supplies 90% of its own load scenario 2687

Microgrid supplies 99% of its own load scenario 1666

27 426 200 2007

48 183 200 1491

222

H. Majidi-Gharehnaz et al.

Table 9.3 Scenarios cost Variable Total cost (MUS$) Capital cost (MUS $) Replacement cost (MUS$) Maintenance cost (MUS$) Unmet load cost (MUS$)

Microgrid supplies 90% of its own load scenario 16.44 10.24

Microgrid supplies 99% of its own load scenario 13.34 7.35

4.26

4.26

1.37

1.17

0.06

0.58

how excess solar-generated electricity can be converted into GH2, which can be subsequently converted back into electricity during periods of demand, and any remaining hydrogen can be sold to the gas grid. In the second case study, the design of a microgrid incorporating a solar farm and GH2 integration was undertaken while considering reliability indicators. GH2 planning can be performed under various scenarios, resulting in improvements in different network parameters, including reliability. The Python program used for the case studies described in this chapter has been shared on GitHub.

References 1. Aminlou, A, Hayati, M. M., & Zare, K. (2023). Local peer-to-peer energy trading evaluation in micro-grids with centralized approach. In 2023 8th International conference on technology and energy management (ICTEM) (pp. 1–6). IEEE. 2. Muttaqi, K. M., Islam, M. R., & Sutanto, D. (2019). Future power distribution grids: Integration of renewable energy, energy storage, electric vehicles, superconductor, and magnetic bus. IEEE Transactions on Applied Superconductivity, 29(2), 1–5. 3. Khan, I. (2018). Importance of GHG emissions assessment in the electricity grid expansion towards a low-carbon future: A time-varying carbon intensity approach. Journal of Cleaner Production, 196, 1587–1599. 4. Nastasi, B., & Basso, G. L. (2016). Hydrogen to link heat and electricity in the transition towards future smart energy systems. Energy, 110, 5–22. 5. Maestre, V., Ortiz, A., & Ortiz, I. (2021). Challenges and prospects of renewable hydrogenbased strategies for full decarbonization of stationary power applications. Renewable and Sustainable Energy Reviews, 152, 111628. 6. Hayati, M. M., Aminlou, A., Zare, K., & Abapour, M. (2023). A two-stage stochastic optimization scheduling approach for integrating renewable energy sources and deferrable demand in the spinning reserve market. In 2023 8th international conference on technology and energy management (ICTEM) (pp. 1–7). IEEE. 7. Gielen, D., Taibi, E., & Miranda, R. (2019). Hydrogen: A renewable energy perspective. International Renewable Energy Agency. 8. Aminlou, A., Hayati, M., & Zare, K. (2023). ADMM-based fully decentralized Peer to Peer energy trading considering a shared CAES in a local community. Journal of Energy Management and Technology.

9

Integration of Solar PV and GH2 in the Future Power Systems

223

9. Mukelabai, M. D., Wijayantha, U. K., & Blanchard, R. E. (2022). Renewable hydrogen economy outlook in Africa. Renewable and Sustainable Energy Reviews, 167, 112705. 10. Lopes, J. A. P., et al. (2020). The future of power systems: Challenges, trends, and upcoming paradigms. Wiley Interdisciplinary Reviews: Energy and Environment, 9(3), e368. 11. Han, J., et al. (2023). Hydrogen-powered smart grid resilience. Energy Conversion and Economics, 4(2), 89–104. 12. Tholen, L., et al. (2021). The green hydrogen puzzle: Towards a german policy framework for industry. Sustainability, 13(22), 12626. 13. Guevara-Ramirez, W., Martinez-de-Alegria, I., & Rio-Belver, R. M. (2023). Evolution of the conceptualization of hydrogen through knowledge maps, energy return on investment (EROI) and national policy strategies. Clean Technologies and Environmental Policy, 25(1), 69–91. 14. Schrotenboer, A. H., Veenstra, A. A., uit het Broek, M. A., & Ursavas, E. (2022). A green hydrogen energy system: Optimal control strategies for integrated hydrogen storage and power generation with wind energy. Renewable and Sustainable Energy Reviews, 168, 112744. 15. Mansour-Saatloo, A., et al. (2021). Robust decentralized optimization of multi-microgrids integrated with power-to-X technologies. Applied Energy, 304, 117635. 16. Cai, T., Dong, M., Liu, H., & Nojavan, S. (2022). Integration of hydrogen storage system and wind generation in power systems under demand response program: A novel p-robust stochastic programming. International Journal of Hydrogen Energy, 47(1), 443–458. 17. Guo, Q., Chen, Y., Xu, Y., Nojavan, S., Bagherzadeh, H., & Valipour, E. (2022). Integration of hydrogen storage system and solar panels in smart buildings. International Journal of Hydrogen Energy, 47(44), 19237–19251. 18. Heris, M.-N., et al. (2020). Evaluation of hydrogen storage technology in risk-constrained stochastic scheduling of multi-carrier energy systems considering power, gas and heating network constraints. International Journal of Hydrogen Energy, 45(55), 30129–30141. 19. Liang, T., Sun, B., Mi, D., & Sha, Z. (2022). Operation optimization of wind-solar complementary hydrogen production system based on MOGEO. In 2022 7th International conference on power and renewable energy (ICPRE) (pp. 821–831). IEEE. 20. Colbertaldo, P., Agustin, S. B., Campanari, S., & Brouwer, J. (2019). Impact of hydrogen energy storage on California electric power system: Towards 100% renewable electricity. International Journal of Hydrogen Energy, 44(19), 9558–9576. 21. Ruiming, F. (2019). Multi-objective optimized operation of integrated energy system with hydrogen storage. International Journal of Hydrogen Energy, 44(56), 29409–29417. 22. Chen, X., et al. (2023). A novel LH2/GH2/battery multi-energy vehicle supply station using 100% local wind energy: Technical, economic and environmental perspectives. Energy, 270, 126871. 23. Cao, Y., Huang, Y., Liu, R., Zhuo, Y., Liu, P., & Nie, J. (2021). Capacity optimization of multienergy complementary microgrid considering green hydrogen system. In 2021 4th International conference on energy, electrical and power engineering (CEEPE) (pp. 765–769). IEEE. 24. Touili, S., Merrouni, A. A., El Hassouani, Y., Amrani, A.-I., & Azouzoute, A. (2019). A technoeconomic comparison of solar hydrogen production between Morocco and Southern Europe. In 2019 International conference on wireless technologies, embedded and intelligent systems (WITS) (pp. 1–6). IEEE. 25. Guandalini, G., Robinius, M., Grube, T., Campanari, S., & Stolten, D. (2017). Long-term power-to-gas potential from wind and solar power: A country analysis for Italy. International Journal of Hydrogen Energy, 42(19), 13389–13406. 26. Hatti, M., Meharrar, A., & Tioursi, M. (2011). Power management strategy in the alternative energy photovoltaic/PEM fuel cell hybrid system. Renewable and Sustainable Energy Reviews, 15(9), 5104–5110. 27. Moradpoor, I., Syri, S., & Santasalo-Aarnio, A. (2023). Green hydrogen production for oil refining–Finnish case. Renewable and Sustainable Energy Reviews, 175, 113159.

224

H. Majidi-Gharehnaz et al.

28. Kaviani, A. K., Riahy, G., & Kouhsari, S. M. (2009). Optimal design of a reliable hydrogenbased stand-alone wind/PV generating system, considering component outages. Renewable Energy, 34(11), 2380–2390. 29. Haodong, M., Chunhua, L., Yu, Z., Kang, Y., & Linyao, F. (2022). Capacity configuration of solar-based battery-hydrogen hybrid energy storage for microgrids. In 2022 China Automation Congress (CAC) (pp. 5346–5350). IEEE. 30. Farhani, S., Grissa, H., & Bacha, F. (2021). Hydrogen production station using solar energy. In 2021 IEEE 2nd international conference on signal, control and communication (SCC) (pp. 301–306). IEEE. 31. Ahangar, H. G., Yew, W. K., & Flynn, D. (2023). Smart local energy systems: Optimal planning of stand-alone hybrid green power systems for on-line charging of electric vehicles. IEEE Access, 11, 7398–7409. 32. Bach, P.-F. http://www.pfbach.dk/ 33. Majidi, H. Solar green hydrogen integration operation. https://github.com/HassanMajidi/ solar_green_hydrogen_integration_operation 34. Gharavi, H., & Ardehali, M. (2013). Imperialist competitive algorithm for optimal design of on-grid hybrid green power system integrated with a static compensator for reactive power management. Journal of Renewable and Sustainable Energy, 5(1), 013115. 35. Jahannoosh, M., Nowdeh, S. A., Naderipour, A., Kamyab, H., Davoudkhani, I. F., & Klemeš, J. J. (2021). New hybrid meta-heuristic algorithm for reliable and cost-effective designing of photovoltaic/wind/fuel cell energy system considering load interruption probability. Journal of Cleaner Production, 278, 123406. 36. Mirjalili, S., Mirjalili, S. M., & Lewis, A. (2014). Grey wolf optimizer. Advances in Engineering Software, 69, 46–61. 37. NASA. https://power.larc.nasa.gov/data-access-viewer/ 38. Majidi, H.. Planning of solar and GH2 integrated microgrid. https://github.com/HassanMajidi/ Planning_of_Solar_GH2_Integrated_Microgrid

Chapter 10

GH2 Networks: Production, Supply Chain, and Storage Mahsa Sedaghat

, Amir Amini

, and Adel Akbarimajd

Climate change and the urgent need for clean energy along with the accelerated improvement of the green economy have added to the importance of hydrogen energy as a strong support for a sustainable society. Thanks to its zero-carbon emission, high energy intensity, high conversion efficiency, and no pollution, hydrogen energy has outperformed traditional fossil fuels. The International Energy Agency (IEA) predicts that it will reduce the CO2 emission to 80 Giga tons by 2050. So far, more than 30 countries and regions have issued green hydrogen (GH2) production strategies. To promote the GH2 market globally and evaluate the lifecycle and emissions of greenhouse gases (GHG), the IEA offers different hydrogen production methods based on official standards. In this chapter, we generally discuss the principles, development rate, significant research points, and technologies and challenges of GH2 production around the world and discuss the general state of global hydrogen energy.

M. Sedaghat · A. Amini (✉) Department of Electrical Engineering, College of Technical and Engineering, West Tehran Branch, Islamic Azad University, Tehran, Iran e-mail: [email protected]; [email protected] A. Akbarimajd Department of Electrical Engineering, College of Technical and Engineering, West Tehran Branch, Islamic Azad University, Tehran, Iran Departmant of Electrical and Computer Engineering, University of Mohaghegh Ardabili, Ardabil, Iran © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 V. Vahidinasab et al. (eds.), Green Hydrogen in Power Systems, Green Energy and Technology, https://doi.org/10.1007/978-3-031-52429-5_10

225

226

10.1

M. Sedaghat et al.

Definition of Green Hydrogen

Various approaches and criteria have been presented to define GH2. [1] called it renewable hydrogen (hydrogen produced by renewable energies such as the sun or wind) while others added more criteria (such as any route that has a purity of 99.9%) and called it GH2. To resolve differences in interpretations and methods of carbon neutralization and reduction of GHG emissions, the definition of hydrogen energy based on the intensity of GHG is widely accepted. Based on CertifHy, hydrogen produced from renewable energies with an emission threshold of 36.4 gCO2/MJH2 at the beginning of production can be defined as GH2 [2]. There is no harmonized definition of GH2 and it is difficult to include it in clean energy policies. Different definitions of GH2 are classified in Table 10.1, which only focus on reducing GHG emissions and zero carbon.

10.2

Characteristics and Initiatives of Green Hydrogen

Despite several technical standards such as “hydrogen fuel quality,” “hydrogen technologies,” and “fuel cells” hydrogen, there is no universal standard for GH2 due to different interpretations. The main characteristics of GH2 are given in Table 10.2, which includes the work done by standardization agencies, certification organizations, and the results of consultations and projects. The only way that will lead to a big leap in the hydrogen energy industry is to rely on the platform of global energy governance and the global market. Therefore, by

Table 10.1 Literature review of GH2 definitions Definitions Renewable energy sources to reduce atmospheric pollution, increasing energy security, and climate problems worldwide. All renewable energy sources.

Renewable energy resources to reduce GHG emissions. Renewable energy sources to achieve zero carbon through carbon capture storage (CCS) and compensating for GHG emissions. All renewable energy and nuclear sources. Renewable/non-renewable energy sources to achieve uncertain emission intensity targets. Low carbon energy sources and low environmental degradation.

References Mauro and Smith [1]

Poullikkas [3], Clark II [4], Clark II and Rifkin [5], Clark et al. [6], Weidong and Zhuoyong [7] Aarnes et al. [8], Viesi et al. [9], Dolci [10] Dawood et al. [11]

Naterer et al. [12] Velazquez Abad and Dodds [13], Dincer [14] Çelik and Yıldız [15]

CertifHy

International

California

In preparation (International Standard) Test Work (Guarantee of Origin scheme)

Active (national standard) Abandoned work (standardization agencies) Active (regulated)

Germany

UK

Method Group working (Guarantee of Origin scheme)

Performer France

Terminology, GO, interfaces, operational management, safety, training, and education. GHG emissions and Renewable energy source

Reduction of air quality and CO2 emissions. Retrieved from CertifHy From natural gas

Gasoline vehicles

From natural gas Not determined

Reduction of CO2 emissions

Reduction of CO2 emissions

GHG None

Main policies Renewable energy sources

Table 10.2 GH2 characterization and initiatives in the world

60% lower 91 gCO2e/MJ H2.

Retrieved from CertifHy

35%–70% below baseline Determining a province based on final consumption 30% lower GHG

Competence level Renewable energy 100%

Any renewable method leads to production with 99.5% purity.

Biomass thermochemical conversion, electrolysis, catalytic cracking Retrieved from CertifHy

Competency processes Any renewable method based on feces to produce hydrogen from waste (including electrolysis power) Renewable methods such as electrolysis Any neutral technology

Retrieved from CertifHy Point of production

Point of use

Point of use Point of production

System boundary Point of production

10 GH2 Networks: Production, Supply Chain, and Storage 227

228

M. Sedaghat et al.

accounting for carbon emissions, mutual recognition and compatibility around the world is necessary to create an international hydrogen standard. GH2 initiatives based on the Table 10.2 refer to four important points: . The definition of GH2 states that renewable energies should be the only source of hydrogen or not. On the other hand, GH2 depends on the amount of GHG emissions and is defined based on the production technology and source of hydrogen. . The quantitative definition of GHG and its life cycle depends on many options as the system boundary, including carbon emission, production point, and consumption point. . The system boundary and national conditions of large countries are the main criteria for carbon emission from hydrogen production or gasoline vehicles. . To obtain the desired green hydrocarbon threshold, the emission of GHG slightly increases due to the reduction of air pollution and the reduction of carbon.

10.3

Hydrogen Supply Chain Network (HSCN)

The purpose of designing a supply chain network is to reduce environmental pollution and minimize CO2 emissions and supply chain costs. This faces several problems such as the capacity and number of manufacturing centers (MCs), distribution centers (DCs), the allocation of factories to MCs and DCs to customers, the flow of materials throughout the system, and the number of products stored in warehouses. The costs are mainly related to purification/separation processes to meet hydrogen purity standards. This issue can be addressed by the following approaches: . . . .

Hydrogen production from steam methane with carbon absorption and storage Installation of new transmission pipelines Changing the use of the gas distribution network “Hydrogen treatment” with a pressure fluctuation absorption system at the refueling place . The transportation industry is responsible for about 17% of CO2 emissions and 18% of primary energy consumption in the world; that is why this industry is under the pressure of decarbonization. One of the best solutions in this regard involves the use of hydrogen as a zero-emission alternate fuel. Hydrogen can be produced from a wide range of sources with minimal loss. It has also the ability to be stored over time [16]. In this regard, the optimization of hydrogen supply chain cost for the transportation sector is presented until 2050. Unlike the heating sector, the use of hydrogen in the transportation sector requires compliance with purity standards that focus on the development of a low-carbon hydrogen supply chain. The optimal method in terms of the costs and technical requirements includes recent technologies of purity and transportation and purification of hydrogen. Therefore, in this section, the main components of

10

GH2 Networks: Production, Supply Chain, and Storage

Production

Transmission

229

Distribution Seperation/ Purification

Refueling

Fig. 10.1 Hydrogen supply chain in the transportation industry

.

.

. .

.

the supply chain and their role in the low-carbon or carbon-free transportation sector are first presented, and then the analysis of the supply chains is discussed. The transfer of hydrogen from MCs to the fueling station requires a hydrogen supply chain which can be divided into five categories (as shown in Fig. 10.1): Production station; due to significant measures to transport hydrogen, large-scale hydrogen production takes place in a centralized facility for economies of scale and higher production efficiency. On the other hand, to reduce the transportation infrastructure, production can be performed near the site of demand based on steam methane reforming (SMR), membrane methane reforming, pyrolysis, gasification, and electrolysis technologies. Transfer; assuming that this sector is like a “motorway,” hydrogen transport refers to the transport of hydrogen from centralized production sites to stations close to demand by truck/trailers, dedicated hydrogen pipeline, and repurposing gas transmission network. Distribution; this section can be also considered like the streets of the city, which is the last stage of the transportation system and delivers small quantities to the end user. Separation/purification; this section should comply with the standards. Depending on the configuration, it can be provided in a centralized form at a large scale to increase the purity after transfer and injection into the distribution system. It can be generally installed in different places of the supply chain depending on the need. Purification of hydrogen requires its pure distribution to the refueling station, which can increase the transportation costs, it can be also achieved locally at the refueling place. Among its main advantages, its inclusion in the gas distribution network reduces the cost of transportation. Purification and separation of hydrogen can be accomplished by technologies of cryogenic distillation, pressure swing adsorption, selective membranes, and electrochemical purification. Refueling; the point of fuel distribution and the end of the chain are refueling stations that are distributed with a pressure of over 700 bar and quickly in less than 3 minutes. There are 4 types of these stations:

230

M. Sedaghat et al.

1. Stations that vaporize and compress liquid hydrogen 2. Stations that receive compressed hydrogen at a pressure of 200 bar from highcompression trailers 3. Stations that receive compressed hydrogen at a pressure of 100 bar from pipelines with a compression unit 4. Stations that produce hydrogen on site through electrolysis and compress it at a pressure of 15–700 bar The degree of hydrogen is selected by purification or non-purification in each part of the supply chain, which can be divided into 4 categories: pure with ~99.97–100%, clean with ~99–99.97%, dirty with ~95–99%, and mixed with ~20% of hydrogen. A hydrogen supply chain for transportation needs in the time horizon of 2020–2050 includes the following decision variables: . . . . . . . .

Capacity Performance of different supply chain technologies Location Year of their installation The total rate of operational decisions is also divided into the following four categories: Production in each region Transportation from one area to another Shipping in any area Hydrogen purification in each area

10.4

Green Hydrogen Production Energy Sources

The use of the GH2 industry in the development of low-carbon energy is one of the most important energy strategies. In this context, avoiding fossil fuel consumption to obtain “GH2 carbon-free” is of great interest. exploiting solar energy, the combination of wind energy with water electrolysis, and renewable energy sources is one of the main goals of this approach. Strategies have been proposed by European and American countries to create a regulatory framework to promote and support the GH2 industry.

10.4.1

Wind Power

One of the approaches for the production of GH2 involves the use of wind power. Wind generators convert wind energy into electrical energy and hydrogen is produced from the electrolysis of water, which by passing through the hydrogen transfer system leads to the production of hydrogen from wind energy [17]. Hydrogen

10

GH2 Networks: Production, Supply Chain, and Storage

231

production in the grid-connected system is divided into 3 categories, which are mainly used in large-scale wind sites. . Grid-connected wind turbines . Receiving electricity from the grid . Hydrogen production and water electrolysis [18] In off-grid systems, the electrical energy generated by the fans directly enters the water electrolysis equipment without passing through the grid to produce hydrogen. This method only meets the demand for local production. Hydrogen was first produced in the United States in 2004 by connecting a producer to an electrolysis reactor. The “Wind to Hydrogen” project was launched at the US National Renewable Energy Laboratory [19], which used photovoltaic and wind energy to generate and store hydrogen. In this plan, the optimal output capacity of the wind hydrogen system and the wind-hydrogen electrolysis technology play the main role. After the United States, the European Union takes its steps toward no-fossil fuel approaches. As a leader in the field of wind energy conversion to hydrogen, it has implemented projects in Spain and Greece to combine hydrogen production technology with wind energy. This project includes hydrogen storage technology and seawater desalination to produce GH2, which simultaneously provides power and freshwater [20].

10.4.2

Solar Energy

The current production of solar hydrogen involves the electrolysis of water with direct and indirect connections between photovoltaic panels [21]. It is an economical and efficient method of indirect connection that does not require photovoltaic modules, MPPT controllers, DC-DC converters, and batteries because it increases the cost of the system and causes power losses and decreases the efficiency of the system. With the advent of photovoltaic electricity technology in the European market, it quickly took a huge share of the market. In 2020 this market reached about 700 gigawatts and it occupies about 70% of the global market. Now China ranks first with its worldwide installation [22]. Fereidooni et al. [23] reported the feasibility of producing hydrogen from photovoltaic systems and setting up a 20 kW power plant in Yazd-Iran. The capacity of their system was about 373 tons per year, but the investment return period is about 17 years. This value should be reduced by three quarters compared to other technologies [24]. In general, there is no significant advantage of a grid-connected photovoltaic system for large-scale hydrogen production.

232

10.5 10.5.1

M. Sedaghat et al.

Green Hydrogen Production Methods Water Electrolysis

Water electrolysis is a promising GH2 production technology which has already matured for industrial applications. In this process, electricity has high potential and is guided by renewable energies and obtained through low-carbon and carbon-free methods [25]. Hydrogen production by electrolysis of water includes two halfreactions, (a) hydrogen evolution reaction (HER) in the cathode and (b) oxygen evolution reaction (OER) in the anode. Water electrolysis technology can be divided into the following 4 types [26]: . Alkaline water electrolysis (AWE): In the early days of industrialization, this technology was reliable and low-cost. Despite its ease of operation, it occupies a large area. The AWE device consists of an electrolyte, electrolysis, anode, cathode, and diaphragm. The mentioned device works at a low operation temperature. Cathode and anode pressure is balanced during operation to prevent explosion caused by hydrogen or oxygen penetration. The conversion rate can be obtained by increasing the current intensity and decreasing the voltage as the electrolysis of alkaline water starts later and the response time becomes longer [27]. . Proton exchange membrane (PEM): This technology had high efficiency and easy operation with high cost (due to the presence of the catalyst). One of its notable features is the coupling with wind and photovoltaic energy due to the small size of the electrolytic cell. The PEM device is a polymeric cationic membrane with an electrolytic cell in the core, where the anode and cathode catalysts are connected to this membrane. The unit conductivity of hydrogen ions, and as a result high safety, is one of the advantages of this technology. The electrolytic cell has a compact structure [28]. . Solid oxide electrolysis (SOE): this technology has certain limitations, including its high operation temperature. The cathode of the device is made of nickel-based porous serum while the anode is made of perovskite oxide. The electrolyte layer is an oxygen ion conductor which requires a high-temperature reaction site. Due to the high-temperature operating conditions, the material standard for indoor devices is high [29]. . Anion exchange membrane (AEM), this novel technology (especially in the field of material science) features an anion exchange membrane, low operating temperature, fast start-up, and non-corrosive reaction environment [30]. These four technologies are shown in Fig. 10.2 below.

10

GH2 Networks: Production, Supply Chain, and Storage

233

Fig. 10.2 The diagram of four technologies of water electrolysis, (a) AWE, (b) PEM, (c) SOE, and (d) AEM. (From Refs. [27–30])

Carrier transport O2/H2O

Valance Band

(3) 1.23 V

(1) Light absorption Conduction Band

Mass transport

(2) Charge separation

H+/H2

Carrier transport

Fig. 10.3 Photocatalytic hydrogen production mechanism [31, 32]

10.5.2

Photocatalysis

In 1972, researchers discovered that hydrogen can be produced by photocatalytic water splitting on TiO2 electrodes. During this process, a catalyst absorbs photons to produce electrons and holes with high energy, i.e., photocatalysis. Photocatalysis leads to the production of hydrogen through electron-hole pairs under light irradiation as the target product, the mechanism of which is shown in Fig. 10.3. This process converts and stores solar energy into chemical energy as a GH2 production technology [31, 32]. Being produced based on light absorption is one of the main bottlenecks in the application of photocatalysis. On the other hand, compounds such as alcohols and biomass are used as photocatalytic raw materials. Photocatalytic materials can be modified by various techniques to adjust their optic, charge separation, and surface properties. For the first time, researchers made photocatalysts with a series of

234

M. Sedaghat et al.

tristyrenes (TP) that are covalently recombined with organic polymers, that the diphenyl benzothiadiazide (DPBT)-TP photocatalyst significantly increased the HER rate proving that the use of a polymeric photocatalyst is sufficient for efficient hydrogen evolution. Besides the challenges associated with photocatalytic GH2 production including weak reaction and low radiation efficiency of the light source, this method enjoys certain advantages. On the other hand, photocatalytic hydrogen production focuses more on the energy band and catalysts, while researchers have made good progress in designing reactors and devices to improve the performance of catalysts, this path is still limited.

10.5.3

Biomass

In this method, pure hydrogen is obtained through gasification, conversion, decarbonization, and separation using biomass (as raw material). Biomass is used in chemical processes to obtain hydrogen compounds with the use of catalysts. This approach technologically uses biomass in hydrogen production to meet the conditions of industrialization. The path of GH2 production by biomass is illustrated in Fig. 10.4. In this method, the GH2 transportation and storage sector accounts for most of the production costs, which is the main obstacle in the widespread use of GH2 in the energy system. Hydrogen production through biomass can overcome this problem and offer a solution to the current development of hydrogen energy [31, 32]. The two main routes of GH2 production through biomass are as follows [33]: . Biological production of hydrogen: Anaerobic fermentation and photosynthetic hydrogen production are performed through hydrogen-producing microorganisms. But their large-scale application is limited due to poor performance and stability. . Thermochemical hydrogen production: It involves the pyrolysis of biomass hydrocarbons with synthesis gas (CO, H2) and the reaction of CO and H2O to produce hydrogen. Studies have revealed that the visible light absorption spectrum of the biomass method is significantly higher than the photocatalysis method. Improving the oxidation capacity of the catalyst can strengthen the adsorption and activation of biomass, thus, improving its efficiency. In general, Biomass-based GH2 production technology is in line with carbon resource standards but needs to be improved for hydrogen production from carbon to achieve carbon-free energy consumption.

10

GH2 Networks: Production, Supply Chain, and Storage

235

Anaerobic fermentation Pretreatment Gasification

Physical process

CO/H2

Chemical process

Energy crops Forestry waste Agricultural waste Starch crop

Methanol

Steam reforming

Steam reforming

Other methods (Dry reforming, Cracking, etc.)

Methanol

Biological process

Biomass feedstock

hydrolysis

Enzymatic

Glucose

H2

Ethanol

Steam reforming

Fructose etc.

Fig. 10.4 The route of GH2 production by biomass

10.6 Green Hydrogen Transportation Gaseous hydrogen is transported in compressed form through high-pressure pipelines, pipe trailers, or pipe railways. Hydrogen transport is divided into two parts: transmission and distribution. Within regions or across one or more countries, transmission occurs over long distances between production sites and distribution stations, and distribution networks use pipelines or trucks to transport hydrogen to the demand location. In hydrogen transportation, dedicated pipelines are used in relatively small networks [34]. For example, currently 2,600 km of pipelines in the United States of America are dedicated to hydrogen transportation, which is very small compared to 300,000 km of natural gas pipelines. Therefore, the idea of replacing these lines for hydrogen transportation or their improvement has been proposed in recent years [35]. But this replacement faces challenges such as: retrofitting, replacing compressors, increasing the number of turbines and compressors, insulating, and preventing leakage due to the age of the lines.

236

10.7

M. Sedaghat et al.

Green Hydrogen Safety

This section presents a brief description of the safety of hydrogen principles in line with the laws and regulations and the requirements and standards of Hydrogen Refueling Station (HRS). Hydrogen is highly flammable with a fast flaming speed (in a wide range of concentrations between 4% and 74% of volume). This rate is, however, lower than gasoline and diesel. Hydrogen is non-toxic with no color or smell [36]. The characteristics of hydrogen are compared with other fuels in Table 10.3. Liquid propane increases the risk of explosion near the ground due to its high density and production of heavy gases. Since the density of hydrogen is about one-fourteenth of air, it disperses quickly in the open air showing a lower risk of explosion near the ground (the explosive range of hydrogen is 11–70% in air). Therefore, the careful use of hydrogen will reduce possible risks. Hydrogen burns under UV irradiation which is invisible to the human eyes. Hydrogen combustion equipment must be thus employed to prevent possible sparks; thus, ignition and explosion are the most important issues in the field of hydrogen safety. The degassing process is a type of burning process which occurs when the initial velocity of either hydrogen-oxygen and hydrogen-air mixtures matches the combustion speed of hydrogen in air or oxygen mixtures. The flame can be defined as a change from a laminar state to a turbulent state which creates a pressure wave and quickly spreads outwards. At elevated temperatures, an explosion when the flame travels faster than the sound. Explosions may be fatal, so safety measures must be implemented to prevent their consequences. One of the main components of the HRS expansion is for fuel cell refrigerators, although the number of conventional fuel stations is relatively more than that, but HRS are also increasing. According to the reports [37] about 320, 375, and 470 stations were established in 2017, 2018, and 2019, respectively. Until 2019, most of these stations existed in Asia (200 stations) while Europe had the second rank with 185 stations. US only had 85 stations in 2019. Storage stations are divided into two types of liquid (LH2) and gaseous (GH2). A hydrogen storage station including supply, intermediate storage, high-pressure storage and distribution sections as shown in Fig. 10.5. The technique used in the supply sector plays an important role. Hydrogen may be supplied on site, or transported to the site by pipe and pipeline trailers, or liquid hydrogen may be brought to the site by truck tanks. Today, the stations have a capacity of 100–520 kg of compressed GH2, and due to the denser nature of LH2, they can produce at least 1000 kg per day. Table 10.3 Characteristics of hydrogen compared with methane [36]

Characteristics (in air) Lower heating value [kWh/kg] Flame temperature [K] Flammable range [vol%] Ignition energy [μJ] Auto-ignition temperature [°C] Density ratio

Hydrogen 33.33 2318.15 4.0–75.0 20 858.15 0.07

Methane 13.89 2148.15 5.3–15.0 290 813.15 0.55

10

GH2 Networks: Production, Supply Chain, and Storage

GH2 tube trailer

H2 Pipeline

237

External tube trailer

On-site Production

Production / Supply 20-50 MPa

1-3 MPa Steam methane reforming Water electrolysis 90-95 MPa 2-5 MPa Storage Compressor GH2 Medium Pressure Storage GH2

Medium Pressure Storage Up to 40-50 MPa

40-50 MPa

High Pressure Storage GH2

Booster Compressor GH2

High Pressure Buffer Tanks GH2

90-95 MPa

Up to 90-95 MPa

90 MPa

Evaporator, HX From LH2 to GH2

High Pressure Storage

H2 Pre-cooling,GH2

H2 dispensing,GH2

-40°C Station Type A

Up to 3.6kg/min 70 MPa

Cooling & Dispensing

Fig. 10.5 Hydrogen storage installed in an HRS

10.8

Green Hydrogen Technologies in Region

According to the construction industry, the combined systems of hydrogen electricity and heat and the use of GH2 as a means of storing renewable energy could be a cost-effective energy source with greater efficiency. Today, more than 30 strategic regions were found for hydrogen production around the world, which are different in scope and scale. In this section, we examine hydrogen energy around the world according to the following five patterns.

238

10.8.1

M. Sedaghat et al.

Patterns

. Strategic storage technology: This model is aimed at engaging new hydrogen energy technologies to achieve energy independence and security in the United States and Canada. For this reason, American presidents prioritized hydrogen technology as the basis of the entire energy technology chain. In 2021, the United States announced trade incentives aimed at decarbonization in line with investment and employment in infrastructure. . Implementation of decarbonization methods; decarbonization of industry, transport, and other sectors of hydrogen energy is necessary in Europe as a pioneer of energy transfer. The aim of carbon neutrality for the European Union has been set for 2050. The “National Hydrogen Energy Strategy” was issued in Germany in 2020 for the rapid development of the market and to create a value market for GH2. . Export of hydrogen energy: Due to the sanctions against Russia in 2022, Australia and Russia are major exporters of energy in the global hydrogen trade. Australia aims to become one of the top exporters of the Asian market by 2025 to achieve the billion-dollar hydrogen market by 2050. . Development and security of industrial systems: Unstable energy prices in the global market for hydrogen and the limitations of natural resources made Korea and Japan seek suitable alternatives to oil and natural gas. Therefore, in 2019, Japan began to plan a strategy as the first country seeking to have a hydrogenbased society. . Rising giant: As hydrogen is a secondary industrial energy source in the world with various scenarios for its applications, China considers hydrogen to be a suitable competitive option. In line with this issue, China developed emerging industrial systems for deep decarbonization, in 2022.

10.8.2

Other Policies Related to GH2 in Different Continents

. Asia In 2019, a strategy and roadmap for hydrogen was published by the Korean government [38]. Based on [39], the national industry has proposed GH2 as a product obtained from renewable energies (through electrolysis) provided until 2020. The renewable portfolio standard (RPS) system obliges producers to install renewable power plants to provide a share of energy through new energies. [40] predicted that by 2030, decarbonization will be carried out in parts of the Japanese industry, including electricity generation and hydrogen. Therefore, to realize the hydrogen economy and as long as it is cost-effective at the same speed as the country wants, GH2 production is considered vital. Although Japan is developing in this issue, it follows global developments and helps the international agency to achieve the economic and efficiency goal.

10

GH2 Networks: Production, Supply Chain, and Storage

239

. Australia Australia seeks to develop a national strategy for hydrogen export in line with regulatory changes, development of standards, realization of business opportunities, and use of 10% hydrogen in the domestic gas network to 100% [41]. To establish standards for the production, distribution, and use of GH2, the Australian Renewable Energy Agency (ARENA) has funded its first green innovation center to recognize GH2 produced through the electrolysis of renewable electricity. . European Union [42] stated that EU RED2 does not necessarily need to comply with standards and regulates fuel quality and biofuel sustainability guidelines. According to the European Union policies, the consumption of gasoline, diesel, and biofuels, should be controlled until 2020, especially in the transportation industry, to achieve a 6% reduction in GHG intensity, because GH2 is considered a renewable fuel of non-biological origin. This goal helps to reduce carbon intensity in processes such as diesel desulfurization. However, the decarbonization of transportation fuels in the framework of RED2 is one of the main goals of the European Union Commission after 2020 to reduce the intensity of GHG by 14% by 2030. Hydrogen has been identified as one of the alternatives to oil in the European Union directives. Regarding the infrastructure of alternative fuel (hydrogen), it can be considered to promote the construction of refueling networks, which is among the voluntary goals of the European Union until 2025. Guidelines including standards and technical specifications of refueling points and GH2 distribution have been provided. In line with this goal, the European Commission has prepared gas packages as a new opportunity under the title of decarbonized and renewable gases to be offered in gas networks. . United States A bill has been presented in the California Air Resources Board (CARB) regarding low-carbon fuel, which guarantees that hydrogen fueling stations have less carbon intensity than gasoline and that one-third of hydrogen is a renewable fuel. The requirement to reduce the intensity of GHG has led to recent reforms in the market-based mechanisms of their commercial systems to use renewable fuel in the transportation industry.

10.9

Standards of Green Hydrogen Refueling Places

For setting up and installing HRSs, short safety and planning programs have been introduced in accordance with the standards. The standards include issues such as the safe and fast entry of HRS technology into the market. Below are the codes of regulations and standards of HRS: . A document that includes the specifications of a particular component/approach/ method and explains it is called “standard”

240

M. Sedaghat et al.

. A document that describes the desired results (for example, a product) is called “code” . Mandatory and mandatory items, contrary to standards and codes, which are called “regulations” The standards determine and guarantee the compatibility and safety of different systems in HRSs. These standards are continuously reviewed and updated to ensure compatibility between vehicle manufacturers, operators, and energy suppliers according to technological progress and safety requirements. Hydrogen standards are determined by the technical committees of two major international standard organizations, ISO (International Standards Organization), and IEC (International Electro-Technical Commission), in which the national standard bodies are also composed of the Society of Automotive Engineers (SAE), Technical Committee (TC), and the European Committee for Standardization (CEN). Hydrogen standards are: . ISO/TC 197; in hydrogen technologies . ISO/TC 220; in cryogenic vessels . ISO/TC 58; in gas cylinders . ISO/TC 22/SC 41; in gaseous fuels-specific

10.10 10.10.1

Challenges and Outlook Greenhouse Gases (GHG)

GHG emission is possible at all stages of the hydrogen supply chain. Therefore, the calculation of loss and GHG emission during hydrogen distribution and storage is challenging due to the existence of several potential distribution routes. The intensity of GHG emissions depends on the location of supply and demand; while the final impact and efficiency of technology add to its complexity.

10.10.2

Continuous and Economical GOs Scheme

Operations such as data collection, auditing, transactions, and investment in equipment in the hydrogen supply and delivery chain involve costs. GOs marketing includes several stakeholders whose presence has added value and increases costs. Costs such as registration in the GOs scheme and the cost of transactions in the maintenance chain can be also added. Of course, transaction costs are low such that they do not have much impact on investment decisions. One of the challenges for small producers in the initial investment is the same costs and obstacles to enter the market. Ideally GOs scheme will keep small contracts until the maximum demand increases and the volume of business increases.

10

GH2 Networks: Production, Supply Chain, and Storage

10.10.3

241

Emission Intensity Threshold

The low-carbon hydrogen or GH2 can be described based on emission intensity thresholds. Therefore, there are different approaches that sometimes determine the province in relation to the emission of fossil fuel gases, and sometimes it is an absolute province. For example, renewable energy directive (RED) considers the minimum requirement for renewable hydrogen to be a 60% emission reduction compared to fossil fuels. On the other hand, the absolute limit may decrease continuously over time. If the province becomes stricter, instead of pursuing clean technologies, producers will not pursue dirty technologies to produce a green province and instead will be encouraged to adapt their supply chain to high-quality, yet more profitable, hydrogen.

10.11

Conclusion

The inherent characteristics of hydrogen introduced it as one of the most basic carbon-free energies in the future. It, however, faces major challenges, especially the lack of infrastructure, deployment, and large-scale use. Moreover, scientific communities believe that hydrogen can be something more than a fuel in the transportation industry with broader applications. Therefore, new technologies of hydrogen supply chain, including hydrogen production from different sources, storage, and distribution are being investigated. Gaps in these technologies include larger boil-off, the lack of development of insulators, lack of smaller design, and data limitations which must be resolved for commercial application of hydrogen supply chain on a global and larger scale.

References 1. Mauro, R. L., & Smith, D. L. (1994). Development of industrial interest in hydrogen (p. 29). 2. CertifHy, C.-S. (2019). Hydrogen criteria, CertifHy scheme subsidiary document. CertifHy Scheme Subsidiary Document. 3. Poullikkas, A. (2007). Implementation of distributed generation technologies in isolated power systems. Renewable and Sustainable Energy Reviews, 11(1), 30–56. 4. Clark, W. W., II. (2008). The green hydrogen paradigm shift: Energy generation for stations to vehicles. Utilities Policy, 16(2), 117–129. 5. Clark, W. W., II, & Rifkin, J. (2006). A green hydrogen economy. Energy Policy, 34(17), 2630–2639. 6. Clark, W. W., Rifkin, J., O'Connor, T., Swisher, J., Lipman, T., Rambach, G., Science, C. H., & Team, T. (2005). Hydrogen energy stations: Along the roadside to the hydrogen economy. Utilities Policy, 13(1), 41–50. 7. Weidong, G., & Zhuoyong, Y. (2012). Research on non-grid-connected wind power/waterelectrolytic hydrogen production system. International Journal of Hydrogen energy, 37(1), 737–740.

242

M. Sedaghat et al.

8. Aarnes, J., Eijgelaar, M., & Hektor, E. A. (2018). Hydrogen as an energy carrier: An evaluation of emerging hydrogen value chains: Group technology & research-position paper 9. Viesi, D., Crema, L., & Testi, M. (2017). The Italian hydrogen mobility scenario implementing the European directive on alternative fuels infrastructure (DAFI 2014/94/EU). International Journal of Hydrogen Energy, 42(44), 27354–27373. 10. Dolci, F. (2018). Green hydrogen opportunities in selected industrial processes. European Union Science Hub. 11. Dawood, F., Shafiullah, G., & Anda, M. (2020). A hover view over Australia’s hydrogen industry in recent history: The necessity for a hydrogen industry knowledge-sharing platform. International Journal of Hydrogen Energy, 45(58), 32916–32939. 12. Naterer, G., Gabriel, K., Wang, Z., Daggupati, V., & Gravelsins, R. (2008). Thermochemical hydrogen production with a copper–chlorine cycle. I: oxygen release from copper oxychloride decomposition. International Journal of Hydrogen Energy, 33(20), 5439–5450. 13. Velazquez Abad, A., Dodds, P. (2016). Green hydrogen standards. 14. Dincer, I. (2012). Green methods for hydrogen production. International Journal of Hydrogen Energy, 37(2), 1954–1971. 15. Çelik, D., & Yıldız, M. (2017). Investigation of hydrogen production methods in accordance with green chemistry principles. International Journal of Hydrogen Energy, 42(36), 23395–23401. 16. Wickham, D., Hawkes, A., & Jalil-Vega, F. (2022). Hydrogen supply chain optimisation for the transport sector–Focus on hydrogen purity and purification requirements. Applied Energy, 305, 117740. 17. Mostafaeipour, A., Dehshiri, S. J. H., Dehshiri, S. S. H., & Jahangiri, M. (2020). Prioritization of potential locations for harnessing wind energy to produce hydrogen in Afghanistan. International Journal of Hydrogen Energy, 45(58), 33169–33184. 18. Li, Y., Wu, F., & Miao, H. (2019). Analysis of research status and development trend of hydrogen storage technology. In Proceedings of the 2019 3rd international conference on information system and data mining. 19. Li, Z., Zhang, W., Zhang, R., & Sun, H. (2020). Development of renewable energy multienergy complementary hydrogen energy system (A Case Study in China): A review. Energy Exploration & Exploitation, 38(6), 2099–2127. 20. Apostolou, D., & Enevoldsen, P. (2019). The past, present and potential of hydrogen as a multifunctional storage application for wind power. Renewable and Sustainable Energy Reviews, 112, 917–929. 21. Chang-qing, G., Li-qi, Y., Chang-feng, Y., Yan, S., & Zhi-da, W. (2019). Optimization of photovoltaic-PEM electrolyzer direct coupling systems. Advances in New & Renewable Energy, 7(3). 22. Chai, S., Zhang, G., Li, G., & Zhang, Y. (2021). Industrial hydrogen production technology and development status in China: A review. Clean Technologies and Environmental Policy, 23(7), 1931–1946. 23. Fereidooni, M., Mostafaeipour, A., Kalantar, V., & Goudarzi, H. (2018). A comprehensive evaluation of hydrogen production from photovoltaic power station. Renewable and Sustainable Energy Reviews, 82, 415–423. 24. Marino, C., Nucara, A., Panzera, M., Pietrafesa, M., & Varano, V. (2019). Energetic and economic analysis of a stand alone photovoltaic system with hydrogen storage. Renewable Energy, 142, 316–329. 25. Xu, Y., & Zhang, B. (2019). Recent advances in electrochemical hydrogen production from water assisted by alternative oxidation reactions. ChemElectroChem, 6(13), 3214–3226. 26. Guo, B., Luo, D., & Zhou, H. (2021). Recent advances in renewable energy electrolysis hydrogen production technology and related electrocatalysts. Chemical Industry and Engineering Progress, 40(6), 2933. 27. Zeng, K., & Zhang, D. (2010). Recent progress in alkaline water electrolysis for hydrogen production and applications. Progress in Energy and Combustion Science, 36(3), 307–326.

10

GH2 Networks: Production, Supply Chain, and Storage

243

28. Wang, T., Cao, X., & Jiao, L. (2022). PEM water electrolysis for hydrogen production: fundamentals, advances, and prospects. Carbon Neutrality, 1(1), 21. 29. Amaya-Dueñas, D. M., Riegraf, M., Nenning, A., Opitz, A. K., Costa, R., & Friedrich, K. A. (2022). Operational aspects of a perovskite chromite-based fuel electrode in solid oxide electrolysis cells (SOEC). ACS Applied Energy Materials, 5(7), 8143–8156. 30. Kandil, U. F. M., Taha, E. O., Mahmoud, E. A., Mahmoud, M., & Taha, M. R. (2022). Quaternary functionalized cellulose-based biopolymer for anion exchange membrane fabrication. Egyptian Journal of Chemistry, 65(10), 47–56. 31. Cao, L., Iris, K., Xiong, X., Tsang, D. C., Zhang, S., Clark, J. H., Hu, C., Ng, Y. H., Shang, J., & Ok, Y. S. (2020a). Biorenewable hydrogen production through biomass gasification: A review and future prospects. Environmental Research, 186, 109547. 32. Cao, S., Piao, L., & Chen, X. (2020b). Emerging photocatalysts for hydrogen evolution. Trends in Chemistry, 2(1), 57–70. 33. Huang, C.-W., Nguyen, B.-S., Wu, J. C.-S., & Nguyen, V.-H. (2020). A current perspective for photocatalysis towards the hydrogen production from biomass-derived organic substances and water. International Journal of Hydrogen Energy, 45(36), 18144–18159. 34. Razi, F., & Dincer, I. (2022). Renewable energy development and hydrogen economy in MENA region: A review. Renewable and Sustainable Energy Reviews, 168, 112763. 35. Erdener, B. C., Sergi, B., Guerra, O. J., Chueca, A. L., Pambour, K., Brancucci, C., & Hodge, B.-M. (2023). A review of technical and regulatory limits for hydrogen blending in natural gas pipelines. International Journal of Hydrogen Energy, 48(14), 5595–5617. 36. Tretsiakova-McNally, S., & Makarov, D. (2016). LECTURE-safety of hydrogen storage. [línea]. Available: http://www.hyresponse.eu/files/Lectures/Safety_of_hydrogen_storage_note s.pdf. Último acceso: 04 Abril 2022. 37. Samsun, R., Antoni, L., & Rex, M. (2020). Mobile fuel cell application: Tracking market trends. IEA technology collaboration programme advanced cell. 38. Moon, J.-I. (2019). Remarks by President Moon Jae-In at presentation for hydrogen economy roadmap and Ulsan’s future energy strategy. Office of the President: Seoul. Korea. 39. Kan, S. (2018). South Korea’s hydrogen strategy and industrial perspectives. 40. Abad, A. V., & Dodds, P. E. (2020). Green hydrogen characterisation initiatives: Definitions, standards, guarantees of origin, and challenges. Energy Policy, 138, 111300. 41. Group, C. E. C. H. W (2019). Australia’s national hydrogen strategy. 42. Plan, R. (2018). Communication from the commission to the european parliament, the european council, the council, the european economic and social committee and the committee of the regions. European Commission: Brussels, Belgium.

Chapter 11

Supply Chains of Green Hydrogen Based on Liquid Organic Carriers Inside China: Economic Assessment and Greenhouse Gases Footprint João Godinho , João Graça Gomes , Juan Jiang, Ana Sousa , Ana Gomes, Bruno Henrique Santos , Henrique A. Matos , José Granjo , Pedro Frade, Shuyang Wang, Xu Zhang, Xinyi Li, and Yu Lin

11.1

Introduction

Owing to the upsurge in greenhouse gas (GHG) emissions, especially carbon dioxide (CO2), in the past two centuries, a significant change in world climate has occurred [1]. The threat of climate change has evolved into one of the most J. Godinho · H. A. Matos · J. Granjo CERENA, DEQ, Instituto Superior Técnico, University of Lisbon, Lisbon, Portugal J. G. Gomes (✉) Sino-Portuguese Centre for New Energy Technologies (Shanghai) Co., Ltd., Shanghai, China Shanghai Investigation Design and Research Institute Co., Ltd., Shanghai, China Future Energy Leaders Portugal/Portuguese Association of Energy, Amadora, Portugal e-mail: [email protected] J. Jiang · S. Wang · X. Zhang · Y. Lin Sino-Portuguese Centre for New Energy Technologies (Shanghai) Co., Ltd., Shanghai, China Shanghai Investigation Design and Research Institute Co., Ltd., Shanghai, China A. Sousa CERENA, DEQ, Instituto Superior Técnico, University of Lisbon, Lisbon, Portugal Future Energy Leaders Portugal/Portuguese Association of Energy, Amadora, Portugal A. Gomes · P. Frade Future Energy Leaders Portugal/Portuguese Association of Energy, Amadora, Portugal B. H. Santos Future Energy Leaders Portugal/Portuguese Association of Energy, Amadora, Portugal Faculty of Engineering, University of Porto, Porto, Portugal X. Li Shanghai Investigation Design and Research Institute Co., Ltd., Shanghai, China © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 V. Vahidinasab et al. (eds.), Green Hydrogen in Power Systems, Green Energy and Technology, https://doi.org/10.1007/978-3-031-52429-5_11

245

J. Godinho et al.

246

devastating crises for the global environment, development, and geopolitical stability, posing pressure on humanity’s social and economic activities [2]. To minimize the effects of climate change, major economies have focused on modifying their energy systems and, consequently, prompted the pursuit of low-carbon energy resources, increased the use of carbon-storage technologies, and implemented energy efficiency measures and several fiscal mechanisms that penalize carbon emissions. Moreover, a global treaty was negotiated in 2015 to ensure that the global mean temperature increase is less than 2 °C to the average temperature of the Pre-industrial Era. The pact, branded the Paris Agreement, was signed by 195 countries under the United Nations Framework Convention on Climate Change and established actions to decrease carbon emissions starting in 2020 [3]. Due to the Paris Agreement, several economic blocks, such as the People’s Republic of China, the European Union, and the United States of America, started to outline strategies to reduce their carbon footprint. China, the world’s second-largest economy, has pledged to peak carbon emissions before 2030 and aims to achieve carbon neutrality by 2060 [4]. This is a formidable challenge: if China achieves these targets, it would be the fastest decline from peak emissions amongst major economies, just 30 years compared to a target of 70 years for the European Union (Carbon neutral target for 2050) [5] and 40 years for the United States (Carbon neutral target for 2050) [6]. By the end of 2019, fossil fuels accounted for 90% of China’s primary energy supply, being responsible for annual emissions of 9.9 gigatons (Gt) of CO2-eq and 28% of global CO2-eq emissions. China’s emissions are divided as follows: electricity generation 53,0%, industry 28,1%, transport 9,1%, and others 9.7%, as illustrated in Fig. 11.1 [7].

1.2% 3.4%

1.0%

0.6%

3.5% 9.1%

53.0%

Electricity and heat producers Industry Transport Other energy industries Residential Commercial and public services Agriculture Other non specified

28.1%

Fig. 11.1 CO2-eq emissions by sector in 2019, People’s Republic of China

11

Supply Chains of Green Hydrogen Based on Liquid Organic Carriers. . .

247

Fig. 11.2 Total energy supply by source, the People’s Republic of China, (a) 2019 and (b) prediction for 2025

To achieve the carbon neutrality objectives by 2060, China founded the world’s largest carbon exchange market in July 2021 [8], creating a national framework to define emissions pricing. Additionally, the country has been implementing FiveYear Plans (FYP) with binding targets for reducing carbon emissions and energy intensity. The most recent, the 14th FYP, issued in March 2021, aims to reduce carbon intensity (the amount of CO2-eq emitted per unit of GDP – Gross Domestic Product) by 18% and energy intensity (the amount of energy consumed per unit of GDP) by 13.5% by 2025 [9]. In addition, the plan sets a goal for 2021–2025 to increase the ratio of low-carbon sources in China’s energy mix to 20% and pledges to build 1200 GW of solar photovoltaic (PV) and wind capacity in the same period. Figure 11.2 shows China’s 2019 energy mix, emphasizing the high dependence on fossil fuels, particularly coal [7]. Figure 11.2 also contains a prediction of China’s energy mix for 2025, worth noting the increase of renewable generation and the reduction of coal supply [10]. In the 14th FYP, hydrogen (H2) emerged as one of the six industries highlighted to help decarbonize transport, heating, and heavy industry and to offer long-duration energy storage. With this aim, many subnational governments and central government authorities drew policies to develop the H2 economy and increased the research and development spending on technology [11]. In 2022, the expanding number of hydrogen projects in China spanned the country’s geography, and 23 of China’s provinces and municipalities established H2 expansion plans and included it as a critical priority for sustainable development [12]. A case worth mentioning was the selection of five economic centres Beijing-Tianjin-Hebei cluster (known as the JingJin-Ji), Guangdong and Henan provinces, and Shanghai Municipality as demonstration areas for H2 development. Depending on the development rate of the H2 industry, these regions may be granted up to 1.7 billion RMB (¥) as a tax subsidy [13]. Moreover, the municipality of Beijing and the province of Jiangsu have accelerated the planning and construction of H2 refuelling stations, whereas the

248

J. Godinho et al.

Zhejiang province has established the use of H2 in cogeneration power plants, implemented fuel cell vehicles in public and port logistics transport, and combined hydrogen production with offshore wind capacity [12]. In the Inner Mongolia Autonomous Region, the provincial energy-planning ministry aims to develop seven wind and solar PV power projects in the cities of Baotou and Ordos that could produce approximately 67,000 tons of low-carbon H2 (H2 produced via renewable or nuclear power) a year and targets a total capacity of 100,000 tons a year of green H2 for the province by 2023. The Inner Mongolia strategy also plans to include 60 H2 refuelling stations and more than 3800 fuel cell vehicles operating in the public transport, mining, and logistics sectors [12]. Most of the projects mentioned above are being led by State-Owned Enterprises (SOE). According to China’s State-Owned Assets Supervision and Administration Commission, more than one-third of SOEs are planning for H2 generation, distribution, storage, and application [14]. Moreover, the country is rapidly emerging as a significant home for installed electrolysis capacity and home for 35% of the worldwide manufacturing capacity of electrolysis equipment. More recently, with the 14th Five-Year “New Energy Storage” Execution Plan, issued in March 2022, new targets for the whole H2 industry were set for 2025 [15]. The plan highlights the expansion of H2 vehicle fuel cell filling stations and the nation’s green H2 production capacity between 100,000 and 200,000 tons per year, a conservative value compared to Inner Mongolia’s targets. Additionally, the plan restricts the development of fossil-fuel-based H2 production (grey), aims to set up an H2 gas market near production facilities, and emphasizes the importance of developing more advanced technologies, including nuclear H2 production (pink hydrogen), seawater electrolysis, photocatalysis, solid oxide electrolyser cell H2 production, and thermocatalytic decomposition. Nevertheless, despite the targets and projects mentioned earlier and China being the largest producer of H2, at about 33.4 million tons (Mt) (nearly 25% of the world’s total), most of the volume is generated from fossil fuels. The demand for H2 is also increasing, and several entities forecast the nation’s H2 market to reach 37 Mt in 2030 and 97 Mt in 2050 [16]. In summary, to address the dichotomy between the current type of H2 production and the carbon neutrality goals for the country, as well as to ensure supplying the growing need for H2, it is vital for green H2, produced via renewables, to scale up in the next few years. Due to the mismatch in China within the demand from more industrialized coastal areas in the east and the supply from more renewable resourcerich provinces in the west and northwest, H2, as a flexible energy carrier, may be key to ensuring the fulfilment of the carbon neutrality targets.

11.1.1

Status Quo, Challenges, and Outlook

Numerous authors previously analyzed the progress of the H2 market in China [17] and examined the feasibility of attaining high volumes of green and pink H2 in the national energy system. There are, however, numerous limitations in the present

11

Supply Chains of Green Hydrogen Based on Liquid Organic Carriers. . .

249

literature. First, most studies omit the growth of Chinese internal demand for H2 and electricity, assuming a large share of the national green H2 production would be exported. A paradigmatic case is the work of Song et al. [18]. In [18], the researchers propose and develop a methodology to address the potential to produce costcompetitive H2 using the power generated from wind offshore deployed on the eastern Chinese coast. The H2 would then be transported to Japan to help meet the country’s decarbonization targets for 2030 and 2050. However, the study disregards crucial characteristics of the Chinese energy system and its targets for 2060. It does not clarify the share of wind offshore power that would be used to supply the load demand. Second, most studies are focused only on the H2 demand and supply analysis, not considering the different transport costs of green H2, a crucial assessment when considering the long distances and geographical barriers between the richest regions in terms of renewable resource availability and China’s leading economic centres. In Zhang S. et al. [19], a long-term analysis of China’s decarbonization plan is performed. The study highlights that electrolytic hydrogen production from renewable energy sources will become mainstream and essential to decarbonising the transport sector. Nevertheless, the study does not analyze in detail the different costs associated with the transport of green H2 (road, pipeline, and railway) for the medium (~1000 km) and long-range (~5000 km), nor does it state the Levelized Cost of Electricity (LCOE) when producing green H2 from different sources, per province, according to technology, which may impair the conclusions. Third, some reports do not scrutinize the impact of power curtailment and energy storage on the Levelized Cost of Hydrogen (LCOH). In [20], the researchers analyze the production cost, cost structure, and regional differences of several hydrogen production methods in China via the LCOH model [20]. The study concludes that coal-to-hydrogen with CO2-eq capture and storage can be a cost-effective option, cheaper than renewable energy-based water electrolysis, especially in remote provinces such as Xinjiang, Gansu, and Inner Mongolia. There is, though, the severe limitation of not analyzing the levels of renewable power curtailment in the abovementioned provinces. The use of the curtailed power in water electrolysis will correspond to an optimization of the already installed capacity of renewables and, therefore, a reduction in their LCOE. Finally, H2 can only be used for storage or transportation when compressed, liquefied, or attached to a carrier. Despite the investment and policies enacted to support the development of H2 in China, it was not identified a specific public outline or analysis, neither at the national nor at the provincial level, for the transport of green H2 from the wealthiest resource availability areas to the main consumption centres.

11.1.2

Problem Definition and Research Aim

This study aims to answer some of the most challenging questions of the Chinese carbon neutrality plan, considering the governmental targets for 2060, the outline of the FYPs, and the techno-geographical constraints of implementing a viable green

250

J. Godinho et al.

H2 production and transport strategy. In this study and following recent literature, green H2 is assumed to be a crucial step in achieving the Chinese decarbonization goals for several industries, such as the steel and chemical industries, process heating, building environment, and transport. This work presents a methodology to characterize the economic and environmental impacts of producing, storing, and long-distance green H2 transportation. The H2 to be transported is stored via two-way Liquid Organic Hydrogen Carriers (LOHC), in particular, the toluenemethylcyclohexane (TOL-MCH) and the dibenzyltoluene-perhydro-dibenzyltoluene (DBT-PDBT) systems. The methodology assumes that the H2 is produced in China’s remote provinces with significant renewable resources, particularly Inner Mongolia, Tibet, Gansu, and Xinjiang, and transported to the country’s main energy consumption centres, in specific the Jing-Jin-Ji cluster, the Pearl River Delta (Guangdong, Hong Kong and Macau) and the Yangtze River Delta (Shanghai, Jiangsu, Zhejiang and Anhui). Several case studies were developed to assess the costs and CO2-eq emissions reduction associated with different H2 transport and production schemes for 2030, 2040, 2050 and 2060. The main novelty of this paper compared to previous research are: (a) Develop a multi-objective methodology, based on LCOE and resource availability, to assess the renewable power potential to generate green H2. (b) Investigate the optimal techno-economic method for the storage and longdistance transportation of green H2. (c) Study the techno-economic viability of developing a green H2 supply chain capable of reducing China’s carbon emissions, particularly in its main economic clusters. (d) Project China’s current and future potentials for domestic H2 supply and demand for the 2060 horizon. Notwithstanding the study’s focus on China, the methodology is valid and relevant to other geographic regions. The following chapter sections are organized as stated: Sect. 11.2 describes the state of the art of hydrogen storage technology and transport; Sect. 11.3 introduces the case study; the methodology is defined in Sect. 11.4; the results are shown and discussed in Sect. 11.5 and the main conclusions are presented in Sect. 11.6.

11.2

Hydrogen Storage and Transport: State of the Art

The increase in power generated from variable renewable sources will necessitate greater flexibility and adaptability within the electricity system [21–23]. Several flexibility strategies include using electrical batteries, hydro-pumped storage, compressed air storage, or surplus renewable electricity generation to produce hydrogen [24]. The most mature technology is hydro pump storage, which has, however, geographical, topographical and environmental restrictions [23]. With the increase in electrification, batteries are expected to be further developed, being anticipated not

11

Supply Chains of Green Hydrogen Based on Liquid Organic Carriers. . .

251

Fig. 11.3 Hydrogen storage methods

only that storage costs will diminish but also that their durability will increase. Nonetheless, batteries are not expected to be economically viable for long-life scales [25]. Furthermore, electrical battery supply chains are responsible for various environmental issues, mainly related to mining activities, such as toxic gas emissions and hazardous waste [26]. In this sense, it is necessary to develop technologies that are more efficient in long-term storage, are easily transportable, and do not have spatial restrictions. Using H2 as a carrier to distribute and store energy is a competitive and adequate solution, solving some of the above-mentioned issues [27]. The volumetric energy density of liquid hydrogen or compressed gas hydrogen is significantly lower than that of fossil fuels. However, the mass-energy density of the H2 is higher [28]. For these reasons, H2 can only be used for storage or transport if it is compressed, liquefied, or added, by hydrogenation, to a carrier. Liquefied hydrogen storage requires very low temperatures, as the boiling point of hydrogen at atmospheric pressure is -252.8 °C. If compressed H2 is stored, storage tanks must resist high pressures, between 350 and 700 bar. H2 can also be held on the surface of solids, by adsorption, or within solids, by absorption. Figure 11.3 represents how hydrogen can be stored [29]. Using adsorbent materials such as metal-organic frameworks (MOFs) is an option, mainly due to high surface area, tunable pore size, adsorption efficiency, and lightweight. Inside this group, MOF-5 stands out as a promising adsorbent for H2 storage due to its porosity [30, 31]. LOHCs are liquid organic compounds that allow safe, long-term, and reversible storage of H2. They use hydrogenation and catalytic dehydrogenation cycles with heat exchange [22, 32]. LOHCs are more attractive options for hydrogen storage when compared to liquid or compressed hydrogen, as they may use existing chemical transport infrastructure [33]. Some of the examples widely used as LOHCs are toluene-methylcyclohexane (TOL-MCH), dibenzyltoluene-perhydrodibenzyltoluene (DBT-PDBT), BN-methyl cyclopentane, and N-ethylcarbazoldodecahydro-N-ethylcarbazole (NEC-DNEC) [22, 29]. Hydrogen is a pivotal feedstock in producing essential chemicals, including ammonia, formic acid, and methanol. Remarkably, these chemicals can release hydrogen through controlled decomposition processes. Notably, ammonia production involves a very energy-intensive procedure, and it can be conveniently stored in liquid form under mild conditions. The energy density per volume unit of H2 in ammonia and methanol is higher when compared to liquefied or gaseous H2 [23, 25, 34], but they are not categorized as LOHCs [29, 33, 35]. In these physical and material-based methods, the H2 is converted into a suitable form of transport. After transportation, it is reconverted back to pure hydrogen. The hydrogen value chain

252

J. Godinho et al.

steps are in different maturity stages, according to the technology readiness level (TRL). Compressed gaseous and liquified hydrogen for small-scale projects have a very high TRL and are, therefore, ready for commercialization. For large-scale projects, the transport of gaseous hydrogen is mature [36]. The transportation of liquified H2, applied on large scales, is still in the demonstration phase. Although ammonia transport is at a more advanced stage of market development than LOCHs, with a higher TRL, LOCHs would be suitable and less costly [37]. Figure 11.4 systematizes the integration of the renewable energy system and the H2 storage and transport process [29]. When there is a surplus of renewable electricity, the excess can be used as input to an electrolyser to produce green H2. Subsequently, the H2 is absorbed by the LOHCs through hydrogenation and transported to the desired location. LOHCs allow hydrogen transport at normal temperature and standard pressure in 20- or 40-foot containers (32.6 m3 and 69 m3) [38]. At the destination, LOHCs are dehydrogenated to release H2. LOHCs have the advantage of being compatible with existing fuel transport infrastructure. Furthermore, it is not expected that there will be any hydrogen losses in storage or transport [39]. There is, however, no single ideal way of transporting hydrogen; the most economical strategy depends on distance, quantity, end use, and available infrastructure [37]. Transporting compressed hydrogen in pipelines for short distances costs less than transporting liquefied H2 and chemical carriers such as LOHCs. Repurposing existing pipelines to use hydrogen is expected to significantly reduce total costs, turning the pipeline option the most competitive in the future. On the other hand, chemical carriers become more competitive the longer the delivery distance due to the lower transportation cost. The following subsections will characterize the different ways to transport H2 between the Chinese provinces.

11.2.1

Road Transportation

Green H2 is a fundamental vector in the energy transition process with an increased impact in the transport sector. Nevertheless, it is precisely in the road transport process that several challenges arise when a technical-economic analysis is provided. The thermodynamic characteristics of H2 make the transport and storage inefficient under standard conditions, which constitutes a significant challenge for advancing the H2 economy. According to European and American legislation, the maximum Gross Weight of assembly (tractor vehicle/semi-trailer or rigid/trailer) with five or more axles, carrying two 20′ ISO containers or one 40′ ISO container, is 44 tons. There are three transport methods by road: compressed gas, cryogenic hydrogen liquid or using hydrogen carriers (Fig. 11.5). As for road transportation, H2 can be carried as compressed gas via tube, with approximately 2,000 L of capacity and with a pressure of between 180 and 250 bars; or in a liquefied state, using tank-containers with a capacity ranging from 20,000 to 50,000 L, with pressure levels between 6 and 10 bar and at -253 °C [40]. Using LOHCs will allow transport at normal

Supply Chains of Green Hydrogen Based on Liquid Organic Carriers. . .

253

Fig. 11.4 LOHC supply chain system

11

254

J. Godinho et al.

Fig. 11.5 Container capacities for different road transport hydrogen approaches

temperature and pressure in standard containers of 10 ft (3.00 m × 2.40 m × 2.60 m), 20 ft (6.00 m × 2.42 m × 2.60 m) or 40 ft (12.00 m × 2.40 m × 2.90 m) [40]. Compressed H2 can be used for short-distance and small-scale projects. Cryogenic liquified hydrogen is suitable for long-distance and high-volume demand. LOHCs can be used for long-term storage and long-distance paths. While the boiloff losses are negligible for LOCH storage, liquified hydrogen losses need to be considered and are around 0.03%/day [32]. The research developed by Hurskainen & Ihonen [39] showed that the LOCH delivery chain could significantly improve the economics of long-distance road transport, especially above 50 km. It was also identified that the heat supply requirement for releasing H2 at the end-user site and the investment costs for LOHC reactors are the most critical points. One crucial advantage of this solution is that the carriers are oil derivatives, which are already traded around the globe with plenty of loading facilities and storage capacity, turning LOHCs appropriate for routes requiring a combination of different transports (ship, rail and/or truck). Another significant advantage is the absence of boil-off losses (different from liquid H2) as the storage stability for long periods without losses or considerable costs [41].

11.2.2

Rail Transportation

Railways are an effective, precise and reliable alternative to transport large H2 volumes while relieving road traffic, especially in metropolitan areas, where an increasing deployment of H2 refuelling infrastructures is expected [42]. Rail transportation containers are similar to those for road transport, having the significant advantage of delivering several containers simultaneously. There are several types of freight wagons, depending on the kind of material they transport, and dimensions may vary per country. For example, the “GABS 8194 181” series has a maximum capacity of 50.9 tons. For diesel locomotives (1800 series), the total load to be assigned cannot exceed 400 tons, and in the case of electric locomotives (2500 series), the maximum load cannot exceed 450 tons [43]. This means that the maximum capacity for rail transport is about ten times higher than the load capacity by road. The main advantages of rail freight transport are that it is safe and fast, reduces the distance-time relation, has a low energy consumption per transported unit and allows for a greater load capacity. Its main disadvantages are rigid transport schedules, system maintenance costs, and little infrastructure flexibility.

11

Supply Chains of Green Hydrogen Based on Liquid Organic Carriers. . .

255

Transporting liquid hydrogen via tank-containers is more economical than gaseous hydrogen transport over long distances and in quantity. This is mainly due to the lower operating cost and higher storage capacity (larger mass for the same volume) [44]. However, if transport is required for a compressed-gaseous hydrogen network, both tube railway and trailer might be used [45]. Nevertheless, it is important to stress that there are fewer information sources on the cost of hydrogen rail transport since most publications concern H2 as a power source and not the logistics process involved. To overcome this challenge of quantifying the cost of rail transport, the study developed at the University of Calgary [32] was used as a proxy value for calculating the costs of transporting LOHCs by rail. In [32], a comparison was made between the cost of pipeline transport and rail transport of crude oil. The total costs considered externalities such as CO2-eq emissions, air pollution, and accident probability during transportation. It was found that the overall cost of using rail transport is about 2.5 times higher than via pipeline [32]. However, it should be noted that the mentioned study did not consider the expected energy transition in the rail transport network.

11.2.3

Pipeline Transportation

Pipeline transportation is a competitive solution for long-distance H2 transport [46]. Most global gas pipeline networks allow the transport of high energy volumes throughout different countries and continents, while ensuring strict safety standards [47]. Such assets can be refurbished to transport increasing blends of green H2 mixtures or even 100% hydrogen mixtures [48], with competitive investment costs. The refurbishments address mainly the interface equipment to ensure that the transmission and medium-pressure pipelines have high-quality stainless steel with inner and outer coating and that the low-pressure distribution is built with highquality polyethylene. Investment in refurbished assets [49] can achieve a unitary cost of 0.2–0.6 M€/km (around 1.4–4.2 M¥/km). In comparison, the dedicated new assets could be developed within a range of 1.22–3.28 M€/km (around 8.54–22.96 M¥/ km), varying with the diameter and specific terrain restrictions. The refurbishment costs are considered low when analyzing the transport costs per unit, which range between 4.8 and 75 €/MWh/1000 km (33.6 and 525 ¥/MWh / 1000 km) in the new H2 pipelines. It is imperative to emphasize that a large share of natural gas pipelines was refurbished for methane transportation. H2 has distinct calorific value thresholds from methane and different densities (about 1/8 of methane) [50]; therefore, to cope with H2 injection in refurbished natural gas pipelines, some operational and technological issues need to be addressed [51]. On the other side, low-pressure polyethylene pipelines only raise questions about their higher H2 permeability rate. H2 pipelines, namely refurbished from natural gas original pipelines, must be optimized according to the injection points, asymmetric usages for H2, and assuming different scenarios for H2 deployment [52]. The most recent H2 blending guidelines state a

256

J. Godinho et al.

theoretical concentration limit for gas appliances of up to 15%. In the following years, it is expected that to satisfy the majority of H2 consumption of household customers, entirely gaseous H2 networks could be developed. In this sense, it is vital to ensure adequate planning, construction and operational management of dedicated gaseous H2 networks, which could be a cost-effective solution to ensure future demand [51]. Pipeline transportation represents a key feature in the decarbonization strategy to create a bridge between supply and distributed demand, using existing infrastructures and technology to ramp up hydrogen transportation. Besides gas pipelines, oil pipelines could also be used to contribute to establishing H2 supply chains. More specifically, H2 can be transported when loaded onto LOHCs, such as the DBT-PDBT system. Since these products have analogous properties to crude oil-based liquids (such as diesel or gasoline), they can be transported in a liquid state under ambient conditions [52]. Furthermore, regarding oil pipelines specifically, DBT and PDBT (1040 kg/m3 and 910 kg/m3, respectively) have similar densities and kinematic viscosity compared to oil products, which allow them to use already existing oil pipeline infrastructures, with minimal adaptation [53]. Notwithstanding, there is still a gap regarding the direct utilization of oil pipelines for the transportation of LOHCs, which needs to be addressed.

11.3

Case Study

The study assumes green H2 is produced in Inner Mongolia, Tibet, Gansu, and Xinjiang provinces. The H2 is then transported to the People’s Republic of China’s three main economic and industrial clusters: the Jing-Jin-Ji cluster, the Pearl River Delta, and the Yangtze River Delta. This section describes the regions’ socioeconomic and energy demand structure and justifies the reasons behind its selection for the study.

11.3.1

China’s Main Consumption Centres: Jing-Jin-Ji, Yangtze River Delta, Pearl River Delta

The economic growth of China in recent years has been mainly guided by the development of three urban conglomerates: the Jing-Jin-Ji cluster, the Yangtze River Delta and the Pearl River Delta (Fig. 11.6). In total, these regions account for 33% of the Chinese population and represent 36% of the Chinese GDP. Nonetheless, the above-mentioned urban conglomerates are heavily dependent on fossil fuels and would have to change their energy supply radically due to the carbon neutrality strategy. Due to the high population density of the cities’ conglomerates, the cost of real estate and being home to a significant cluster of industries, the options to unfold a green energy generation base in those regions are scarce.

11

Supply Chains of Green Hydrogen Based on Liquid Organic Carriers. . .

257

Legend: Jing-Jin-Ji Cluster Yangzte River Delta Pearl River Delta Beijing

0

500km

Fig. 11.6 Main economic clusters of the People’s Republic of China [53]

The Jing-Jin-Ji cluster is the most recent urban conglomerate comprising the Beijing and Tianjin Municipalities and the Hebei Province. In total, it covers a territory of approximately 217,200 km2, has a population of around 110 million people and a GDP of ¥9.64 trillion [54–56] (€1.35 trillion [57]). The coordination and integration plan of the cities in this new megaregion started in 2014 to solve the region’s environmental pollution, overcrowding, water shortage, and uneven development. To exemplify the uneven development of the area, it is worth mentioning that Hebei has a GDP per capita (¥50 thousand) of nearly one-third of the Beijing and Tianjin Municipalities. With Beijing at its core, the cluster is China’s political and cultural centre, also an essential part of northern China’s manufacturing and logistics base. The main industries are metallurgical machinery, information technology, aviation, steel, and chemicals [58]. A crucial component of the region’s integration is developing a highly efficient internal transport network that would allow a more efficient transit of goods and make the geographical distribution of industries and labour in the area more even [59]. Regarding its energy supply, the Jing-Jin-Ji cluster still heavily relies on fossil fuels. As depicted in Fig. 11.7, fossil fuels account for more than 90% of the region’s energy consumption (2962 TWh). They are vital to supply the distributed heating network of Tianjin and Beijing, as well as its industry. Compared with the other economic clusters, it is noticeable that the Jing-Jin-Ji has a much higher coke gas final consumption (470 TWh), which its robust chemical and industrial sector may explain. The most significant renewable source is wind power, which accounts for 6.2% of the total electricity consumption. Wind power is mainly generated in the Hebei province [54–56].

258

J. Godinho et al.

Coal Gangue and Briquettes Solar PV power Wind power Hydro power Thermal power Coke Gas Petroleum Products Natural Gas Heat Electricity Other

10.1%

11.1%

0.6%

3.8% 6.2% 0.4%

20.2% 20.9% 89.6% 15.9% 8.3%

12.9%

Fig. 11.7 Total final energy consumption of the Jing-Jin-Ji cluster, 2020

The Yangtze River Delta is comprised of the Shanghai Municipality and the Jiangsu, Anhui, and Zhejiang provinces. It covers 355,240 km2, and the total population is 236 million, factors that turn the cluster into the largest megaregion in the world. The GDP of the region is ¥27 trillion [60–63] (€3.81 trillion [57]), accounting for 25% of the Chinese GDP, despite holding only 16% of the total population. The Yangtze River Delta is China’s financial and economic hub, contains the largest stock exchange in mainland China, and represents nearly 40% of the country’s total foreign direct investment inflows. Moreover, the region is a vital hub for the scientific and technological communities, holding several of the world’s top 100 universities and a wide range of national laboratories and engineering centres. To develop the region’s integration level, local governments are committed to increasing the urbanization rate to 70% and investing heavily in digital infrastructures, such as 5G networks [64]. The region has a high fossil fuel dependency despite the positive economic indicators. Overall, the region’s total final energy consumption (TFC) is 4,285 TWh, the highest consumption of the areas in the study. The Yangtze River Delta has the largest share of electricity (35.5%, 1,520 TWh) in the overall final consumption of the three economic clusters. In the electricity sector, fossil fuels account for 85.9% (1,306 TWh) of the total. Low-carbon power accounts for 14.1% (214.3 TWh) of the total electricity consumption. Nevertheless, with the reduced share of low-carbon sources in the power mix, the regional governments are committed to enhancing their renewable sector mainly by developing wind offshore [65]. The final consumption mix of the regions is displayed in Fig. 11.8 [60–63]. Lastly, the study also focused on the Pearl River Delta, a region with an area of 180,901 km2 and a cumulative population of 135 million people. The region comprises Guangdong province and the Special Administrative Regions of Macau and Hong Kong, whose coordinated development has significantly boosted efficiency in the region. The Pearl River Delta has a total GDP of ¥15.8 trillion (€2.24 trillion) and was one of the earliest regions in China to open to foreign

11

Supply Chains of Green Hydrogen Based on Liquid Organic Carriers. . . 0.9%

4.4% Coal Gangue and Briquettes Nuclear Solar PV power Wind power Hydro power Thermal power Coke Gas Petroleum Products Natural Gas Heat Electricity Other

259

11.2% 7.0%

2.8% 2.2% 2.0%

35.5%

34.8%

85.9%

6.6% 2.4% 4.3%

Fig. 11.8 Total final energy consumption of the Yangtze River Delta, 2020 0.6% Coal Gangue and Briquettes Nuclear Solar PV power Wind power Hydro power Thermal power Coke Gas Petroleum Products Natural Gas Heat Electricity Other

4.4% 1.5% 20.3%

49.9%

27.5% 71.6%

2.3% 5.4%

1.3% 1.8% 5.0%

8.3% 8.4%

Fig. 11.9 Total final energy consumption of the Pearl River Delta, 2020

trade under the reforming policies implemented by Deng Xiaoping [66–68]. The region has a diverse economic sector with Guangdong, especially the cities of Shenzhen and Guangzhou, being known for high technology and digital services, Macau focuses on tourism, cultural and gambling activities, and Hong Kong is regarded as one of the most services-oriented economies in the world, with services sectors accounting close to 90% of the GDP. The Pearl River Delta has a lower dependency on fossil fuels than the previous megaregions, primarily due to nuclear power. In the electricity sector, low-carbon sources (renewable + nuclear) accounted for 28.4% of the mix by the end of 2020 (571.9 TWh) (Fig. 11.9). The overall energy consumption of the region was 2081.6 TWh in 2020, and leading suppliers of energy were petroleum products. The region has a low heat demand compared to the above-mentioned economic clusters.

260

11.3.2

J. Godinho et al.

China’s Main Renewable Power Producer Provinces: Xinjiang, Tibet, Inner Mongolia, Gansu

A viable decarbonization path for China’s Main Consumption Centers, Jing-Jin-Ji, Yangtze River Delta and the Pearl River Delta, is to allocate a large share of the green energy generation industry to more remote areas. After analyzing China’s solar irradiation, wind speed temporal and geographical distribution, and population density, it was assessed that most of the green H2 could be produced in the province of Gansu, Inner Mongolia, Tibet and Xinjiang Uygur Autonomous Regions. The four regions include some of China’s richest areas in terms of solar and wind resources, as illustrated in Fig. 11.10. Figure 11.11 shows the location of the provinces under study. This section provides a short description of each province, focusing on the relevant features within the context of the green H2 supply chain. The Tibet Autonomous Region covers an area of 1,228,400 km2, accounting for 1/8 of the total area of China, but it only has a population of 3.66 million. Tibet’s GDP in 2021 was ¥208.02 billion (€29.6 billion [57]). In the region’s GDP, the added value of the primary, secondary and tertiary industries accounted for 7.9%, 36.4% and 55.7%, respectively. Tibet’s main economic activities are dairy and cattle farming, ecological tourism, and renewable energy [71]. Nevertheless, the fast development that the region had in the last decades, its small population, and geographical barriers, such as the average altitude of more than 4000 m, mountain ranges and large canyons, still impair the construction of transport networks and, consequently, hinder its economic growth. Due to a lack of available data, the scrutiny of the region’s energy consumption is focused only on the power sector. The analysis of the electrical structure of the region emphasizes the potential of Tibet’s renewable resources to supply China’s energy needs. In 2020, Tibet generated 9.15 TWh, 95.5% of which was provided by renewable power plants, and the majority was devoted to the supply of other provinces (Fig. 11.12) [72]. By 2025, renewable power plants are expected to have a capacity of 15 GW, and the exporting capacity of electricity will grow due to investments in ultra-high voltage direct current (UHVDC) transmission [73]. The Xinjiang Uygur Autonomous Region is the largest province of China, covering an area of 1,664,900 km2. It has a permanent resident population of 25.89 million and a GDP of ¥1,598.37 billion (€226 billion [57]). The added value of the primary industry accounted for 14.7% of the regional GDP, the secondary industry’s added value corresponds to 37.4%, and the added value of the tertiary sector accounted for 47.9%. Xinjiang is a significant petroleum and natural gas producer with robust iron, steel, and cement industries. It is home to a vital pipeline network transporting natural gas and oil from the province to China’s east coast [74]. The region relies heavily on coal, but it is rapidly increasing its share of renewable sources, especially in the power sector. In 2020, the percentage of wind and solar PV power increased by 7.3% and 14.3%, respectively, compared to 2019. To help achieve the country’s decarbonization targets, the State Grid of China plans

11

Supply Chains of Green Hydrogen Based on Liquid Organic Carriers. . .

261

Fig. 11.10 (a) China’s average global horizontal irradiation [69] and (b) land-based wind resource potential [70]

262

J. Godinho et al.

Fig. 11.11 Green hydrogen production provinces studied, People’s Republic of China [53] Fig. 11.12 Total final electricity consumption of the Tibet Autonomous Region, 2020

4.5% 3.1% 15.5%

Thermal power Wind power Solar power Hydro power

76.8%

to construct new UHVDC power lines connecting Xinjiang to other provinces [75]. Xinjiang’s total final consumption for 2020 was 1507 TWh; its distribution is shown in Fig. 11.13. Gansu is located in Northwest China, has a total area of 425,800 km2, and a resident population of 24.9 million. In 2021, the GDP of Gansu was ¥1,024.33 billion (€145.6 billion), with a ratio of the primary, secondary and tertiary industries of

11

Supply Chains of Green Hydrogen Based on Liquid Organic Carriers. . .

263

2.7% Coal Gangue and Briquettes Solar PV power Wind power Hydro power Thermal power Coke Gas Petroleum Products Natural Gas Heat Electricity Other

10.8% 0.6%

11.5%

0.8% 5%

5.4%

2.5% 8.4%

21.1% 84.1% 47.2%

Fig. 11.13 Total final energy consumption of the Xinjiang Uygur Autonomous Region, 2020

1.6% Coal Gangue and Briquettes Solar PV power Wind power Hydro power Thermal power Coke Gas Petroleum Products Natural Gas Heat Electricity Other

14.9%

8.9%

0.4% 9.7%

4.2%

4.4%

6.8% 24%

49.1% 36.8%

39.3%

Fig. 11.14 Total final energy consumption of the Gansu Province, 2020

13.3%, 33.8%, and 52.9%, respectively. Gansu has an important mineral extraction and fossil fuel industry. Gansu is one of the wealthiest renewable sources areas in China. In the power sector, renewables account for 50.9% of the final consumption [76]. Gansu’s total final consumption for 2020 was 572 TWh; its distribution is shown in Fig. 11.14. The last province considered in the case study is the Inner Mongolia Autonomous Region. Inner Mongolia is located in Northern China. It has an area of 1,183,000 km2 and an aggregated population of 24 million. The annual GDP of Inner Mongolia in 2021 reached ¥2,051.42 billion (€291.5 billion) with a ratio of the primary, secondary and tertiary industries of 9.0%, 39.3%, and 51.7%, respectively. Recent years have seen an increase in the number of projects related to green hydrogen and wind power in the province. The region has a high wind resource potential and a solar irradiation index well above the country average, making it one of the most promising areas of China to develop a green power supply industry. Nevertheless, the province is still highly dependent on coal in the total energy consumption mix, and it’s one of the world’s most relevant coal mining regions.

264

J. Godinho et al. 0.2% 2.2%

3.5% 1.1%

Coal Gangue and Briquettes Solar PV power Wind power Hydro power Thermal power Coke Gas Petroleum Products Natural Gas Heat Electricity Other

5.2%

4.7%

0.5%

75.9%

18.2% 1.4%

11.2% 75.8%

Fig. 11.15 Total final energy consumption of the Inner Mongolia Autonomous Region, 2020

By the end of 2020, renewable energy sources still accounted for 18.3% of the electric power production of the province [77]. The region’s total energy consumption in 2020 was 3,561 TWh (Fig. 11.15). The review of the socio-economic indicators per province mentioned above and the comparison with the indicators present in Sect. 11.3.1 reinforce the notion that the development of a green energy industry in those areas can contribute to reducing the regional imbalance between China’s west and east, create more job opportunities and promote equitable development and resource distribution in the country [78].

11.4 Methodology The current section presents the methodology to access the amount of green H2 that can be produced in China’s remote provinces, the most cost-effective strategy of transportation and the amount of CO2-eq emissions saved per year due to the implementation of a decarbonization strategy based on hydrogen. The section is divided into three major parts: 11.4.1 Supply Chain Steps, 11.4.2 Green H2 CostPrice, and 11.4.3 CO2-eq Emission Savings.

11.4.1

Supply Chain Steps

The supply chain steps comprise the production of H2 via electrolysis, supported by renewable power plants in Tibet, Xinjiang, Gansu and Inner Mongolia, and its transportation to the Jing-Jin-Ji cluster, the Yangtze and the Pearl River Deltas. This section explains the process of analysing the production potential of green H2, the demand for H2 in the economic centres and the framework to decide the transportation methods and routes.

11

Supply Chains of Green Hydrogen Based on Liquid Organic Carriers. . .

11.4.1.1

265

Potential for Green H2 Production

As presented in the case study section, the proposed supply chain of green H2 within China will be established across a considerably large region of the country. Starting with the supply potential, data was collected to determine the total amount of green H2 produced within the considered production provinces and available for transportation. To this effect, several steps were required. Firstly, considering only the potential to create green H2, it was necessary to collect data regarding both the projected levels of electricity consumption, as well as the forecasted levels of electricity production via renewable energy sourcesRES in the study’s provinces. By collecting data regarding future projections for the consumption of electricity in China [79] and adjusting these values to forecasts for population growth of the production provinces [80] within the same timeframe, it was possible to extrapolate the respective values until 2060. This last step was required since it has been stated that changes in population within China will occur unevenly, depending on the region of the country. Regarding the production provinces considered, Tibet’s population is projected to peak within the next decade, with a subsequent drop of nearly 25% until 2060. Inner Mongolia’s population is projected to peak around 2030. This is followed by a decline, which projects the population to be slightly lower in 2060 than the levels for 2010. Finally, both provinces of Gansu and Xinjiang are considered within the same region of China (Northwest China) and thus were assumed to have the same population trend, due to the lack of more detailed data. The population is projected to peak between 2030 and 2040 for these two regions, with a considerably low decline until 2060. However, it is noteworthy that these regions are expected to have the lowest variation in population within the considered timeframe. By collecting this data, it was possible to assess the projected levels of electricity consumption (ConsumEl) within the considered production provinces (pp) for each year between 2020 and 2060, through Eq. 11.1:

ConsumEl,year =

ConsumEl,year - 1=Pop

pp,year - 1

×

ConsumEl,yearMC=Pop

ConsumEl,year - 1MC=Pop

MC,year

× Poppp,year

MC,year - 1

8pp, year ð11:1Þ where Pop = Projected level for population MC = China Note that for the years before 2020, historical data was retrieved from [81]. Following the projection for electricity produced using RES within the production provinces, it was necessary to source the expected capacity levels until 2060. For

266

J. Godinho et al.

Table 11.1 Projected values for solar PV and wind capacity within the production provinces for 2025 and 2050 and for the LOW and HIGH scenarios 2025

Region Gansu Inner Mongolia Tibet Xinjiang

Solar PV [MW] 41,690 45,000 10,000 49,440

2050 Wind (hydro for Tibet) [MW] 38,530 89,000 15,000 32,960

Solar PV [MW] Low High 100,000 500,000 200,000 550,000 10,000 250,000 350,000 600,000

Wind (hydro for Tibet) [MW] Low High 100,000 150,000 925,000 1,500,000 15,000 15,000 150,000 250,000

this sub-step, it was assumed that only electricity production through wind energy and solar PV would be included, except for the region of Tibet, where data was collected regarding hydropower capacity. The assumed values were collected from the expected power capacity for both 2025 and 2050. Accounting for the lack of historical data, electricity production from RES was only calculated from 2025 onwards. Furthermore, to account for the higher levels of uncertainty, two different scenarios were considered for the projected RES capacity in 2050. The two scenarios are designated as “LOW” and “HIGH,” depending on whether the projected capacities are either low or high. These values were then used to compute the numbers for the remaining years through linear interpolation, while for the years between 2050 and 2060, it was assumed that the respective capacity would remain fixed to the levels of 2050. This latter assumption was considered since the levels of available green H2 for transportation are projected to be high enough in 2050 to satisfy the projected consumption levels for the selected provinces, as presented in the results section. The respective values for projected solar PV and wind capacity for each of the production provinces and capacity scenarios are shown in Table 11.1. Henceforth, the formulas presented in the methodology section are adjusted for each scenario. The following required step was to account for the respective RES alternatives’ capacity factors (or load factors). Since RES are naturally variable, these capacity factors depend heavily on geographical regions and weather conditions. This means that, for large countries such as China, it is crucial to assess whether the availability of both wind and solar energy varies, depending on the region of the country. To this effect, it was collected historical data on the capacity factors of wind production for the province of Inner Mongolia and Gansu, which is assumed to be similar to the achievable numbers within the two other production provinces for 2020. For wind energy, a capacity factor of 55% was considered fixed across the accounted timeframe. Figure 11.16 shows the average daily profile for a typical wind farm in Inner Mongolia [82]. Regarding solar energy, it was identified that the annual availability varies significantly throughout different regions of China. Due to lower levels of yearly irradiance, the autonomous region of Inner Mongolia was considered to have a 19%

Supply Chains of Green Hydrogen Based on Liquid Organic Carriers. . .

11

Fig. 11.16 Average hourly wind generation per MWp installed in Inner Mongolia

267

Wind generation (MWh/MWp)

0.65

0.60

0.55

0.50

0.45

0.40

00 00 00 00 00 00 00 00 00 00 00 00 00: 02: 04: 06: 08: 10: 12: 14: 16: 18: 20: 22:

0.9

0.9

0.8

0.8

Solar PV generation (MWh/MWp)

Solar PV generation (MWh/MWp)

Period (h)

0.7 0.6 0.5 0.4 0.3 0.2

0.7 0.6 0.5 0.4 0.3 0.2

0.1

0.1

0.0

0.0

00 00 00 00 00 00 00 00 00 00 00 00 00: 02: 04: 06: 08: 10: 12: 14: 16: 18: 20: 22:

Period (h)

00 00 00 00 00 00 00 00 00 00 00 00 00: 02: 04: 06: 08: 10: 12: 14: 16: 18: 20: 22:

Period (h)

Fig. 11.17 Average hourly solar PV generation of each MWp installed in Inner Mongolia (left) and Western regions (right)

capacity factor for 2020. Contrarily, the provinces of Gansu, Xinjiang and Tibet, which belong to the country’s Western regions, are assumed to have a higher annual capacity factor of 31% in 2020. These values were collected through historical data [83]. Figure 11.17 presents the average hourly solar PV generation profile for Inner Mongolia (left) and the Western Chinese Regions (right). In Fig. 11.17, it is verified that Inner Mongolia has an average daily production profile starting at 6 h and finishing at 19 h; the generation reaches its peak at 12 h. The Western regions present an atypical generation profile with a zenith of production at 10 h and other at 15 h; this behaviour may be justified by the long distances encompassed in the region, which are separated thousands of km apart, and as an effect of the Chinese single time zone.

268

J. Godinho et al.

It is noteworthy that, according to the literature, the annual capacity factors for solar energy were assumed to improve slightly in the coming decades by a factor of 0.9% per decade. This is explained since there are expected improvements in air quality, mainly due to reduced atmospheric pollution [84] and technological innovations. Thus, for the calculation of the total potential for electricity production through RES, for both the LOW and HIGH scenarios, Eq. 11.2 was used:

hours year 8year

ProdEl,year = CapW,year × CFW þ CapPV,year × CFPV,year × 8760

ð11:2Þ where Prod El = Projected level for electricity production via RES [Wh] Cap = Projected level for total power capacity [W] CF = Capacity factor [%] W = Wind energy PV = Solar PV By assessing the expected levels for production and consumption, i.e., supply and demand, it was further necessary to calculate the remaining electricity, which could potentially be used for green H2 production. Nonetheless, before this assessment, it was assumed that energy storage through pumped hydropower storage (PHS) and electricity transmission through UHVDC gridlines would be prioritized over the power demand of electrolyser units. PHS systems were considered the first application for surplus renewable electricity. This decision assumed that each province aimed to decarbonise its own energy consumption first, thus ensuring that stored energy was available whenever there was not enough renewable electricity to satisfy the local needs. Furthermore, it is noteworthy that the assessed energy stored for 1 year is assumed to be converted to electricity for consumption the following year. Finally, based on historical data on the Tiantai PHS system in Zhejiang province, China, a 15% capacity factor was assumed for the storage operation [85]. The projected storage capacity levels that were retrieved for this study for each of the respective production provinces are presented in Table 11.2. As Table 11.2 highlights, the Gansu province is projected to slightly increase storage capacity throughout the next decade. This is due to new PHS installations expected to be operational by 2030. In the case of Inner Mongolia, the storage

Table 11.2 Projected storage capacity per province

Region Gansu Inner Mongolia Tibet Xinjiang

Until 2026 (MW) 8,742 2,424 90 2,400

Until 2030 (MW) 13,000 240,549 90 2,400

11

Supply Chains of Green Hydrogen Based on Liquid Organic Carriers. . .

269

capacity levels are expected to increase by a factor of roughly one hundred. However, it is also assumed that beyond 2025, this province will have a new requirement of minimum electrochemical storage capacity (mainly utility-scale batteries). The storage requirement is supposed to be 15% of the total RES electricity production capacity until 2060 [86]. However, considering that capacity factors for utility-scale batteries typically fall within the 9–16% range [87], we maintained consistency by applying the same capacity factor used for the Pumped Hydro Storage systems in this scenario as well. Collecting this data and adjusting the storage levels across the accounted timespan made it possible to calculate the amount of excess electricity after accounting for PHS storage; this was performed by satisfying the following conditional Equation 11.3: ProdEl,year; pp - ConsumEl,year; pp - StorCAP,pp,year - StorCAP,pp,year - 1 × CFStor × 8760

hours >0 year 8year; pp

ð11:3Þ where StorCAP = Projected level for storage capacity [W] Stor = Energy storage When local renewable electricity production for a specific year, plus the stored energy from the previous year, is assumed to be sufficient to satisfy the total electricity demand, electricity transmission through UHVDC is deemed the next most viable option. In this regard, data were collected regarding the existing infrastructure for 2020, which can connect the production provinces directly with the consumption regions [88]. It was assessed that all the production provinces, except for Tibet, have existing transmission infrastructure lines connecting these regions with Central and Eastern China, including the Jing-Jin-Ji and the Yangtze River Delta regions. Furthermore, accounting for the projected investments [89, 90] in constructing other UHVDC transmission lines, it was assumed that the respective capacity for each assessed region would increase by 25% each decade until 2060. The data collected for the UHVDC transmission capacity is presented in Table 11.3. According to Table 11.3, it was assumed that electricity transmission through UHVDC would be the first option for the remaining electricity produced through RES since UHVDC transmission is the most efficient alternative for the transportation of electricity [91]. Compared to other options, which mainly involve converting electrical energy to either chemical energy (batteries and hydrogen) or gravitational energy, in the case of PHS, the transmission grid allows for the direct transportation of electricity with minor losses. More specifically, it was assumed that a 3% total loss of electricity would occur regarding UHVDC transmission [92]. Additionally, to ensure that the grid has enough available capacity to match the renewable electricity generation, the grid capacity factor was assumed to be equivalent to the wind power capacity factor. This entails that no amount of produced renewable electricity is wasted before transmission. Thus, for every year that there was a surplus of electricity, the following conditional Eq. 11.4 was applied:

270

J. Godinho et al.

Table 11.3 UHVDC transmission capacity for each production province and their respective destinations for 2020

To Henan province (between Yangtze River Delta and Jing-JinJi) Anhui province (Yangtze River Delta) Hunan province (close to Yangtze River Delta) Henan province (goes through Gansu to Yangtze River Delta) Zhejiang province (Yangtze River Delta) Shandong province (between Yangtze River Delta and Jing-JinJi) Total

From Gansu (GW) –

Inner Mongolia (GW) –

Tibet (GW) –

Xinjiang (GW) 8

12 –

8 8 –

10 10

16

30

IFðUHVDCCAP,pp × UHVDCEff × CapMultyear × CFW × 8760

20

hours Þ year

< SurplusEl,year,pp THEN : ðUHVDCCAP,pp × UHVDCEff × CapMultyear hours × CFW × 8760 Þ year ELSE : SurplusEl,year,pp 8year, pp

ð11:4Þ

where UHVDCCAP = Total capacity for electricity transmission through UHDVC [W] UHVDCEff = Transmission efficiency (97%) CapMult = Capacity multiplier, which increases 25% each decade SurplusEl = Excess electricity from RES, after accounting for consumption and PHS storage [Wh] Equation 11.4 allows the computation of the projected levels for electricity transmission through UHVDC. By deducting this amount from the previously calculated SurplusEl, it was possible to finally quantify the projected amount of electricity available to produce green H2 within the production provinces. Assuming that all the available electricity is then used to power electrolyser units, the total potential to create green H2 was computed through Eq. 11.5:

11

Supply Chains of Green Hydrogen Based on Liquid Organic Carriers. . .

TotH2 ,year,pp =

ðSurplusEl - UHVDCEl Þyear,pp ElectrolyserEff,year 8year, pp

271

ð11:5Þ

where TotH2 = Total potential for green H2 production [kg] UHVDCEl = Projected amount of electricity transmitted through UHVDC ElectrolyserEff = Electrolyser efficiency [kWh/kg-H2] It was deemed that electrolyser efficiency would increase between 2020 and 2050, from 66.50 kWh/kg-H2 or around 50% to 42.50 kWh/kg-H2 or roughly to 78%, respectively [93]. In this study, it was assumed that efficiency for the years in between would be calculated through linear interpolation. The same linear trend was assumed for the years between 2050 and 2060, culminating in an energy requirement of 34.50 kWh/kg-H2, or roughly 97%. For the latter years, the following linear Eq. 11.6 was used: ElectrolyserEff,year = - 0:8 × Year þ 1, 682:50 82050 < Year ≤ 2060

ð11:6Þ

Following the calculation for the total amount of green H2 produced, further details regarding the energy consumption of the production provinces were considered. These values were extrapolated until 2060, following the same method used for electricity consumption. Furthermore, gasoline and diesel consumption were also considered in terms of energy density. These were assumed to account for the energy needs for transportation. An overall internal combustion engine efficiency of 20% was assumed to account for the final energy consumption for transport [94]. These values were considered the total final energy consumption for each production province. By factoring in the electricity generated from fossil fuels and accounting for the heat and final energy consumption derived from gasoline and diesel, it becomes feasible to calculate the overall volume of final energy consumption requiring decarbonization. Before assessing the total available green H2 for transportation, it was necessary to fully account for the amount required to decarbonise the producing provinces. Following the same reasoning considered for renewable electricity, this allows each producing region to prioritize the decarbonisation of their energy mix first. Each year, the available H2 is deemed to be divided across each end-use. However, these shares were assessed based on the respective values for each production province’s total primary energy supply (TPES). Thus, for the remaining electricity and heat demand, the produced green H2 was assumed to be used to power gas turbines at CHP plants, which are deemed to operate at a 60% energy-based efficiency for the entire timespan of the study. Contemporary CHP systems, both existing and newly installed, can incorporate a significant proportion of hydrogen, ranging from 20% to 100%, as indicated by equipment manufacturers. It is noteworthy that, in the coming decade, emerging CHP systems are expected to

272

J. Godinho et al.

have the capability to combust 100% green hydrogen.Regarding transportation, the produced green H2 was used to power fuel-cell electric vehicles, which were considered to convert 25% of the input energy. To assess the amount of green H2 required, in mass terms, Eq. 11.7 was used:

DEMH2 ,year,pp =

ðElFF þ HeatFF Þyear,pp TranspFFyear,pp þ ηCHP ηFCEV Wh 33:ð3Þ =g - H2 8year, pp

ð11:7Þ

where DEMH2 = Total demand for green H2 [kg] ElFF = Electricity produced using fossil fuels [Wh] HeatFF = Heat produced using fossil fuels [Wh] TranspFF = Final energy use for transportation produced using fossil fuels [Wh] By assessing both the total potential production of green H2, as well as the specific demand within each of the producing provinces, it was possible to compute the total amount of green H2 available for transportation using Eq. 11.8: TRANSPH2 ,year,pp = TotH2 ,year,pp - DEMH2 ,year,pp 8year, pp

ð11:8Þ

where TRANSPH2 = Total amount of H2 available for transportation [kg]

11.4.1.2

Projected Demand for Green H2

After calculating the amount of green H2 available for transportation, i.e., the supply profiles for each producing province, it was further required to highlight the demand profile for each consumption region. As for the case of the production provinces, data was collected regarding the historical consumption of electricity, heat, and transportation (gasoline and diesel). The collected values were respective to 2020 and were extrapolated until 2060, following the same methodology expressed in previous sections. For heat and transportation, it was assumed that the extrapolated consumption values were entirely targeted for decarbonisation. For the case of electricity, however, local RES production was accounted for, similarly to the producing provinces. Firstly, considering the historical data for 2020, RES contributed for each consumption region, even though it was relatively small. More specifically, it was assumed that renewable electricity accounted for roughly 10.4% of the overall electricity consumption for the Jing-Jin-Ji region. This amounted to approximately 14.1% and 28.4% for the Yangtze and Pearl River

11

Supply Chains of Green Hydrogen Based on Liquid Organic Carriers. . .

273

Table 11.4 Projected values for solar PV and wind capacity within the consumption regions for 2025 and 2050 and for the low and high scenarios 2025 Region Jing-Jin-Ji Yangtze River Delta Pearl River Delta

Solar PV [MW] 48,600 53,500

Wind [MW] 56,000 32,410

2050 Solar PV [MW] Low High 60,000 60,000 231,000 232,000

10,000

15,000



5000

Wind [MW] Low High 125,000 325,000 30,000 210,000 –

30,000

Deltas regions, respectively [81]. Furthermore, future projections for deploying RES capacity within these regions were also accounted for and are presented in Table 11.4 [79]. Converging with the methods used for the production provinces, the values for wind and solar capacity between 2025 and 2050 were computed through linear interpolation. Similarly, from 2050 to 2060, the capacity values were fixed. Furthermore, regarding the associated capacity factors for solar, it was assumed that all three consumption regions would follow the same trend used for the province of Inner Mongolia. This assumption was based on the findings for average irradiation, presented in Fig. 11.17, which shows that these regions have a relatively lower irradiation profile, especially compared to the provinces of Tibet, Xinjiang and Gansu. After discounting the expected supply of renewable electricity, it was also necessary to subtract the levels of transmitted electricity originating from the considered production provinces. By encompassing all these values, Eqs. 11.9 and 11.10 were used to calculate the total final energy consumption deemed for decarbonisation, i.e., total demand for the transported green H2:

DEMH2 ,year,cr =

ðElFF þHeatFF Þyear,cr ηCHP

þ

TranspFFyear,cr ηFCEV

33:ð3Þ Wh=g - H2

ð11:9Þ

where ElFF,year,cr = TotEl,year,cr - RESEl,year,cr þ UHVDCEl,year,cr 8year, cr

ð11:10Þ

TotEl = Total demand for electricity [Wh] RESEl = Electricity produced via RES [Wh] cr = Consumption regions

11.4.1.3

Transportation

With the available supply and demand levels for green H2 assessed, the following step relates to transportation planning. Two transportation options were considered:

274

J. Godinho et al.

Option 1: Focuses on transporting pure green H2, either in its gaseous state or by converting the H2 through liquefaction. The transportation strategy revolves around giving precedence to existing gas pipelines, enabling direct conveyance of H2 from the electrolyser’s exit point. Before transportation, it is only necessary to have a compression phase since green H2 usually leaves the electrolyser at a standard pressure of 30 bar. For transportation, a pressure between 40 and 80 bar is required [95]. In the case of insufficient pipeline capacity to transport the total amount of green H2, this first option assumed that the remaining amount was subject to liquefaction and further transported, firstly through railway and finally via road transportation. It was assumed that the production provinces would be able to integrate enough capacity for the compression and the liquefaction of the transported green H2. At the consumption point, the respective regions could incorporate the necessary infrastructure for decompression or regasification of the transported H2. Option 2: The second option referred to the alternative of transporting green H2 when stored within LOHCs. More specifically, this study focused on the DBT-PDBT system, the most financially viable candidate via two-way LOHCs [96]. Similarly, different transportation modes were accounted for. First, refined and crude oil pipelines were prioritized, followed by railway and road transportation. In this case, it was considered that the required hydrogenation capacity would be available within each production province. Also, following the same method for the last option, the consumption regions were deemed to be able to incorporate the necessary infrastructure capacity for the storage and dehydrogenation of the LOHCs, before consuming the transported H2. When considering the different transportation modes, it is important to mention several assumptions taken. Starting with gas pipelines, data were collected regarding the capacity of existing connections between the production provinces and the consumption regions [97], which is presented in Table 11.5. Only two assessed production provinces were considered to transport capacity via gas pipelines. In the case of Tibet, this is explained since investment in transportation infrastructure is low, mainly due to geographical constraints. This study assumes no gas pipelines are built from Tibet outwards for the entire timespan. Regarding Gansu, no gas pipeline capacity was considered. This is justified since all the existing pipelines that either pass or start in Gansu, currently, begin or can connect to pipelines starting in the province of Xinjiang. Further assumptions include a 95% annual load factor for the use of the pipelines [98] and a capacity multiplier, which allowed projecting the impact of this specific transportation mode until 2060. To this effect, the retrieved total capacity levels were assumed to increase by 50% until 2040 and by 100% until 2060, with the respective capacity levels increasing linearly between milestones. Finally, this study considered transporting either a blend composed of 20% pure H2 and 80% natural gas and another option considering only repurposed pipelines transporting 100% pure H2. A similar approach was selected for the oil pipelines, starting with the existing refined petroleum products and crude oil pipeline capacity between the production

11

Supply Chains of Green Hydrogen Based on Liquid Organic Carriers. . .

275

Table 11.5 Current gas pipeline capacity between the production provinces and the consumption regions, with the respective length in km, for 2020

To Zhejiang (YRDa) Shanghai (YRDa) Hong-Kong (PRDa) Guangdong (PRDa) Beijing (JJJa) Beijing Outer Beijing Jilin (JJJa) Heilongjiang (JJJa) Total a

Capacity (m3/day) From Gansu Inner Mongolia – – – – – – – – – 15,007 – 11,969 – 558 – 372 – 496 – 28,402

Tibet – – – – – – – – – –

Xinjiang 17,984 7,193 17,984 17,984 – – – – – 61,146

Length [km] 4,159 4,200 4,895 5,278 1,082 1,279 254 232 156

YRD Yangtze River Delta region, PRD Pearl River Delta region, JJJ Jing-Jin-Ji region

Table 11.6 Current oil pipeline capacity between the production provinces and the consumption regions, for 2020

Refined

Crude

a

To Lanzhou (Gansu) Shanghai (YRDa) Beijing (JJJb) Hong-Kong (PRDc) Lanzhou (Gansu) Hong-Kong Beijing Beijing and Shanghai area Shanghai Total

Capacity in m3/day From Gansu Inner Mongolia – – – 27,584 27,584 – – 18,442 – – – 27,663 – 55,208 – 13,831

Tibet – – – – – – – –

Xinjiang 36,885 – – – 55,208 – – –

Length [km] 1,930 4,429 4,374 3,512 1,737 880 2,471 578

– 101,274

– –

– 92,093

979

80,964 150,004

YRD Yangtze River Delta region, bPRD Pearl River Delta region, cJJJ Jing-Jin-Ji region

and consumption regions. This data is presented in Table 11.6. It is noteworthy that for the oil pipeline grid in China, the city of Lanzhou in the province of Gansu represents a splitting point. Here, the oil pipeline infrastructure originating from Xinjiang connects to the central oil pipeline grid, which supplies both the Jing-Jin-Ji and the Yangtze River Delta regions and the southwestern grid, which connects to the Guangdong in the Pearl River Delta region. For this reason, it was assumed that the LOHCs transported from Xinjiang and Gansu would be evenly distributed between the three mentioned consumption regions, i.e., 33.3% for each. Apart

276

J. Godinho et al.

Table 11.7 Current railway infrastructure connections between the production provinces and the consumption regions for 2020 Departure Region/city Inner Mongolia

Tibet

Xinjiang

Hohehot Hohehot Baotou Hohehot Lasa Lasa Lasa Nyingchi Urumqi Urumqi Urumqi Hami Hami

Arrival City Beijing (JJJ) Shanghai (YRD) Guangzhou (PRD) Guangzhou (PRD) Beijing Shanghai Guangzhou Beijing Beijing Shanghai Guangzhou Beijing Shanghai

Length [km] 660 2,346 3,112 2,947 3,757 4,373 4,980 3,512 3,144 3,600 4,684 2,600 3,524

from this, it is also important to note that, as for the case with gas pipelines, the Tibet Autonomous Region is also assumed not to have any available transmission capacity through oil pipelines across the entire timespan. The annual load factor considered for the transmission was 60% [98]. Additionally, following the same approach for the gas pipelines, the overall transmission capacity was assumed to increase by 50% by 2040 and 100% by 2060. Railway was the next mode considered for transportation of the demanded green H2. Converging with the approach followed for the pipeline infrastructure, data was first gathered regarding the existing railway connections between the production provinces and the consumption regions. This data is presented in Table 11.7 and provides the specific cities of departure and arrival. Since all the connections that cross the province of Gansu originate in Xinjiang, it was assumed that the total capacity of these respective connections would be allocated to the supply of green H2 from Xinjiang. To address the annual capacity of the railway infrastructure, it was necessary to outline the amount of cargo each train can carry, as well as the number of trips that could be travelled for each rail connection, annually. Firstly, each train can carry several rail cars. Each rail car can contain a maximum of ten containers, with two rows of five containers stacked on top of each other [99]. An illustration of this configuration is shown in Fig. 11.18. For this study, it was considered that each train would have a maximum capacity of 12,500 tonnes, with each rail car carrying a maximum of 60 tonnes or 10 containers [100]. This allows for a total of 209 rail cars per train, assumed for the entire timespan of this study. In each rail car, the number of containers transported was constrained by the mass limit of the rail car as well as the volumetric capacity of the containers. As mentioned in Sect. 11.2.2, the volumetric capacity of a cryogenic container in roughly 50 m3, while for LOHCs, this capacity

11

Supply Chains of Green Hydrogen Based on Liquid Organic Carriers. . .

277

One car

One car

Fig. 11.18 Example train configuration, displaying the maximum layout of 10 containers per rail car

Table 11.8 Current length for road transportation between the production provinces and the consumption regions. Values for 2020 and in km

To Beijing Shanghai Guangdong

Length [km] From Gansu Inner Mongolia 1,441 480 2,002 1,726 2,260 2,304

Tibet 3,856 3,954 3,555

Xinjiang 2,766 3,924 4,181

is slightly lower at 36 m3. Considering a density value of 70.85 kg/m3 for LH2 and of 1040 kg/m3 and 910 kg/m3 [101] for DBT and PDBT, respectively, this results in a maximum of 10 containers per rail car for LH2 and approximately two containers for the DBT-PDBT system. The annual load factor of the railway infrastructure was deemed to equal the projected schedules for green H2 production, thus assuming that transportation would be available whenever the electrolysers are operating. Furthermore, it was assumed that the speed at which cargo trains are expected to travel will increase, following a linear trend from an average of 200 km/h in 2020 to an average speed of 600 km/h in 2060 [102]. This will result in shorter amounts of time spent per trip, which, in turn, will allow for a considerably larger number of trips per year, i.e., the higher overall capacity of the railway infrastructure per year. Apart from the data highlighted in Sect. 11.2.2, assessing the respective distances between the production provinces and the consumption regions was essential. These data are presented in Table 11.8. The final transportation mode considered was via road, using trucks, which, as mentioned, is assumed to be fully capable of transporting the surplus green H2 after exhausting the capacity of both pipelines and railways. Furthermore, an increase in average speed is considered for the approach followed for railway transportation. For the case of road transportation, a linear trend was assumed, starting at an average speed of 75 km/h in 2020 to 200 km/ h in 2060. It's essential to note that this projection is subject to significant uncertainty, contingent upon the development and widespread implementation of smart vehicles equipped with autonomous driving capabilities. The main transportation routes established for this study are presented in Fig. 11.19. Inner Mongolia will prioritise transportation to the Jing-Jin-Ji area. The Xinjiang and Gansu provinces are expected to first transport green H2 to the Yangtze River Delta, with any surplus being redirected to the Pearl River Delta. Lastly, Tibet

278

J. Godinho et al.

Fig. 11.19 Projected transportation routes between production and consumption regions

is projected to transport exclusively to the Pearl River Delta region. These routes were selected based either on proximity, as is the case for the course Inner Mongolia – Jing-Jin-Ji, or on the expected volume for green H2 production, with the routes Xinjiang – Yangtze River Delta and Xinjiang – Pearl River Delta as examples. This latter assumption was based on the projected cost reductions for both the production of H2 and the respective conversion costs, i.e., liquefaction of H2 or LOHC hydrogenation, by applying the concept of economies of scale. This topic is further explained in Sect. 11.5. Furthermore, it is assumed that, for both the LOW and the HIGH scenarios, once the TPES of each of the consumption regions is entirely CO2-eq emission-free, the surplus of green H2 that could potentially be produced within the respective production province is transported to the other consumption regions, prioritizing distance. This approach ensures that transportation costs are minimised since these costs increase linearly concerning the length travelled, regardless of the transportation mode. This is further explained in Sect. 11.4.2.

11.4.2

Green H2 Cost-Price

For this study, the calculation of the total cost-price for green H2 was composed of three different expenses, with each of them also following a LOW and a HIGH scenario, being directly related to the overall production capacity of the production

11

Supply Chains of Green Hydrogen Based on Liquid Organic Carriers. . .

279

Table 11.9 Learning rates, progress ratios, and respective experience indexes for the production cost of green H2 between 2025 and 2060, for both the LOW and HIGH scenarios Low High

Learning rate 2.0% 4.0%

Progress ratio 98.0% 96.0%

Experience index b -0.0589 -0.0291

provinces, as expressed in Sect. 11.4.1.2. The first expense relates to the production of green H2. The initial values set for the production cost of green H2 in 2025 were ¥31.5/kg-H2 (€4.50/kg-H2) and ¥24.5/kg-H2 (€3.50/kg-H2) for the LOW and the HIGH scenarios, respectively, which are within the projected range [93]. After 2025, the projected financial scaling benefits were assessed through learning rates and progress ratios [103]. This assumption was based on the relative novelty of the technology. However, since the study covers an extended period, the assumed learning rates were considered conservative, ensuring that most of the projected cost decreases occur in the earlier years of deployment. Eq. 11.11 was used to determine the subsequent production costs until 2060, using the data in Table 11.9 for each considered scenario. CostProdH2 ,year,pp = CostProdH2 ,year‐1,pp ×

CumProdyear,pp b CumProdyear‐1,pp 8year, pp

ð11:11Þ

where Cost ProdH2 = Production cost of green H2 [€/kg-H2] CumProd = Cumulative production of green H2 [kg] b = Experience index Before the produced green H2 is ready for transportation, it needs to be converted, depending on the transportation option chosen. For Option 1, it was necessary to encompass the respective costs for liquefaction and regasification, projected to occur within the production provinces and the destined consumption regions, accordingly. For Option 2, the transported H2 is assumed to be loaded and unloaded into the considered LOHC via hydrogenation and dehydrogenation processes. Data was retrieved regarding these specific costs and the associated transportation and storage costs [37]. Figure 11.20 shows the costs per distance travelled, resulting from the scenarios carried out by the European Commission [37]. In this study, we assume that the costs associated with transport in China follow the same trend as in Europe. Additionally, it is notable that the values here shown were calculated assuming the transportation of 1 Mt/year of green H2 and are an average of the two scenarios considered in the reference. Eqs. 11.12, 11.13, and 11.14 referred to the figures’ values and were used to compute the base costs of transporting H2 through either the gas pipeline or multi-modal transportation for both LH2 and LOHCs.

280

J. Godinho et al.

Fig. 11.20 H2 transportation costs relative to distance travelled [37]

Where CostPipe,pp = 0:000233 × AVERAGE DistPipe,pp þ 0:055769

ð11:12Þ

CostLH2 ,pp = 3:62 × 10 - 5 × AVERAGE DistRail&Road,pp þ 0:655232

ð11:13Þ

CostLOHC,pp = 3:90 × 10 - 5 × AVERAGE DistRail&Road,pp þ 0:662021 8pp

ð11:14Þ

Cost Pipe = Transportation cost when using gas pipelines [€/kg-H2] CostLH2 = Transportation cost when using multi-modal transportation for LH2 [€/ kg-H2] CostLOHC = Transportation cost when using multi-modal transportation for LOHCs [€/kg-H2] Dist = Total distance transported [km] However, to extrapolate these costs appropriately, it was necessary to discriminate between CAPEX and OPEX. This is relevant since OPEX was assumed to surge linearly with the production volume, while CAPEX is expected to benefit from economies of scale by applying a scale factor. The respective operational expenses include the transportation and storage costs and the OPEX of the respective conversion processes. The capital expenses are represented by the CAPEX required for the conversion processes. The individual shares of CAPEX and OPEX per transportation option are expressed in Table 11.10. With this data, it was possible to determine the total projected costs for transportation, storage, and the respective conversion processes for each scenario, consumption region, and options for transportation until 2060. For this purpose, Eqs. 11.15, 11.16, and 11.17 were used.

Supply Chains of Green Hydrogen Based on Liquid Organic Carriers. . .

11

281

Table 11.10 Relative shares for CAPEX and OPEX for each considered transportation option [103] CAPEX 69% 84%

Option 1 Option 2

OPEX 31% 16%

For Option 1: CostTSCP,year,pp,cr = ShareOPEX × CostPipe,pp × TranspPipe,pp þ CostLH2 ,pp × TranspLH2 ,pp þ ShareCAPEX × CostPipe,pp × TranspPipe,pp þ CostLH2 ,pp × TranspLH2 ,pp × SFyear,pp TRANSPH2 ,year,cr ð11:15Þ

For Option 2: CostTSCP,year,pp,cr = ShareOPEX × CostLOHC,pp × TranspLOHC,pp þ

ð11:16Þ

ShareCAPEX × CostLOHC,pp × TranspLOHC,pp × SFyear,pp TRANSPH2 ,year,cr and

SFyear,pp =

TRANSPH2 ,year,pp R 1 Mt

TRANSPH2 ,year,pp

× 100%

ð11:17Þ

8year, pp, cr where CostTSCP = Costs for transportation, storage, and the respective conversion processes [€/kg-H2] SF = Scale factor [%] R = Constant capital scale factor [2/3] [103] Hence, to calculate the projected cost-price for green H2, Eq. 11.18 was used. The production cost for green H2 within each consumption region is computed through the mean average of the production costs from each of the incoming transporting

282

J. Godinho et al.

provinces. Finally, it is essential to mention that all calculations were performed in euros. However, the respective results in this study account for a standard conversion rate of 7 Yuan/Euro (¥/€) [104]. CostH2 ,year,pp,cr =

CostProdH2 ,year,pp ×

TRANSPH2 ,year,pp TRANSPH2 ,year,cr

þ CostTSCP,year,cr 8year, pp, cr ð11:18Þ

11.4.3

CO2-eq Emission Savings

The projected savings for CO2-eq emissions generated by the considered TFC of each consumption region were assessed based on a Business-as-usual scenario (BAU). For this scenario only, it was assumed that the energy mix allocated to consumption for 2020 for each consumption region was fixed until 2060. Thus, following the chronological profiles for the computed TFC computed, Sect. 11.4.1.2, the BAU scenario assumed that heat and fossil-based electricity would be generated through CHP plants, with the mix of coal and natural gas presented in Table 11.11 [81]. Transportation was assumed to be fully powered by gasoline and diesel. Regarding the respective emission factors, this study considered that coal emits 352.8 kgCO2-eq/MWh (LHV) and natural gas emits 201.6 kgCO2-eq/MWh (LHV). An average of the respective emission factors for gasoline and diesel was used for transportation, resulting in 257.4 kgCO2-eq/MWh (LHV) [103]. When accounting for all these variables, Eq. 11.19 was used to compute the projected CO2-eq emissions for each of the considered scenarios across all consumption regions and the entire timespan: CO2 Emyear, cr = EnergyFF, year, cr × TFCH:E, year, cr × Mixcoal, year, cr × EFcoal þ MixNG, year, cr × EFNG þTFCT, year, cr × EFGas 8year, cr ð11:19Þ

Table 11.11 Relative shares of used coal and natural gas for heat and electricity production within the consumption regions, 2020 Coal NG

Jing-Jin-Ji region 82.2% 17.8%

Yangtze River Delta 86.1% 13.9%

Pearl River Delta 75.6% 24.4%

11

Supply Chains of Green Hydrogen Based on Liquid Organic Carriers. . .

283

where EnergyFF = Energy content of fossil fuels used [MWh] EF = Emission factor [kgCO2-eq/MWh] H = Heat E = Electricity T = Transportation

11.5

Results and Discussion

The results obtained from the proposed methodology show the projections for each of the considered consumption regions, mentioned in Sect. 4, for both assumed scenarios, LOW and HIGH, between 2025 and 2060. The first set of results highlights the projected cost-price for green H2. Secondly, an overview of the projected values for the TPES is provided, followed by the respective values for total final consumption. Section 11.5.3 shows results regarding the projected savings for CO2-eq emissions when compared to the BAU scenario.

11.5.1

Final Cost-Price of H2 Per Region

Starting with the projected cost-price for the delivered green H2, Figs. 11.21, 11.22, and 11.23 provide a detailed breakdown for each of the consumption regions in both 2040 and 2060. The values for 2040 were chosen since this was the earliest year in which green H2 is projected to be transported to all consumption regions . Furthermore, the values presented are an average of the two transportation options assumed for this study since the results for each are relatively close. This occurs partly 27.0 24.66

18.0 14.50

15.0 12.0 9.0

14.0 12.0 10.0 7.93

8.0 6.0

6.0

4.0

3.0

2.0

0.0

Transportation Conversion processes H2 production

16.83

16.0

H2 total costs (¥/kg-H2)

21.0

H2 total costs (¥/kg-H2)

18.0

Transportation Conversion processes H2 production

24.0

0.0 Low

High Scenario

Low

High Scenario

Fig. 11.21 Projected breakdown of the cost-price for green H2 transported to the Jing-Jin-Ji region. Values for 2040 (left) and 2060 (right) for the LOW and HIGH scenarios

284

J. Godinho et al. 20.0

27.0 24.86

Transportation Conversion processes H2 production

H2 total costs (¥/kg-H2)

21.0 18.0

14.88

15.0 12.0 9.0

18.0

Transportation Conversion processes H2 production

17.52

16.0

H2 total costs (¥/kg-H2)

24.0

14.0 12.0 10.0 7.63

8.0 6.0

6.0

4.0

3.0

2.0 0.0

0.0 Low

Low

High

High Scenario

Scenario

Fig. 11.22 Projected breakdown of the cost-price for green H2 transported to the Yangtze River Delta region. Values for 2040 (left) and 2060 (right) for the LOW and HIGH scenarios

20.0 Transportation Conversion processes H2 production

28.12

27.0

H2 total costs (¥/kg-H2)

24.0 21.0 18.0

15.26

15.0 12.0 9.0

18.0

Transportation Conversion processes H2 production

17.37

16.0

H2 total costs (¥/kg-H2)

30.0

14.0 12.0 10.0 7.72

8.0 6.0

6.0

4.0

3.0

2.0 0.0

0.0 Low

High Scenario

Low

High Scenario

Fig. 11.23 Projected breakdown of the cost-price for green H2 transported to the Pearl River Delta region. Values for 2040 (left) and 2060 (right) for the LOW and HIGH scenarios

because the assumed cost functions for transportation were multi-modal, which does not account for the substantial variation in transportation requirements. An example of this can be noted in the case of Option 2, where road transportation is much more prevalent than in Option 1. Nonetheless, it is notable that Option 2 was the cheapest alternative across all consumption regions and for both the LOW and HIGH scenarios. More specifically, it was calculated that, for the entire timespan, the cost-price for the established supply chain would be roughly 1.8% and 3.7% cheaper when using Option 2 for the LOW and HIGH scenarios, respectively. For 2040, the region of Jing-Jin-Ji is projected to have the cheapest available green H2 in both scenarios, with roughly ¥24.66/kg-H2 (€13.52/kg-H2) in the LOW and ¥14.50/kg-H2 (€2.07/kg-H2) for the HIGH. This can be explained by the relatively low production costs for green H2 within Inner Mongolia for 2040, which according to the results can supply the Jing-Jin-Ji cluster entirely. Contrarily, the cluster of the Pearl River Delta is expected to have the most expensive green H2

11

Supply Chains of Green Hydrogen Based on Liquid Organic Carriers. . .

285

in 2040, at roughly ¥28.12/kg-H2 (€4.02/kg-H2) for the LOW scenario and ¥15.26/ kg-H2 (€2.18/kg-H2) for the HIGH scenario. Again, this can be credited to the high production costs of the supplying provinces. For the LOW scenario, this region is projected to be only supplied by the Tibet Autonomous Region, where the production costs are highest. In the HIGH scenario, this region was projected to receive H2 from Tibet and surplus from Inner Mongolia. In 2060 and for the LOW scenario, the region of Jing-Jin-Ji is also projected to have the lowest cost per kg-H2, at roughly ¥16.83/kg-H2 (€2.40/kg-H2) in 2060, followed by the Pearl River Delta region at ¥17.37/kg-H2 (€2.48/kg-H2) and finally the Yangtze River Delta region at ¥17.52/ kg-H2 (€2.50/kg-H2). However, when looking at the values for the HIGH scenario, the order is reversed, ranging between ¥7.63/kg-H2 (€1.09/kg-H2) for the Yangtze River Delta region to ¥7.93/kg-H2 (€1.13/kg-H2)for the Jing-Jin-Ji cluster. This is primarily explained since the generation costs of green H2 and the respective conversion processes are projected to be substantially lower for the HIGH scenario. Moreover, for 2060 and this study, the province of Xinjiang is projected to be able to fully supply both the regions of the Yangtze and the Pearl River Deltas. This is noteworthy, considering that Xinjiang’s results show the lowest costs for both H2 production and the conversion process costs for transportation. When analysing the individual costs, it is noticeable that green H2 production represents most of the total costs, contributing, on average, between 66% and 87% for the HIGH and LOW scenarios, respectively. Both transportation costs and the associated conversion process costs are expected to be relatively similar, with an average share of 5–8% and 7% for the LOW scenario and of 9–18% and 8–16% for the HIGH scenario, respectively. Even though the costs for the conversion processes were considered to represent most of the supply chain costs, i.e., transportation, storage, and conversion, it is possible to assess that these costs are projected to benefit slightly from scaling. Notwithstanding, further investigation could provide better insight regarding the impact of each assessed cost by performing a sensitivity analysis on the relevant parameters. Apart from this, although the results presented in this study fall within the projected ranges to produce green H2 [93], it is essential to emphasize that more location-specific data would allow for a better assessment of the final cost. Similarly, as mentioned in Sect. 11.4, the supply chain costs were adapted from a study performed for the European Union. A more detailed set of data points could enable a more accurate estimate.

11.5.2

Total Primary Energy Supply Per Region

The projected values for TPES per region are shown in Figs. 11.24, 11.25, 11.26, 11.27, 11.28, and 11.29. For the Jing-Jin-Ji cluster, it is noteworthy that fossil fuels are expected to be entirely displaced by 2040 for the LOW scenario. In contrast, this is anticipated to occur for the HIGH setting by 2035. The electricity transmission through UHVDC is expected to contribute to the displacement of fossil fuels in the

Total primary energy supply (TWh)

286

J. Godinho et al.

1,600 1,400 1,200 1,000 800 600 400 200 2025

Hydrogen

2030

2035

Renewable electricity

2040

2045

2050

Electricity transmitted

Gasoline/Diesel

2055

Natural Gas

2060

Coal

Total primary energy supply (TWh)

Fig. 11.24 Total primary energy supply for the Jing-Jin-Ji region, LOW scenario 1,600 1,400 1,200 1,000 800 600 400 200 2025 Hydrogen

2030

2035

Renewable electricity

2040

2045

Electricity transmitted

2050

Gasoline/Diesel

2055 Natural Gas

2060 Coal

Total Primary Energy Supply (TWh)

Fig. 11.25 Total primary energy supply for the Jing-Jin-Ji region, HIGH scenario 4,000 3,500 3,000 2,500

2,000 1,500 1,000 500 2025

Hydrogen

2030

2035

Renewable electricity

2040

2045

Electricity transmitted

2050

Gasoline/Diesel

2055

Natural Gas

Fig. 11.26 Total primary energy supply for the Yangtze River Delta region, LOW scenario

2060

Coal

Total primary energy supply (TWh)

11

Supply Chains of Green Hydrogen Based on Liquid Organic Carriers. . .

287

4,000 3,500 3,000 2,500

2,000 1,500 1,000 500 2025 Hydrogen

2030

2035

Renewable electricity

2040

2045

Electricity transmitted

2050 Gasoline/Diesel

2055

2060

Natural Gas

Coal

Total primary energy supply (TWh)

Fig. 11.27 Total primary energy supply for the Yangtze River Delta region, HIGH scenario 1,600 1,400 1,200 1,000 800 600 400 200 2025 Hydrogen

2030

2035

Renewable electricity

2040

2045

Electricity transmitted

2050 Gasoline/Diesel

2055 Natural Gas

2060 Coal

Total primary energy supply (TWh)

Fig. 11.28 Total primary energy supply for the Pearl River Delta region, LOW scenario 1,600 1,400

1,200 1,000 800 600 400 200 2025 Hydrogen

2030

2035

Renewable electricity

2040

2045

Electricity transmitted

2050 Gasoline/Diesel

2055 Natural Gas

Fig. 11.29 Total primary energy supply for the Pearl River Delta region, HIGH scenario

2060 Coal

288

J. Godinho et al.

LOW scenario. The explanation for the phenomena mentioned above is that the levels of available renewable electricity produced within the Jing-Jin-Ji region are expected to be lower. Transportation via Option 1 is performed mainly by rail, with a small contribution from gas pipelines, for both scenarios. However, due to the cargo constraints of the assumed railway capacity, transportation via Option 2 requires roughly 80% of road transportation across the considered modes. Once fossil fuels are entirely displaced, green H2 is expected to supply approximately 58% of the TPES, while locally-produced renewable electricity will provide the remaining 42%. The results in the Yangtze River Delta show that fossil fuels are expected to be displaced by 2047 and 2040 for the LOW and HIGH scenarios, correspondingly. Furthermore, contrarily to the Jing-Jin-Ji region, electricity transmission is expected to play a substantially more significant role in both scenarios. In fact, for the LOW setting, it is projected that there will be enough demand for transmitted electricity until 2060, even though its contribution is relatively low due to the constraints in transmission capacity. From 2047 onwards, green H2 is projected to supply between 51% and 58% of the TPES, while the remaining comprises the electricity share fulfilled through transmission or local RES. For the HIGH scenario, green H2 is expected to satisfy 51% of the TPES after 2039. Until 2043, electricity transmission will still provide a minor contribution, although locally renewable electricity is projected to supply the remaining 49% of TPES from 2043 to 2060. In the Pearl River Delta, the projected profile for TPES is considerably different compared to the other clusters analysed. Firstly, it is noteworthy that there is no UHVDC capacity to supply this region with renewable electricity produced in either of the assessed production provinces. Furthermore, for the LOW scenario, it is expected that there will be no locally produced renewable electricity. Conversely, while for the HIGH scenario, it is predicted that there will be a renewable electricity supply, this will nevertheless provide a small contribution towards the displacement of fossil fuels. Overall, the results show the total removal of fossil fuels by 2049 and 2040 for the LOW and HIGH scenarios, respectively. Figure 11.30 provides insight into the amount of green H2 each consumption region is projected to import from the respective provinces. Considering the amounts imported, these values show the amount of green H2 each province is expected to supply between 2025 and 2060. The Jing-Jin-Ji region is projected to be fully supplied by the green H2 produced in Inner Mongolia for both the LOW and HIGH scenarios. Furthermore, it is relevant to highlight that both Gansu and Xinjiang are projected to provide a much higher contribution to the green H2 market within China for the HIGH scenario, which is related to the projections for RES capacity in these provinces.

11.5.3

Total Final Energy Consumption Per Region

The results for the TFC are represented from Figs. 11.31 to 11.36; the figures highlight the projected consumption shares for electricity, heat, and transportation. Furthermore, a distinction is provided between the amount of energy fulfilled

11

Supply Chains of Green Hydrogen Based on Liquid Organic Carriers. . .

Low

High

100%

100%

289

Jing-Jin-Ji

6%

Yangtze River Delta

17%

28%

21% 62%

66%

17%

29%

24%

Pearl River Delta 29%

18%

83%

Gansu

Inner Mongolia

Tibet

Xinjiang

Total primary energy supply (TWh)

Fig. 11.30 Breakdown of the supplied green H2 by province for each consumption region. Values averaged for 2025–2060

1,200 1,000 800 600 400 200 2025

2030

2035

2040

2045

2050

2055

Transportation - ZE

Heat - ZE

Electricity - ZE

Transportation - FF

Heat - FF

Electricity - FF

Fig. 11.31 Total final consumption for the Jing-Jin-Ji region, LOW scenario

2060

Total final energy consumption (TWh)

290

J. Godinho et al. 1,200 1,000

800 600 400 200 2025

2030

Transportation - ZE

2035

Heat - ZE

2040

Electricity - ZE

2045

2050

Transportation - FF

2055

Heat - FF

2060

Electricity - FF

Total final energy consumption (TWh)

Fig. 11.32 Total final consumption for the Jing-Jin-Ji region, HIGH scenario

2,500 2,000 1,500 1,000 500 2025

2030

Transportation - ZE

2035

Heat - ZE

2040

Electricity - ZE

2045

2050

2055

Transportation - FF Heat - FF

2060

Electricity - FF

Fig. 11.33 Total final consumption for the Yangtze River Delta region, LOW scenario

through fossil fuels (FF) or renewable energy, i.e., electricity produced via RES and green H2 with zero emissions (ZE). As detailed, the region of Jing-Jin-Ji (Figs. 11.31 and 11.32) primarily depends on coal and natural gas for its heat demand, which alludes to the heavy industries in this region, as highlighted in Sect. 11.2. Most of the electricity demand is supplied through local power plants or UHVDC transmission from the Inner Mongolia Autonomous Region. The delivered green H2 is projected to supply roughly 25% of the TFC for this region between 2025 and 2060 for the LOW scenario. This contribution is expected to be around 31% of the TFC for the HIGH scenario. Conversely, following the TFC for the Yangtze River Delta region (Figs. 11.33 and 11.34), electricity consumption demands the most significant share of fossil fuels, amounting to around 49% of the total TFC in 2025. Nonetheless, for both

Total final energy consumption (TWh)

11

Supply Chains of Green Hydrogen Based on Liquid Organic Carriers. . .

291

2,500 2,000 1,500 1,000 500

2025

2030

2035

2040

2045

2050

2055

Transportation - ZE

Heat - ZE

Electricity - ZE

Transportation - FF

Heat - FF

Electricity - FF

2060

Total final energy consumption (TWh)

Fig. 11.34 Total final consumption for the Yangtze River Delta region, HIGH scenario 700 600 500 400

300 200 100 2025

2030

Transportation - ZE

2035

Heat - ZE

2040

Electricity - ZE

2045

2050

Transportation - FF

2055

Heat - FF

2060

Electricity - FF

Fig. 11.35 Total final consumption for the Pearl River Delta region, LOW scenario

scenarios, it can be stated that electricity produced via fossil fuels will be mitigated at a larger pace than heat and transportation due to the increasing capacity of RES within the region and the UHVDC transmission originating from the provinces of Gansu and Inner Mongolia. Electricity consumption represents the majority of the TFC, at around 68% between 2025 and 2060, mainly due to an abundant service industry and household lighting [60–62] within this region, as mentioned in section 11.3. Thus, green H2, transported mainly from Inner Mongolia and Xinjiang provinces, is projected only to fulfil around 14% and 20% of the TFC for this region for the LOW and HIGH scenarios, respectively. This H2 is expected to fully supply the demand for both heat and the transportation sector.

Total final energy consumption (TWh)

292

J. Godinho et al. 700 600 500 400 300 200 100 2025

2030

2035

2040

2045

2050

2055

Transportation - ZE

Heat - ZE

Electricity - ZE

Transportation - FF

Heat - FF

Electricity - FF

2060

Fig. 11.36 Total final consumption for the Pearl River Delta region, HIGH scenario

Lastly, regarding the region of the Pearl River Delta (Figs. 11.35 and 11.36), the results show that electricity consumption is the most prevalent among the three consumption regions considered for this study, at nearly 70% of the TFC, for both scenarios. The transported H2 was shown to supply roughly 33% of the TFC for the LOW scenario and around 46% for the HIGH scenario. It is noteworthy that the disparity of the Pearl River Delta's results between scenarios is the largest across all consumption regions, which can be explained through two main reasons. Firstly, no UHVDC transmission is assumed to be available for the entire timespan for both scenarios. Also, renewable electricity produced locally is only available for the HIGH scenario and is projected to contribute roughly 39% of all electricity consumption between 2025 and 2060. As seen from the results of the TPES, this region is expected to be the most dependent on imported green H2 from the production provinces. The second reason for the disparity between scenarios is that, since the distance was prioritized for transportation, this region was subject to the production level of each of the respective production provinces and their surplus of green H2 after supplying the remaining regions. Notwithstanding, for both scenarios, the Pearl River Delta region is projected to receive green H2 from Tibet starting in 2025. However, this contribution is relatively small for the LOW scenario, representing around 17% of the TFC, with Inner Mongolia supplying the remaining 83%. For the HIGH scenario, the region is projected to import H2 from all the production provinces, with Gansu and Tibet providing the majority, at 29% and 28%, respectively.

11

Supply Chains of Green Hydrogen Based on Liquid Organic Carriers. . .

293

700 600

Total CO2-eq emissions (Mt)

500 400 300 200 100 2025

2030

2035

2040 BAU

LOW

2045

2050

2055

2060

2055

2060

HIGH

Fig. 11.37 Total CO2-eq emissions for the Jing-Jin-Ji region, per scenario

Total CO2-eq emissions (Mt)

1,400 1,200 1,000

800 600 400 200 2025

2030

2035 BAU

2040 LOW

2045

2050

HIGH

Fig. 11.38 Total CO2-eq emissions for the Yangtze River Delta region, per scenario

11.5.4

CO2-eq Emission Savings Per Region

The final set of results concerns the potential savings in CO2-eq emissions when choosing either the LOW or the HIGH scenario. Figures 11.37, 11.38, and 11.39 illustrate the contrast between these scenarios and the BAU projection mentioned in Sect. 11.4.3. Starting with the Jing-Jin-Ji region, it is clear that locally produced renewable electricity is prone to benefit the region substantially towards mitigating related CO2-eq emissions, amounting to nearly 70% of the total electricity consumption in 2025. Nevertheless, the delivered green H2 is projected to help the region become CO2-eq emission-free by reducing emissions from around 300 Mt of CO2-eq in 2025 to zero by 2039, the latest, i.e., for the LOW scenario.

294

J. Godinho et al.

Total CO2-eq emissions (Mt)

450 400 350 300 250 200 150 100 50 2025

2030

2035

2040 BAU

LOW

2045

2050

2055

2060

HIGH

Fig. 11.39 Total CO2-eq emissions for the Pearl River Delta region, per scenario

For the Yangtze River Delta region, the abatement of CO2-eq emissions through the incoming green H2 is more demanding, with a targeted level of roughly 1,000 Mt of CO2-eq for decarbonisation, in 2025. Overall, the LOW scenario is expected to prevent the emission of more than 22.8 Gt of CO2-eq between 2025 and 2060. The HIGH scenario is projected to prevent around 28.9 Gt of CO2-eq from being emitted throughout the same period, anticipating the complete carbon abatement by 7 years. Concerning the Pearl River Delta region, mitigating CO2-eq emissions presents the toughest challenge among the considered consumption regions. As seen from the results shown in Sects. 11.5.2 and 11.5.3, the contrast between scenarios is the most striking for this region, with the LOW scenario presenting a relatively low contribution from transported H2 until around 2048. For this region, the LOW scenario can only abate roughly 30% of lifetime emissions or approximately 4.5 Gt of CO2-eq between 2025 and 2060. Contrarily, the HIGH setting is projected to nearly double this contribution, at 8.6 Gt of CO2-eq emissions mitigated between the same time, or around 56% of lifetime emissions.

11.6

Conclusion

This study presents a novel methodology to characterize the environmental and economic impacts of producing, storing, and transporting H2 over long distances. The H2 is assumed to be produced by water electrolysis via electricity generated by solar PV, wind, and hydropower plants. The green H2 is considered to be transported and stored either through a mix of gaseous and liquefied H2 or within two-way LOHCs, in particular the DBT-PDBT system. To corroborate the advantages of the method developed, the framework is implemented in the context of China. This

11

Supply Chains of Green Hydrogen Based on Liquid Organic Carriers. . .

295

choice is driven by several factors: the nation’s substantial CO2-eq emissions, its rapid growth in the hydrogen and renewable power sectors, the extensive geographical distances, and inherent constraints, as well as the renewable resource imbalances prevalent between its coastal and inland regions. Attending to the country’s commitment to reduce its socioeconomic and energy generation regional disparity, the framework considers green H2 production in Inner Mongolia, Tibet, Gansu, and Xinjiang, and its consumption in the Jing-Jin-Ji cluster, and the Pearl and Yangtze River Deltas, between the years 2025 and 2060. Furthermore, following data regarding the projections for wind and solar PV capacity deployment, two separate scenarios were considered for this study, respectively assuming a LOW and a HIGH level of future capacity. The results show that developing green H2 production and supply chains is a viable way to reduce China’s CO2-eq emissions, reducing emissions ranging between 52% for the LOW scenario and 67% for the HIGH scenario in the regions studied. Moreover, the model concludes that, for China, the most financially viable option is the transportation of green H2 via the DBT-PDBT system. More specifically, the LOCH option is projected to be roughly 2% cheaper than transporting pure green H2, either in its gaseous state or by converting the H2 through liquefaction, for the LOW scenario and 4% cheaper for the HIGH scenario for the considered timeframe. This study could provide governmental and regional planning departments with a cost-effective strategy to reduce emissions while enhancing the renewable power potential in China’s most remote areas. For future studies, the following directions are vital: an in-depth analysis of the generation of H2 with systems integration of specific energy sources, such as wind offshore; it would also be vital to investigate the effect of water scarcity, mainly in the northern region, in the process of H2 production. Summary The Chinese government’s commitment to carbon neutrality by 2060 has emphasized hydrogen’s pivotal role in energy transition. This study investigates the potential of electrolysis to provide hydrogen, sourced from remote Chinese provinces. It analyzes the economic and environmental impacts of transporting green hydrogen to major industrial centres. Liquid organic hydrogen carriers facilitate cross-regional hydrogen storage and transportation. Data is sourced from literature, expert interviews, and databases. Results suggest feasible green hydrogen contribution towards China’s carbon neutrality targets.

References 1. Jewell, J., Vinichenko, V., Nacke, L., & Cherp, A. (2019). Prospects for powering past coal. Nature Climate Change, 9, 592–597. https://doi.org/10.1038/s41558-019-0509-6 2. Abel, G. J., Brottrager, M., Crespo Cuaresma, J., & Muttarak, R. (2019). Climate, conflict and forced migration. Global Environmental Change, 54, 239–249. https://doi.org/10.1016/J. GLOENVCHA.2018.12.003 3. UNFCC. (2016). Report of the conference of the parties on its twenty-first session, Paris, France. UNFCC. FCCC/CP/2015/10/Add.1.

296

J. Godinho et al.

4. Zeng, N., Jiang, K., Han, P., Hausfather, Z., Cao, J., Kirk-Davidoff, D., et al. (2022). The Chinese carbon-neutral goal: Challenges and prospects. Advances in Atmospheric Sciences, 1. https://doi.org/10.1007/s00376-021-1313-6 5. Croatian Presidency of the Council of the European Union. (2020). Long-term low greenhouse gas emission development strategy of the European Union and its Member States. Zagreb. 6. United States of America Government. (2023). The United States of America nationally determined contribution reducing greenhouse gases in the United States: A 2030 emissions target. 7. International Energy Agency. (2023). China – Countries & Regions – IEA. https://www.iea. org/countries/China. Accessed 4 June 2022. 8. Nogrady, B. (2021). China launches world’s largest carbon market: But is it ambitious enough? Nature, 595, 637. https://doi.org/10.1038/d41586-021-01989-7 9. Chinese State Council. (2021). The fourteenth five year plan for national economic and social development of the people’s Republic of China and the outline of long-term objectives for 2035. http://www.gov.cn/xinwen/2021-03/13/content_5592681.htm. Accessed 4 June 2022. 10. Venditti, B. (2022). Visualizing China’s evolving energy mix. https://elements.visualcapitalist. com/visualizing-chinas-energy-transition/#:~:text=Underthenewplan%2Cby2060%2C Chinais,enoughsolarpowerfor33millionhomes%28110GW%29. Accessed 22 Aug 2022. 11. Liu, W., Wan, Y., Xiong, Y., & Gao, P. (2022). Green hydrogen standard in China: Standard and evaluation of low-carbon hydrogen, clean hydrogen, and renewable hydrogen. International Journal of Hydrogen Energy. https://doi.org/10.1016/j.ijhydene.2021.10.193 12. Nakano, J. (2020). China’s hydrogen industrial strategy. Center for Strategic and International Studies. https://www.csis.org/analysis/chinas-hydrogen-industrial-strategy#:~:text= Notwithstandingtheabsenceofanationalhydrogenstrategy%2C,developmentplans%2 Cincludingintheirownfive-yearplans. Accessed 4 June 2022. 13. Yu, Y. (2020). China hydrogen policy: A summary of provincial plans – Energy iceberg. Energy Iceberg. https://energyiceberg.com/china-hydrogen-policy-provincial-summary/. Accessed 4 June 2022. 14. Zheng, X. (2021). Beijing sews up hydrogen energy plan. China Daily. https://www. chinadaily.com.cn/a/202108/18/WS611c5d3da310efa1bd669724.html. Accessed 4 June 2022. 15. National Development and Reform Commission. (2022). 14th five-year “new energy storage” execution plan. https://mp.weixin.qq.com/s/o4XrINzW72zb66fBTysxqQ. Accessed 4 June 2022. 16. China Hydrogen Alliance. (2020). White paper of hydrogen energy and fuel cell industry in China. 17. International Energy Agency. (2019). The future of hydrogen. 18. Song, S., Lin, H., Sherman, P., Yang, X., Nielsen, C. P., Chen, X., et al. (2021). Production of hydrogen from offshore wind in China and cost-competitive supply to Japan. Nature Communications, 12, 1–8. https://doi.org/10.1038/s41467-021-27214-7 19. Zhang, S., & Chen, W. (2022). Assessing the energy transition in China towards carbon neutrality with a probabilistic framework. Nature Communications, 13, 1–15. https://doi.org/ 10.1038/s41467-021-27671-0 20. Fan, J. L., Yu, P., Li, K., Xu, M., & Zhang, X. (2022). A levelized cost of hydrogen (LCOH) comparison of coal-to-hydrogen with CCS and water electrolysis powered by renewable energy in China. Energy, 242, 123003. https://doi.org/10.1016/j.energy.2021.123003 21. Yan, X., Zhang, X., Gu, C., & Li, F. (2018). Power to gas: Addressing renewable curtailment by converting to hydrogen. Frontiers in Energy, 12, 560–568. https://doi.org/10.1007/s11708018-0588-5 22. Eypasch, M., Schimpe, M., Kanwar, A., Hartmann, T., Herzog, S., Frank, T., et al. (2017). Model-based techno-economic evaluation of an electricity storage system based on Liquid Organic Hydrogen Carriers. Applied Energy, 185, 320–330. https://doi.org/10.1016/j. apenergy.2016.10.068

11

Supply Chains of Green Hydrogen Based on Liquid Organic Carriers. . .

297

23. Tso, W. W., Demirhan, C. D., Lee, S., Song, H., Powell, J. B., & Pistikopoulos, E. N. (2019). Energy carrier supply chain optimization: A Texas case study. Computer Aided Chemical Engineering, 47, 1–6. https://doi.org/10.1016/B978-0-12-818597-1.50001-1 24. Wang, S., Tarroja, B., Schell, L. S., Shaffer, B., & Samuelsen, S. (2019). Prioritizing among the end uses of excess renewable energy for cost-effective greenhouse gas emission reductions. Applied Energy, 235, 284–298. https://doi.org/10.1016/j.apenergy.2018.10.071 25. Aakko-Saksa, P. T., Cook, C., Kiviaho, J., & Repo, T. (2018). Liquid organic hydrogen carriers for transportation and storing of renewable energy – Review and discussion. Journal of Power Sources, 396, 803–823. https://doi.org/10.1016/j.jpowsour.2018.04.011 26. Dehghani-Sanij, A. R., Tharumalingam, E., Dusseault, M. B., & Fraser, R. (2019). Study of energy storage systems and environmental challenges of batteries. Renewable and Sustainable Energy Reviews, 104, 192–208. https://doi.org/10.1016/j.rser.2019.01.023 27. Lanphen, S. (2019). Hydrogen import terminal. 28. Dawood, F., Anda, M., & Shafiullah, G. M. (2020). Hydrogen production for energy: An overview. International Journal of Hydrogen Energy, 45, 3847–3869. https://doi.org/10.1016/ j.ijhydene.2019.12.059 29. Andersson, J., & Grönkvist, S. (2019). Large-scale storage of hydrogen. International Journal of Hydrogen Energy, 44, 11901–11919. https://doi.org/10.1016/j.ijhydene.2019.03.063 30. Sillar, K., Hofmann, A., & Sauer, J. (2009). Ab initio study of hydrogen adsorption in MOF-5. Journal of the American Chemical Society, 131, 4143–4150. https://doi.org/10.1021/ ja8099079 31. Saha, D., Deng, S., & Yang, Z. (2009). Hydrogen adsorption on metal-organic framework (MOF-5) synthesized by DMF approach. Journal of Porous Materials, 16, 141–149. https:// doi.org/10.1007/s10934-007-9178-3 32. Reuß, M., Grube, T., Robinius, M., Preuster, P., Wasserscheid, P., & Stolten, D. (2017). Seasonal storage and alternative carriers: A flexible hydrogen supply chain model. Applied Energy, 200, 290–302. https://doi.org/10.1016/j.apenergy.2017.05.050 33. Makepeace, J. W., He, T., Weidenthaler, C., Jensen, T. R., Chang, F., Vegge, T., et al. (2019). Reversible ammonia-based and liquid organic hydrogen carriers for high-density hydrogen storage: Recent progress. International Journal of Hydrogen Energy, 44, 7746–7767. https:// doi.org/10.1016/j.ijhydene.2019.01.144 34. Niermann, M., Drünert, S., Kaltschmitt, M., & Bonhoff, K. (2019). Liquid organic hydrogen carriers (LOHCs)-techno-economic analysis of LOHCs in a defined process chain. Energy and Environmental Science, 12, 290–307. https://doi.org/10.1039/c8ee02700e 35. Abdin, Z., Tang, C., Liu, Y., & Catchpole, K. (2021). Large-scale stationary hydrogen storage via liquid organic hydrogen carriers. IScience, 24, 102966. https://doi.org/10.1016/j.isci.2021. 102966 36. Blanco, H. (2022). What’s best for hydrogen transport: Ammonia, liquid hydrogen, LOHC or pipelines? EnergypostEu. 37. European Commission. (2021). Assessment of hydrogen delivery options – Science for policy briefs. 38. Van Hoecke, L., Laffineur, L., Campe, R., Perreault, P., Verbruggen, S. W., & Lenaerts, S. (2021). Challenges in the use of hydrogen for maritime applications. Energy and Environmental Science, 14, 815–843. https://doi.org/10.1039/d0ee01545h 39. Hurskainen, M., & Ihonen, J. (2020). Techno-economic feasibility of road transport of hydrogen using liquid organic hydrogen carriers. International Journal of Hydrogen Energy, 45, 32098–32112. https://doi.org/10.1016/j.ijhydene.2020.08.186 40. Hydroville. (2022). How is hydrogen transported? 41. IRENA. (2022). Global hydrogen trade to meet the 1.5 °C climate goal: Part II – Technology review of hydrogen carriers. 42. Diermann, R. (2020). Hydrogen can be transported by rail, German railway company says. PV Magazine. 43. IMTT. (2011). Instrução de exploração técnica no. 51.

298

J. Godinho et al.

44. Energy.gov. (2022). Liquid hydrogen delivery. 45. Almansoori, A., & Shah, N. (2006). Design and operation of a future hydrogen supply chain: Snapshot model. Chemical Engineering Research and Design, 84, 423–438. https://doi.org/ 10.1205/cherd.05193 46. Miao, B., Giordano, L., & Chan, S. H. (2021). Long-distance renewable hydrogen transmission via cables and pipelines. International Journal of Hydrogen Energy, 46, 18699–18718. https://doi.org/10.1016/j.ijhydene.2021.03.067 47. Centre for European Policy Studies. (2019). The future of gas in Europe: Review of recent studies on the future of gas. 48. Cerniauskas, S., Jose Chavez Junco, A., Grube, T., Robinius, M., & Stolten, D. (2020). Options of natural gas pipeline reassignment for hydrogen: Cost assessment for a Germany case study. International Journal of Hydrogen Energy, 45, 12095–12107. https://doi.org/10. 1016/j.ijhydene.2020.02.121 49. ACER. (2021). Transporting pure hydrogen by repurposing existing gas infrastructure: Overview of existing studies and reflections on the conditions for repurposing. 50. Messaoudani, Z., Rigas, F., Binti Hamid, M. D., & Che Hassan, C. R. (2016). Hazards, safety and knowledge gaps on hydrogen transmission via natural gas grid: A critical review. International Journal of Hydrogen Energy, 41, 17511–17525. https://doi.org/10.1016/j. ijhydene.2016.07.171 51. Gondal, I. A. (2019). Hydrogen integration in power-to-gas networks. International Journal of Hydrogen Energy, 44, 1803–1815. https://doi.org/10.1016/j.ijhydene.2018.11.164 52. Marcogaz. (2019). Technical association of the European gas industry. 53. Planning Department – Shanghai Investigation Design and Research Institute. (2022). Map of People’s Republic of China. 54. Beijing Municipal Bureau of Statistics. (2022). Beijing Municipal Statistical Bulletin on national economic and social development in 2021. http://tjj.beijing.gov.cn/tjsj_31433/ tjgb_31445/ndgb_31446/202203/t20220301_2618672.html. Accessed 13 June 2022. 55. Tianjin Municipal Bureau of Statistics. (2022). Statistical bulletin of Tianjin National Economic and social development in 2021. http://stats.tj.gov.cn/tjsj_52032/tjgb/202203/ t20220314_5828586.html. Accessed 13 June 2022. 56. Hebei Survey Department of the National Bureau of Statistics. (2022). Statistical bulletin of national economic and social development of Hebei Province in 2021. http://www.hetj.gov. cn/cms/preview/hetj/app/tjgb/101642400676359.html. Accessed 13 June 2022. 57. XE. (2022). Chinese Yuan Renminbi to Euros Exchange Rate. https://www.xe.com/ currencyconverter/convert/?Amount=1&From=CNY&To=EUR. Accessed 10 June 2022. 58. Xinhua News. (2019). “Jing-jin-ji”: China’s regional city cluster takes shape. https://global. chinadaily.com.cn/a/201902/25/WS5c73b633a3106c65c34eb4c8.html. Accessed 13 June 2022. 59. Mao, R., Bao, Y., Duan, H., & Liu, G. (2021). Global urban subway development, construction material stocks, and embodied carbon emissions. Humanities and Social Sciences Communications, 8. https://doi.org/10.1057/S41599-021-00757-2 60. Shanghai Municipal Bureau of Statistics. (2022). Shanghai Municipal Statistical Bulletin on national economic and social development in 2021. http://tjj.sh.gov.cn/tjgb/20220314/e0 dcefec098c47a8b345c996081b5c94.html. Accessed 13 June 2022. 61. General Office of Jiangsu Provincial People’s Government. (2022). Statistical bulletin of national economic and social development of Jiangsu Province in 2021. http://www.jiangsu. gov.cn/art/2022/3/31/art_64797_10398993.html. Accessed 13 June 2022. 62. Zhejiang Provincial Bureau of Statistics. (2022). Statistical bulletin of national economic and social development of Zhejiang Province in 2021. http://tjj.zj.gov.cn/art/2022/2/24/art_122 9129205_4883213.html. Accessed 13 June 2022. 63. Rabbit Island Online. (2021). The latest GDP ranking of cities in Anhui Province in 2020: Hefei has exceeded trillion, with one city dominating the other. https://www.163.com/dy/ article/G2E9IV5E0512FSNF.html#:~:text=2020. Accessed 13 June 2022.

11

Supply Chains of Green Hydrogen Based on Liquid Organic Carriers. . .

299

64. The State’s Council of The People’s Republic of China. (2019). China outlines integrated development of Yangtze River Delta. https://english.www.gov.cn/policies/latestreleases/201 912/01/content_WS5de3a5c0c6d0bcf8c4c181e2.html. Accessed 13 June 2022. 65. Baiyu, G. (2020). Offshore wind takes off in China 2020. https://chinadialogue.net/en/energy/ china-offshore-wind-power-growth/#:~:text=Chinanowhas23%25oftheworld’soffshore, Fujianadded0.35GWand0.2gigawattsrespectively. Accessed 13 June 2022. 66. Guangdong Provincial Bureau of Statistics. (2022). Statistical bulletin of national economic and social development of Guangdong Province in 2021. http://stats.gd.gov.cn/attachment/0/4 83/483178/3836135.pdf. Accessed 13 June 2022. 67. Statistics and Census Service. (2022). Gross domestic product – Statistics. https://www.dsec. gov.mo/en-US/Statistic?id=901. Accessed 13 June 2022. 68. Census and Statistics Department TG of the HKSAR. (2022). Statistical products. https:// www.censtatd.gov.hk/en/. Accessed 13 June 2022. 69. The World Bank S. (2020). Global Solar Atlas 2.0. Solar Resource Maps of China, Global Solar Atlas 2.0. https://solargis.com/maps-and-gis-data/download/china Accessed 2 Nov 2022. 70. Thieriot, H. (2016). Wind & Sun: Relief For China’s Dry North – China Water Risk. https:// www.chinawaterrisk.org/resources/analysis-reviews/wind-sun-relief-for-chinas-dry-north/. Accessed 14 June 2022. 71. Statistics Bureau of the Tibet Autonomous Region. (2022). Statistical bulletin on national economic and social development of the Tibet Autonomous Region in 2021. http://www. xizang.gov.cn/zwgk/zfsj/ndtjgb/202204/t20220426_295819.html. Accessed 14 June 2022. 72. National Bureau of Statistics. (2022). Statistical analysis of power generation and power generation structure in Tibet Autonomous Region in 2021. https://www.huaon.com/channel/ distdata/782773.html. Accessed 14 June 2022. 73. Li, M. (2022). Focus on ecological protection and promote green development with high https://www.tibet3.com/news/zangqu/xz/2022-03-07/262868.html. Accessed quality. 14 June 2022. 74. General Office of the People’s Government of Xinjiang Uygur Autonomous Region. (2022). Statistical bulletin on national economic and social development of Xinjiang Uygur Autonomous Region in 2021. http://www.xinjiang.gov.cn/xinjiang/tjgb/202203/2325b36ee55d4c64 9313e764078a17e4.shtml. Accessed 15 June 2022. 75. Zheng, X. (2021). Xinjiang powers up with clean energy efforts. https://global.chinadaily.com. cn/a/202112/08/WS61b01601a310cdd39bc7a0be.html. Accessed 15 June 2022. 76. Gansu Province Bureau of Statistics. (2022). Statistical bulletin of national economic and social development of Gansu Province in 2021. http://tjj.gansu.gov.cn/tjj/c1094 57/202203/2000738.shtml. Accessed 15 June 2022. 77. Statistics Bureau of Inner Mongolia Autonomous Region. (2022). https://www.nmg.gov.cn/ tjsj/sjfb/tjsj/tjgb/202202/t20220228_2010485.html. Accessed 15 June 2022. 78. Zhu, S., Diab, N., Goodrich, J., Lin, D., Wen, Y., Zhang, S., et al. (2020). A story of struggle and success: China’s 832 poorest counties. China Global Television Network. https://news. cgtn.com/event/2020/China-s-battle-against-poverty/index.html#:~:text= Accordingtothecharacteristicsofregionalpoverty%2CChina,southwestern Chinawithfewerclustersincentralprovinces. Accessed 14 June 2022. 79. Energy Research Institute, Natural Development and Commission Reform. (2015). China 2050 high renewable energy penetration scenario and roadmap study. 80. Chen, Y., Guo, F., Wang, J., Cai, W., Wang, C., & Wang, K. (2020). Provincial and gridded population projection for China under shared socioeconomic pathways from 2010 to 2100. Scientific Data, 7, 1–13. https://doi.org/10.1038/s41597-020-0421-y 81. National Bureau of Statistics. (2021). China energy statistics 2021 edition. 82. China Three Gorges Corporation. (2020). Inner Mongolia wind farm planning report. 83. China Three Gorges Corporation. (2020). Solar PV power plants design guideline. 84. Chen, S., Lu, X., Nielsen, C. P., Geng, G., He, K., McElroy, M. B., et al. (2022). Improved air quality in China can enhance solar-power performance and accelerate carbon-neutrality targets. One Earth, 5, 550–562. https://doi.org/10.1016/j.oneear.2022.04.002

300

J. Godinho et al.

85. China Three Gorges Construction Engineering. (2021). Tiantai Pumped Storage Power Station Project Report. 86. Inner Mongolia Autonomous Region Energy Bureau. (2021). Notice on matters related to the development and construction of wind power and photovoltaic power generation in 2021. http://nyj.nmg.gov.cn/zwgk/zfxxgkzl/fdzdgknr/zcjd_16463/202110/t20211013_1902163. html. Accessed 2 Nov 2022. 87. National Renewable Energy Laboratory. (2021). Utility-scale battery storage | Electricity. https://atb.nrel.gov/electricity/2021/utility-scale_battery_storage. Accessed 2 Nov 2022. 88. Shu, Y., & Chen, W. (2018). Research and application of UHV power transmission in China. High Voltage, 3, 1–13. https://doi.org/10.1049/HVE.2018.0003 89. Marc Champion. (2021). Future of world energy lies in UHVDC transmission lines. Bloomberg. https://www.bloomberg.com/news/features/2021-06-09/future-of-world-energylies-in-uhvdc-transmission-lines. Accessed 2 Nov 2022. 90. International Energy Agency. (2011). Technology roadmap – China wind energy development roadmap 2050. https://www.iea.org/reports/technology-roadmap-china-wind-energy-develop ment-roadmap-2050. Accessed 2 Nov 2022. 91. Pan, J., Nuqui, R., Srivastava, K., Jonsson, T., Holmberg, P., & Hafner, Y. J. (2008). AC grid with embedded VSC-HVDC for secure and efficient power delivery. In 2008 IEEE energy 2030 conference, ENERGY 2008. https://doi.org/10.1109/ENERGY.2008.4781042 92. Sadovskaia, K., Bogdanov, D., Honkapuro, S., & Breyer, C. (2019). Power transmission and distribution losses – A model based on available empirical data and future trends for all countries globally. International Journal of Electrical Power & Energy Systems, 107, 98–109. https://doi.org/10.1016/J.IJEPES.2018.11.012 93. IRENA. (2020). Green hydrogen cost reduction. 94. Singh, A. P., & Agarwal, A. K. (2021). Introduction to novel internal combustion engine technologies for performance improvement and emission reduction. Energy, Environment, and Sustainability, 3–6. https://doi.org/10.1007/978-981-16-1582-5_1 95. IRENA (International Renewable Energy Agency). (2022). Global hydrogen trade to meet the 1.5°C climate goal: Technology review of hydrogen carriers. 96. Niermann, M., Timmerberg, S., Drünert, S., & Kaltschmitt, M. (2021). Liquid organic hydrogen carriers and alternatives for international transport of renewable hydrogen. Renewable and Sustainable Energy Reviews, 135, 110171. https://doi.org/10.1016/J.RSER.2020. 110171 97. Open Infrastructure Map. (2022). https://openinframap.org/#2.93/38.26/108.83. Accessed 2 Nov 2022. 98. Shashi Menon, E. (2015). Transmission pipeline calculations and simulations manual. 99. BNSF. (2022). Intermodal transportation. https://www.bnsf.com/about-bnsf/virtual-traintour/flatbed.html. Accessed 3 Nov 2022. 100. Quorum Corporation. (2010). Railway capacity background & overview. 101. Niermann, M., Beckendorff, A., Kaltschmitt, M., & Bonhoff, K. (2019). Liquid Organic Hydrogen Carrier (LOHC) – Assessment based on chemical and economic properties. International Journal of Hydrogen Energy, 44, 6631–6654. https://doi.org/10.1016/j.ijhydene. 2019.01.199 102. van Leijen, M. (2021). China presents freight train to break speed record with 350 kmph. RailFreight. https://www.railfreight.com/railfreight/2021/01/05/new-claim-for-fastest-freighttrain-in-the-world-350-kmph/?gdpr=accept&gdpr=deny. Accessed 3 Nov 2022. 103. Kornelis, B., & Evert, N. (2020). Introduction to energy analysis (3rd ed.). Routledge. https:// doi.org/10.1038/001647a0 104. European Central Bank. (2022). Chinese yuan renminbi (CNY). https://www.ecb.europa.eu/ stats/policy_and_exchange_rates/euro_reference_exchange_rates/html/eurofxref-graph-cny. en.html. Accessed 3 Nov 2022.

Chapter 12

Green Hydrogen Research and Development Projects in the European Union Hossein Biabani , Ali Aminlou , Mohammad Mohsen Hayati Hassan Majidi-Gharehnaz , and Mehdi Abapour

12.1

,

Introduction

The utilization of renewable electricity to generate hydrogen through the process of water electrolysis is the backbone of the Power-to-Gas concept. This hydrogen can be directly employed as the ultimate energy carrier [1]. In the context of worldwide climate change, the severity of greenhouse gas emissions has escalated [2]. The EU has set ambitious targets to reduce greenhouse gas emissions and achieve carbon neutrality by 2050. To decarbonize the energy sector and meet these goals, the EU is investing in Horizon projects focused on green hydrogen. Green hydrogen, produced from renewable sources, can be used as a clean fuel for transportation, heating, and industry, contributing to the EU’s climate objectives and sustainable energy systems. The Horizon projects related to green hydrogen have significant implications for the future of EU energy systems, enabling the integration of renewables, reducing emissions, and promoting economic growth. Europe aims to shift from natural gas-derived hydrogen to renewable hydrogen produced through electrolysis with renewable electricity. By 2030, the European Commission proposed producing and importing ten million tons of renewable hydrogen, offering opportunities for decarbonizing industries and transportation and providing long-term storage for the electricity sector, enhancing overall energy efficiency. The EU’s strategy for energy system integration focuses on optimizing and modernizing the entire energy system by connecting different energy carriers and end-use sectors. By linking electricity, heat, cold, gas, and fuels, the EU aims to achieve efficiency and

H. Biabani · A. Aminlou · M. M. Hayati · H. Majidi-Gharehnaz · M. Abapour (✉) Faculty of Electrical and Computer Engineering, Energy Systems Research Institute (ESRI), Smart Energy Systems Lab, University of Tabriz, Tabriz, Iran e-mail: [email protected] © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 V. Vahidinasab et al. (eds.), Green Hydrogen in Power Systems, Green Energy and Technology, https://doi.org/10.1007/978-3-031-52429-5_12

301

302

H. Biabani et al.

decarbonization through cross-sectoral integration, utilizing technologies, digitalization, and flexibility markets. Energy system integration will help achieve cost-effective decarbonization and empower consumers to make energy choices while strengthening cross-sectoral links and reducing emissions [3]. The EU demonstrates its commitment to hydrogen development through the Important Projects of Common European Interest (IPCEIs) program, providing substantial investment support for hydrogen-related projects. IPCEI Hy2Tech and IPCEI Hy2Use focus on developing innovative technologies, constructing hydrogen infrastructure, and integrating hydrogen into the industrial sector. The Clean Hydrogen Partnership further supports research, innovation, and collaboration among stakeholders to drive the development and deployment of clean hydrogen technologies. The EU has also adopted clarifying rules under the Renewable Energy Directive to define renewable hydrogen criteria and calculate life-cycle emissions, fostering the development of renewable hydrogen in the EU’s energy system. The EU’s investment, regulatory frameworks, and collaborative efforts highlight its commitment to hydrogen as a key component of the clean energy transition, promoting innovation and deployment across various sectors [4].

12.2 12.2.1

The Hydrogen Technology Ecosystem Hydrogen as a Sustainable Energy Source

Hydrogen is an abundant element that can be obtained from various sources, including water (through electrolysis) and renewable resources (such as biomass or solar energy) [5]. It is considered a promising energy carrier due to its high energy density and the ability to produce electricity with only water vapor as a byproduct when used in fuel cells [6].

12.2.2

Fuel Cells

Fuel cells are electrochemical systems that, without burning fuel, turn the chemical energy of hydrogen and oxygen (from the air) into electricity. They consist of an anode, a cathode, and an electrolyte. Different types of fuel cells, such as proton exchange membrane (PEM) fuel cells, solid oxide fuel cells (SOFCs), and alkaline fuel cells (AFCs), have unique characteristics and applications [7].

12.2.2.1

Working Principles of Fuel Cells

Hydrogen gas is fed to the anode of a PEM fuel cell, where it is divided by the catalyst into protons (H+) and electrons (e-). The protons move through the proton

12

Green Hydrogen Research and Development Projects in the European Union

303

exchange membrane while the electrons are routed through an external circuit to produce an electric current. On the cathode side, water is the only consequence of oxygen’s combination with protons and electrons [8].

12.2.2.2

Advantages of Hydrogen and Fuel Cells

• Clean and Sustainable: Hydrogen used in fuel cells produces electricity with no emissions of greenhouse gases or air pollutants, contributing to a cleaner and healthier environment [9]. • High Efficiency: Fuel cells have high energy conversion efficiency, typically greater than traditional combustion-based technologies, reducing energy waste [10]. • Versatility: Hydrogen can be produced from diverse sources and used in a wide range of applications, including transportation, stationary power generation, and portable devices [11]. • Energy Storage: Hydrogen can be used as energy storage, allowing excess renewable energy to be stored and used later when demand is high [5, 12].

12.2.3

Challenges and Considerations

• Infrastructure: Developing a robust hydrogen production, storage, and distribution infrastructure is essential for the widespread adoption of hydrogen as an energy carrier. • Cost: Reducing the cost of fuel cells, hydrogen production, and storage technologies is crucial to make them more economically competitive. • Safety: Ensuring safe handling, storage, and transportation of hydrogen is a key consideration for widespread deployment. • Sustainability: The environmental impact of hydrogen production methods should be carefully evaluated, emphasizing utilizing renewable energy sources and minimizing carbon emissions. Continued research, development, innovation, supportive policies, and international collaboration are crucial for unlocking the full potential of hydrogen and fuel cells in the future of energy. By integrating hydrogen and fuel cells into various sectors, we can move toward a sustainable, low-carbon, and decentralized energy system [13].

304

12.3

H. Biabani et al.

Hydrogen Policy of the European Union (EU)

The European Union introduced its hydrogen policy in July 2020 to achieve a carbon-neutral economy. Previous policies from countries like Australia and Japan have also addressed hydrogen, but the EU’s approach stands out for its comprehensive nature and integrated strategy to reduce carbon emissions. This policy, part of the EU’s COVID-19 pandemic recovery efforts, emphasizes the concept of “rebuilding better” to effectively combat climate change. Central to the EU plan, as well as earlier plans, is the utilization of carbon-neutral hydrogen as the primary currency of the new energy economy. Carbon-free energy can be stored in carbon-neutral hydrogen and later exchanged as a commodity between producers and consumers. By producing carbon-neutral hydrogen from diverse renewable energy sources and employing it in various applications such as residential, commercial, industrial, power generation, and transportation, the carbon footprint of these applications can be reduced or eliminated [14]. This approach offers two significant advantages: 1. Carbon-neutral hydrogen serves as an energy source, powering sectors of the economy that are challenging to decarbonize through electrification. This is achieved through fuel cells and the production of synthetic fuels from hydrogen. 2. Carbon-neutral hydrogen production facilitates the storage of intermittent renewable energy, enabling renewable sources to become a mainstream energy option.

12.3.1

European Hydrogen Bank

The European Commission has proposed the establishment of the European Hydrogen Bank, which is part of its hydrogen accelerator measures. This bank aims to promote investment security and business opportunities for renewable hydrogen production in Europe and globally. President Von der Leyen announced this initiative in her State of the Union speech in 2022, and it is included in the Commission’s work program for 2023. The European Hydrogen Bank’s primary objective is to attract private investments in hydrogen value chains by addressing initial investment challenges and connecting renewable energy supply to EU demand. It will create a market for renewable hydrogen, offer growth opportunities and jobs, and support renewable hydrogen production through an auction under the Innovation Fund. Green hydrogen partnerships will also be formed to import renewable hydrogen from third countries and incentivize decarbonization. These efforts aim to ensure fair competition between EU production and imports. The Hydrogen Energy Network is an informal group that assists national energy authorities in leveraging the opportunities presented by hydrogen. It serves as a platform for sharing information, experiences, and developments in hydrogen and enables collaboration on specific issues [15, 16].

12

Green Hydrogen Research and Development Projects in the European Union

12.4

305

Advantages of Hydrogen in Connection with Renewable Energy Sources

Hydrogen is widely acknowledged as a crucial catalyst in the transition toward sustainable energy systems. Its versatility and flexibility make it an ideal energy carrier to address the challenges associated with renewable energy adoption. By enabling sector coupling, hydrogen facilitates the transfer of energy across various sectors, such as electricity, heating, and transportation. It can be produced from renewable sources and utilized in diverse applications, including power generation, transportation, and industrial processes. One significant contribution of hydrogen is its support for integrating renewable energy into the grid. With the increasing share of renewables in the energy mix, there is a need for long-term energy storage solutions to balance supply and demand over different time scales. Unlike shortterm storage options like batteries, hydrogen offers the capability for large-scale and extended-duration energy storage. It can be stored in multiple forms, such as gas, liquid, or other compounds. Additionally, hydrogen plays a vital role in longdistance energy transportation, benefiting regions with abundant renewable resources far from demand centers. It can facilitate renewable energy transportation from one region to another using pipelines, ships, or trucks. This approach is often more cost-effective than constructing new power transmission lines. Moreover, hydrogen supports the development of new renewable energy technologies by storing excess energy from intermittent sources like wind and solar. This stored energy can later power fuel cells or other devices during periods of high demand, reducing the need for backup power from fossil fuels and enabling greater deployment of renewable energy capacity. Fig. 12.1 shows the benefits of hydrogen for the EU based on an ambitious vision up to 2050.

Fig. 12.1 Benefits of hydrogen for the EU according to an ambitious vision up to 2050

306

12.5

H. Biabani et al.

GH2 Projects in the World

Green hydrogen (GH2) is seen as a promising solution to global energy and climate challenges due to its clean and sustainable nature. It can be produced using renewable energy sources like wind, solar, and hydropower, and has versatile applications in transportation, industry, and power generation. One of the key advantages of green hydrogen is its versatility, finding use in various sectors. It can power fuel cell electric vehicles in transportation, emitting only water vapor. In industry, it can replace fossil fuels in processes like steel production, reducing greenhouse gas emissions. Moreover, in power generation, it can provide grid balancing, offering reliable energy when renewables are unavailable. • Australia is at the forefront of green hydrogen production with projects like the Western Green Energy Hub in Western Australia, which aims to generate green hydrogen through a 50 GW wind and solar energy plant. This hydrogen will be used for industrial processes, powering data centers, and export, leading to significant green energy production and emission reduction [17]. • Germany’s Siemens Gas and Power is working on a 100 MW electrolyzer plant project powered by renewable energy. This plant will produce green hydrogen to help balance the country’s electricity grid, marking a significant step toward a more sustainable energy system [18]. • The Netherlands is also investing in green hydrogen production, particularly through developing a green hydrogen hub in the northern part of the country, led by Gasunie and various partners. Renewable energy sources will be utilized to produce hydrogen for applications in transportation and industry, contributing to increased green energy production and reduced emissions [19]. • In Japan, a consortium collaborates on a 10 MW green hydrogen production plant in Fukushima. Renewable energy will be used to generate hydrogen for fuel cells in power generation and transportation, representing a significant stride toward a sustainable energy system [20]. • The United States is likewise involved in green hydrogen production, such as the Los Angeles Department of Water and Power project, which aims to construct a green hydrogen facility at the Haynes Generating Station in California. With a daily hydrogen production capacity of up to 3.6 tons, this initiative can contribute to powering the city’s grid and generating substantial amounts of green energy while reducing greenhouse gas emissions [21]. The Los Angeles Department of Water and Power has committed to operate its fossil-fueled power plants, totaling 4300 MW, with a portion of that power, around 3500 MW within the LA area, sourced from green hydrogen by the middle of the decade. This usage is expected to increase gradually, aiming for 100 percent reliance on green hydrogen within approximately 10 years [22].

12

Green Hydrogen Research and Development Projects in the European Union

12.6

307

Hydrogen Roadmap in Europe

The Hydrogen Roadmap Europe provides a comprehensive and detailed analysis of the potential for hydrogen deployment in the EU, focusing on how hydrogen can enable Europe’s energy transition. The report outlines the unique versatility of hydrogen as an energy carrier and its potential to contribute to a sustainable and low-carbon economy. It also discusses various hydrogen solutions for Europe’s economy and environment, including transportation, industry, and power generation. By realizing this objective, the EU will be on track to cut its CO2 emissions by 560 Mton by 2050. According to the scientific document, the potential for hydrogen production in the EU is approximately 2250 terawatt hours (TWh), which covers almost a quarter of the EU’s total energy demand. The process of transitioning Europe to a decarbonized energy system is currently underway. This transition will have a profound impact on the methods by which the European Union generates, distributes, stores, and consumes energy. It necessitates the adoption of power generation methods that are virtually free of carbon emissions, as well as a significant increase in energy efficiency. Additionally, the decarbonization of transportation, buildings, and industry is crucial to achieving this transition. Stakeholders involved in this endeavor must actively explore all available options to effectively limit energy-related CO2 emissions to less than 770 megatons (Mt) annually by the year 2050. Also, the successful implementation of the energy transition in the EU will heavily rely on the widespread utilization of hydrogen. Hydrogen is indispensable for the decarbonization of the gas grid, which connects Europe’s industrial sector and provides heating for over 40% of EU households and contributes to 15% of EU power generation [23]. Additionally, the figure below illustrates the increasing demand for hydrogen across various industries. According to Fig. 12.2, by the year 2050, hydrogen has the potential to fulfill approximately 24% of the total energy demand in the EU, equivalent to around 2250 terawatt hours (TWH) of energy.

12.6.1

Horizon 2019

The 3Emotion project, funded by the European Union, aims to revolutionize public transport in Pau, France, by deploying a fleet of eight advanced hydrogen fuel cell Van Hool Exquis tram-buses over the next 2 years. This initiative marks the introduction of hydrogen fuel cell electric buses in France and the world’s first comprehensive Bus Rapid Transit (BRT) system with 18-meter articulated tram buses. The project integrates hydrogen technology as an energy source, demonstrating significant engineering progress. The tram buses feature a zero-emission electric hybrid powertrain that combines PEM fuel cells, lithium batteries, and electric motors. They efficiently harness braking energy, ensuring additional power availability. The project showcases that sustainable choices can enhance the quality and efficiency of public transportation. Key stakeholders, including Van Hool,

308

H. Biabani et al.

Final energy 14,100 demand 2% Thereof H2

11,500 4%

9,300 6%

8%

24% 2,251 112

675

579

780 665 65 33

70 62

481

8

237

43

85 207

257

427

427

391

391

Business as usual

Ambitious

Business as usual

Ambitious

2030

4

5

6 7

Power generation, buffering

Transportation

Heating and power for buildings

Industry energy New Industry feedstock

1

325

2015

1-3

Existing industry feedstock

2050

Fig. 12.2 The increasing demand for hydrogen across various industries [23]

SMTUPPP, Engie, and ITM Power, collaborate to integrate the hydrogen fuel cell tram-buses into Pau’s urban transportation network. This aligns with the city’s urban redevelopment project and provides an environmentally friendly transportation solution. The Bus Rapid Transit service line offers an alternative to traditional railbased transport, reducing Pau’s carbon footprint and improving air quality. The deployment of these cutting-edge tram-buses signifies a defining moment in Pau’s transportation history, positioning the city as a trailblazer in eco-friendly public transportation. The project sets an example for cities worldwide, encouraging sustainable solutions that prioritize the well-being of citizens and the environment [24, 25].

12.6.2

Horizon 2020

Horizon 2020 was the EU’s research and innovation program for the 2014–2020 period. Several projects related to green hydrogen were funded under this program. For example, the REFHYNE project aimed to demonstrate the feasibility and

12

Green Hydrogen Research and Development Projects in the European Union

309

scalability of renewable hydrogen production through a 10 MW electrolyzer powered by wind energy in Germany. The project successfully produced green hydrogen at a competitive cost and demonstrated large-scale green hydrogen production’s technical and economic viability. Another Horizon 2020 project, the GenComm project, aimed to develop an integrated green hydrogen energy system for off-grid communities in Ireland, Scotland, and Spain. The project developed a modular system that combines wind and solar power with hydrogen storage and fuel cells to provide a reliable and sustainable source of energy for remote communities. The project demonstrated the potential of green hydrogen to enable energy independence and improve the quality of life in remote areas.

12.6.2.1

REFHYNE Project

The REFHYNE project is at the forefront of providing clean refined hydrogen for Europe. At the Shell Rhineland Energy and Chemicals Park in Wesseling, Germany, the project aims to build and run Europe’s largest PEM electrolyzer with assistance from the European Commission’s Fuel Cells and Hydrogen Joint Undertaking. This groundbreaking initiative aligns with the European Union’s ambitious strategy to combat climate change and achieve carbon neutrality by 2050. Its primary focus is on producing green hydrogen using renewable energy sources to contribute to carbon emissions reduction. Shell, a prominent energy company, and ITM Power, a renowned manufacturer of electrolyzers in the Rhineland region of Germany, have formed a collaborative joint venture for the REFHYNE project. At the heart of the REFHYNE project lies the establishment of a large-scale electrolyzer plant that utilizes wind and solar energy to generate green hydrogen. With a capacity of 10 MW, this state-of-the-art facility is projected to produce approximately 1300 tons of hydrogen annually. Figure 12.3 illustrates the expected layout of the ITM PEM electrolyzer. The produced hydrogen, which is decarbonized, can be fully integrated into refinery processes, including the desulfurization of conventional fuels. The ultimate aim is to create a reliable and sustainable source of green hydrogen that can be utilized in various sectors, such as fuel and chemical production, transportation, and industry. Additionally, the project aligns with the European Union’s objectives of transforming the transportation sector by providing a low-carbon alternative for hydrogen-powered vehicles, reducing harmful emissions. Furthermore, integrating green hydrogen into energy storage systems enhances grid stability. It facilitates the efficient utilization of renewable energy sources, contributing to the broader goal of establishing a more sustainable and resilient energy landscape. Figure 12.4 depicts the planned hydrogen production and consumption chain in the REFHYNE project. Beyond its immediate impact, the REFYNE project has significant implications for the commercial viability of green hydrogen production. As renewable energy sources become more cost-effective and widespread, green hydrogen is expected to play a pivotal role in the global transition to a low-carbon economy.

310

H. Biabani et al.

Step Down Transformers

Rectifiers Pump Skid

Control Panels

Phase Separation

PEM Stack Skid

Air Blast Coolers

Control Room

Water Trap Hydrogen Purifier

Water Purification

Refrigerant Chillers

Fig. 12.3 The ITM power PEM electrolyzer expected layout [26]

Fig. 12.4 Planned hydrogen production and consumption chain in the REFHYNE project

By demonstrating the feasibility and economic viability of large-scale green hydrogen production, the REFHYNE project not only sets the stage for future investments but also instills confidence in the transformative potential of this clean energy solution. With the goal of achieving 6 gigawatts (GW) of electrolyzer

12

Green Hydrogen Research and Development Projects in the European Union

311

capacity by 2024 and scaling up to an astonishing 40 GW in Europe by 2030, the project highlights the European Union’s commitment to driving sustainable hydrogen production. Currently, the electrolyzer industry operates on a smaller scale, with the largest PEM electrolyzer plants under construction typically reaching around 10 megawatts (MW), and global production remaining below 100 MW per year. The REFHYNE project serves as a guiding light, illuminating the path toward achieving these ambitious targets and advancing the renewable hydrogen sector through technological innovation and large-scale implementation. The project will utilize the produced hydrogen for various purposes, including: – Processing and upgrading products at the Wesseling refinery site – Testing the PEM technology on the largest scale achieved thus far – Exploring potential applications in other sectors, such as industry, power generation, heating for buildings, and transportation The REFHYNE project began in January 2018 and was originally expected to continue for 5 years until December 2022. This project has received financing from the Horizon 2020 research and innovation program of the European Union [27–29].

12.6.2.2

GENCOMM Project

The EU Interreg North-West Europe (NWE) program provided €9.3 million in funding for the GENCOMM project. Its goal was to demonstrate hydrogen (H2) as a reliable, clean, and secure energy storage technology for communities by the end of 2020. H2 can be used to store or transform intermittent renewable energy sources like wind, solar, and biomass into a variety of energy forms, including electricity, fuel for cars, biomethane, heat, and valuable chemicals. This approach will enhance the adoption of renewable energy, decrease emissions, and improve energy security in European communities [30, 31]. The objectives of the GenComm project are as follows: • Establishing and operating three pilot-scale renewable H2 energy storage sites in a safe manner • Developing technical and economic models to evaluate the performance of these facilities • Creating an online decision support tool called H2GO, based on map visualization, to aid other communities in implementing renewable H2 systems • Establishing the Europe-wide Community H2 Forum (CH2F) to promote H2 energy storage initiatives throughout Europe

312

12.6.3

H. Biabani et al.

Horizon Europe Projects

Firstly, it is necessary to mention that Horizon 2020 program has concluded its funding period from 2014 to 2020. The successor program, Horizon Europe, is now in place for the 2021–2027 period, continuing the EU’s commitment to research and innovation. Several projects related to green hydrogen have already been funded under this program. For example, the HIPERION project aims to develop a highly efficient and cost-effective electrolyzer for green hydrogen production. The project will leverage advanced materials and manufacturing techniques to reduce the cost and increase electrolysis efficiency. The project aims to achieve a hydrogen production cost of less than 2€/kg, making green hydrogen competitive with fossil fuels [32, 33]. Another Horizon Europe project, the HEAVENN project, aims to develop a sustainable and scalable green hydrogen supply chain for the aviation sector. The project will focus on the development of a green hydrogen production facility, the use of green hydrogen as a fuel for aircraft, and the development of a logistics and distribution network for green hydrogen. The project aims to demonstrate the technical and economic feasibility of green hydrogen as a fuel for the aviation sector, a major contributor to global emissions [34]. The Northern Netherlands region is making significant strides toward becoming a leader in the production and use of hydrogen energy. By 2026, the area hopes to establish itself as a “Hydrogen Valley,” which will house the full hydrogen value chain. This will include the production, storage, distribution, and local end-use of hydrogen energy. The ambitious plan for the region has recently been awarded a grant of €20 million from the Fuel Cell and Hydrogen Joint Undertaking, a publicprivate cooperation between the European Commission, industry organization Hydrogen Europe, and the sector’s research organization Hydrogen Europe Research. The HEAVENN project, which brings together 29 public and private stakeholders from seven European nations, received the money. The project aims to develop the infrastructure required to support the widespread use of hydrogen energy in the region. When the project is finished, there will be at least 10 fuel-cell-powered buses, 105 vehicles overall (including cars and vans), eight heavy-duty trucks, four garbage trucks, and one inland ship on the roads. This will significantly reduce the region’s carbon footprint and help to create a more sustainable future. The research will also aid the aviation industry since “drop-in e-fuels” created from hydrogen will be able to be utilized in current aircraft without altering the engine. As a result, carbon emissions from the aviation sector will be reduced, improving the sustainability of air travel. On the other hand, while paying attention to transportation, this project will also focus on the heating and electricity supply of residential buildings. One hundred new residential structures in Hoogeveen (northeastern Netherlands) are anticipated to be heated using fuel cells and hydrogen boilers. Trucks will initially provide hydrogen to the homes, but later on a dedicated pipeline may be financially viable. Between 250 and 1200 homes in neighboring communities will be connected to heat them using a mixture of natural gas and hydrogen. However, this will require

12

Green Hydrogen Research and Development Projects in the European Union

313

existing boilers to be made hydrogen-ready, which will be an interesting endeavor. Additionally, the project will use already-existing natural gas pipelines to transfer hydrogen between industrial locations, and Veendam’s (The northeastern part of the Netherlands, in the province of Groningen) Hystock storage facility will have subterranean storage available [35, 36]. This will create a robust infrastructure that will support the widespread use of hydrogen energy in the region. In fact, the Northern Netherlands region’s plan to become a “Hydrogen Valley” is an ambitious and exciting project that will significantly reduce the region’s carbon footprint and create a more sustainable future [37]. With the support of the €20 million grant from the Fuel Cell and Hydrogen Joint Undertaking, the region is well on its way to achieving its goal [38]. The project will not only benefit the region but also serve as a model for other regions looking to transition to a more sustainable future [39].

12.6.4

Clean Hydrogen Joint Undertaking

Europe is becoming the front line of the clean energy revolution in the world. The Clean Hydrogen Joint Undertaking (CH JU) is a public-private collaboration advancing hydrogen technologies in Europe. It supports the EU’s energy and climate targets, fostering climate neutrality by 2050. The CH JU promotes economic growth, job creation, and the hydrogen industry, benefiting small and medium-sized businesses (SMEs). It enhances research, knowledge creation, and breakthrough discoveries, driving scientific and technological progress. The CH JU aims to achieve a sustainable, decarbonized energy system, implementing the EU’s hydrogen strategy and shaping the future energy landscape. The benefits of the Clean Hydrogen Joint Undertaking encompass a wide spectrum of stakeholders, offering profound advantages on multiple fronts. Firstly, from a public perspective, this collaborative effort plays a pivotal role in achieving the European Union’s ambitious 2030 energy and climate targets. More significantly, it serves as a transformative pathway toward attaining climate neutrality by 2050, a cornerstone of global sustainability [40].

12.6.4.1

Green Hysland Project

The Green Hysland project aims to establish a cutting-edge “green hydrogen ecosystem” in the Balearic Islands, particularly Mallorca, Spain, by generating, distributing, and utilizing at least 300 tons of renewable hydrogen annually. This initiative holds great promise for sustainable development across multiple sectors. The project envisions using green hydrogen as a fuel source for fuel cell buses, rental vehicles, and maritime transportation, while also meeting the heat and power demands of commercial and public buildings. Furthermore, the project seeks to integrate green hydrogen into the island’s gas pipeline network to decarbonize the gas supply. The collaboration between esteemed entities such as the Spanish Ministry of Industry, Trade and Tourism, the Ministry for the Ecological Transition and the Demographic

314

H. Biabani et al.

Challenge, and the Balearic Government demonstrates their shared commitment to advancing sustainable energy practices. Spain’s strategic position in the renewable hydrogen landscape offers significant potential for achieving climate neutrality, generating employment opportunities, and fostering economic growth. With a timeline from 2021 to 2025, the Green Hysland project aims to substantially reduce CO2 emissions on Mallorca by up to 20,700 tons per year. The project underscores the importance of coordination and cooperation among various stakeholders, including governmental entities, research institutions, and industry partners, to drive the transition toward cleaner and more sustainable energy systems. It serves as a milestone in Southern Europe’s pursuit of sustainable energy solutions, leading to cleaner transportation, energy-efficient buildings, and a notable reduction in carbon emissions. As the project progresses, it will inspire similar endeavors and contribute to a more sustainable and prosperous future [41, 42]. This research has been funded by the Fuel Cells and Hydrogen 2 Joint Undertaking, Hydrogen Europe, and Hydrogen Europe Research, all of which are supported by the European Union’s Horizon 2020 Research and Innovation program [43].

12.6.4.2

H2FUTURE Project

The H2Future initiative is a pioneering project that aims to establish a sustainable hydrogen economy in Austria and contribute to the global transition toward a sustainable energy future. The project uses renewable energy sources such as wind and solar power to develop a large-scale electrolysis plant, hydrogen storage and transportation technologies, and a hydrogen refueling infrastructure. It is a collaborative effort involving various companies and research institutions, including VERBUND, Siemens, voestalpine, and the Austrian Institute of Technology, with coordination by VERBUND. A primary objective of the H2Future project is the construction of a 6 MW electrolysis plant that utilizes renewable electricity to produce green hydrogen on an industrial scale. This hydrogen will be utilized as a clean energy source for transportation, industry, and heating applications, thereby reducing dependence on fossil fuels and addressing the impacts of climate change. The project is a significant step toward Austria’s goal of achieving carbon neutrality by 2040. It aligns with the European Union’s efforts to reduce greenhouse gas emissions and transition to a low-carbon economy. A major milestone of the H2Future project is the installation and operation of a 6 MW PEM electrolysis system at the Voestalpine Linz steel plant in Austria. This system, the largest of its kind worldwide, can produce up to 1200 cubic meters of hydrogen per hour, sufficient to supply approximately 50 fuel cell buses or 600 passenger cars [44]. The system offers high efficiency and reliability based on Siemens Energy’s advanced PEM electrolysis technology, making it well-suited for large-scale green hydrogen production. Additionally, the H2Future project encompasses the development of hydrogen storage and transportation technologies. These technologies are vital for ensuring the efficient and safe transport of hydrogen, a highly flammable gas. The project investigates various storage solutions, including underground salt

12

Green Hydrogen Research and Development Projects in the European Union

315

caverns, and explores transportation options such as pipeline networks, truck transport, and shipping in liquid form. The H2Future project distinguishes itself through its strong emphasis on research and development. Collaborating with research institutions like the Dutch research center TNO and K1-MET in Austria, the project aims to replicate experimental results on larger scales in EU28 for the steel industry. This research will enhance the efficiency of the electrolysis process and identify new applications for green hydrogen. Beyond Austria and the European Union, the significance of the H2Future project extends globally. It is consistent with the Paris Agreement’s objectives to keep global warming well below two degrees Celsius. The project aids in limiting the effects of climate change by lowering greenhouse gas emissions and encouraging the use of renewable energy sources. Furthermore, it supports the United Nations’ Sustainable Development Goals, particularly Goal 7 (affordable and clean energy) and Goal 13 (climate action) [45–47].

12.6.4.3

Hybalance Project

The Hybalance project is a pioneering research and development initiative focused on addressing the need for efficient and reliable energy storage solutions. Its primary objective is to develop a hydrogen fuel cell system specifically designed for stationary energy storage applications [48]. Generously funded by the European Union, the project brings together a consortium of partners from industry, academia, and research institutions. The overarching goal of the Hybalance project is to engineer an economically viable and highly efficient hydrogen fuel cell system capable of seamlessly harnessing and storing renewable energy from sources like wind and solar power [49]. This system aims to provide a reliable and robust energy source for various sectors, including residential, commercial, and industrial applications. To achieve this ambitious vision, the Hybalance project involves the development of key components that enhance performance and reliability. The overall schematic of this project is presented in Fig. 12.5. These components include an advanced hydrogen storage system, a cutting-edge fuel cell stack, and a sophisticated power management system. The hydrogen storage system employs innovative design principles to ensure efficient hydrogen gas storage at high pressure, maximizing the system’s energy storage potential. Simultaneously, the fuel cell stack utilizes state-of-the-art technologies to convert stored hydrogen gas into a steady stream of electricity. The power management system facilitates seamless integration and optimized utilization, ensuring reliability, stability, and optimal performance by precisely managing energy flow from the fuel cell system to end-users. One remarkable advantage of the Hybalance system is its ability to store significant quantities of energy over extended durations. This attribute makes it an ideal solution for areas with intermittent or unreliable power supplies, such as remote communities or regions with substantial renewable energy generation capacities. By harnessing the full potential of the Hybalance system, these communities can overcome the limitations of conventional energy storage solutions and embrace a future powered by clean, dependable, and sustainable energy sources. The implications of the

316

H. Biabani et al.

Fig. 12.5 The schematic of the Hybalance Project [53]

Hybalance project extend far beyond its immediate scope, signaling a transformative shift in the hydrogen fuel cell industry. The project plays a crucial role in shaping a sustainable and environmentally conscious energy landscape by demonstrating the viability and unlocking the full potential of this cutting-edge technology for energy storage applications. The successful integration of the Hybalance system into existing power grids enables the seamless assimilation of renewable energy sources, fostering a greener, more resilient, and prosperous future while preserving the delicate balance of our planet [50–52].

12.7

Implications and Conclusion

The Horizon projects related to green hydrogen demonstrate the EU’s commitment to promoting sustainable and decarbonized energy systems. Green hydrogen has the potential to play a key role in the energy transition by enabling the integration of renewable energy sources and reducing greenhouse gas emissions. The Horizon projects have contributed to the development of innovative technologies and solutions for green hydrogens. The Horizon Green Hydrogen Project in the European Union is a groundbreaking initiative that has the potential to revolutionize the energy sector and pave the way for a more sustainable future. This project is a significant

12

Green Hydrogen Research and Development Projects in the European Union

317

step toward achieving the EU’s ambitious climate goals and reducing carbon emissions. The project aims to produce green hydrogen through renewable sources such as wind, solar, and hydropower. This will not only reduce carbon emissions but also help in the fight against climate change. Green hydrogen is a clean and sustainable alternative to fossil fuels and can be used in a wide range of applications, including transportation, industry, and heating. Moreover, the Horizon Green Hydrogen Project has the potential to create new job opportunities and boost economic growth in the region. The project will require skilled workers and investment in infrastructure, which will create new jobs and stimulate economic activity. This will not only benefit the EU but also help to address global challenges such as unemployment and poverty. However, the success of the project will depend on the availability of funding, supportive policies and regulations, and the willingness of stakeholders to invest in green hydrogen technology. The EU must provide adequate funding and support to ensure the project’s success and encourage private-sector investment in green hydrogen technology. The Horizon Green Hydrogen Project is a promising initiative that could pave the way for a cleaner and more sustainable energy future in the EU. It is a significant step toward achieving the EU’s climate goals and reducing carbon emissions. The project has the potential to create new job opportunities, boost economic growth, and promote sustainable development. The EU must continue supporting and investing in green hydrogen technology to ensure a sustainable future for future generations.

References 1. Wulf, C., Linßen, J., & Zapp, P. (2018). Review of power-to-gas projects in Europe. Energy Procedia, 155, 367–378. 2. Zhang, X., Jiao, K., Zhang, J., & Guo, Z. (2021). A review on low carbon emissions projects of steel industry in the world. Journal of Cleaner Production, 306, 127259. 3. Hayati, M. M., Aminlou, A., Zare, K., & Abapour, M. (2023). A two-stage stochastic optimization scheduling approach for integrating renewable energy sources and deferrable demand in the spinning reserve market. In 2023 8th International Conference on Technology and Energy Management (ICTEM), IEEE, pp. 1–7. 4. EU Hydrogen strategy. Available: https://energy.ec.europa.eu/topics/energy-systems-integra tion/hydrogen_en. Accessed 1 June 2023. 5. Daneshvar, M., Mohammadi-Ivatloo, B., Zare, K., & Asadi, S. (2021). Transactive energy management for optimal scheduling of interconnected microgrids with hydrogen energy storage. International Journal of Hydrogen Energy, 46(30), 16267–16278. 6. Mansour-Saatloo, A., Agabalaye-Rahvar, M., Mirzaei, M. A., Mohammadi-Ivatloo, B., Abapour, M., & Zare, K. (2020). Robust scheduling of hydrogen based smart micro energy hub with integrated demand response. Journal of Cleaner Production, 267, 122041. 7. Sammes, N. (2006). Fuel cell technology: Reaching towards commercialization. Springer Science & Business Media. 8. Du, Z., Li, H., & Gu, T. (2007). A state of the art review on microbial fuel cells: A promising technology for wastewater treatment and bioenergy. Biotechnology Advances, 25(5), 464–482.

318

H. Biabani et al.

9. Mohammadi, M., Noorollahi, Y., Mohammadi-Ivatloo, B., & Yousefi, H. (2017). Energy hub: From a model to a concept–A review. Renewable and Sustainable Energy Reviews, 80, 1512–1527. 10. Nazari-Heris, M., Abapour, S., & Mohammadi-Ivatloo, B. (2017). Optimal economic dispatch of FC-CHP based heat and power micro-grids. Applied Thermal Engineering, 114, 756–769. 11. Mansour-Saatloo, A., et al. (2021). Multi-objective IGDT-based scheduling of low-carbon multi-energy microgrids integrated with hydrogen refueling stations and electric vehicle parking lots. Sustainable Cities and Society, 74, 103197. 12. Nojavan, S., Zare, K., & Mohammadi-Ivatloo, B. (2017). Application of fuel cell and electrolyzer as hydrogen energy storage system in energy management of electricity energy retailer in the presence of the renewable energy sources and plug-in electric vehicles. Energy Conversion and Management, 136, 404–417. 13. Thomas, J. M., Edwards, P. P., Dobson, P. J., & Owen, G. P. (2020). Decarbonising energy: The developing international activity in hydrogen technologies and fuel cells. Journal of Energy Chemistry, 51, 405–415. 14. Phadke, M. (2020). The birth of a hydrogen economy? Tribology & Lubrication Technology, 76(11), 18–22. 15. Commission outlines European Hydrogen Bank to boost renewable hydrogen. Available: https://energy.ec.europa.eu/news/commission-outlines-european-hydrogen-bank-boost-renew able-hydrogen-2023-03-16_en. Accessed 1 June 2023. 16. Bernd Weber, C. (2023). Design Options for a European Hydrogen Bank, Konrad-AdenauerStiftung. Germany. Retrieved from https://policycommons.net/artifacts/3676911/designoptions-for-a-european-hydrogenbank/4482753/ on 25 Jan 2024. CID: 20.500.12592/tznwz4. 17. Kar, S. K., Sinha, A. S. K., Bansal, R., Shabani, B., & Harichandan, S. (2023). Overview of hydrogen economy in Australia. Wiley Interdisciplinary Reviews: Energy and Environment, 12(1), e457. 18. Green hydrogen production. Available: https://www.siemens-energy.com/global/en/offerings/ renewable-energy/hydrogen-solutions.html. Accessed 1 May 2023. 19. Delpierre, M., Quist, J., Mertens, J., Prieur-Vernat, A., & Cucurachi, S. (2021). Assessing the environmental impacts of wind-based hydrogen production in the Netherlands using ex-ante LCA and scenarios analysis. Journal of Cleaner Production, 299, 126866. 20. The hydrogen energy infrastructure development in Japan,” presented at the International Conference Green Energy and Smart Grids (2018), Russian Federation, 2018. 21. Hernandez, D. D., & Gençer, E. (2021). Techno-economic analysis of balancing California’s power system on a seasonal basis: Hydrogen vs. lithium-ion batteries. Applied Energy, 300, 117314. 22. Los Angeles plans to jump-start a green hydrogen market in the US. Available: https://www. canarymedia.com/articles/hydrogen/los-angeles-seeks-to-jump-start-a-green-hydrogen-marketin-the-u-s. Accessed 10 May 2023. 23. Cells, F., & Undertaking, H. J. (2019). Hydrogen roadmap Europe: A sustainable pathway for the European energy transition. Publications Office. 24. 3Emotion project. Available: https://www.3emotion.eu/. Accessed 1 June 2023. 25. Barrett, S. (2017). Van Hool hydrogen fuel cell tram-buses for Pau in 3Emotion. Fuel Cells Bulletin, 2017(9), 14. 26. Project overview presentation-10 MW electrolyser Rhineland refinery. Available: https:// refhyne.eu/wp-content/uploads/2020/06/REFHYNE-project-overview-Sep18.pdf. Accessed 19 Mar 2023. 27. REFHYNE project overview. Available: https://www.refhyne.eu/about/. Accessed 19 Mar 2023. 28. Smith, W. A., Burdyny, T., Vermaas, D. A., & Geerlings, H. (2019). Pathways to industrialscale fuel out of thin air from CO2 electrolysis. Joule, 3(8), 1822–1834. 29. Dragoon, K., et al. (2022). Hydrogen as part of a 100% clean energy system: Exploring its decarbonization roles. IEEE Power and Energy Magazine, 20(4), 85–95.

12

Green Hydrogen Research and Development Projects in the European Union

319

30. GENCOMM H2 energy storage. Available: https://www.universityofgalway.ie/therme/pro jects/active/gencomm/. Accessed 4 June 2023. 31. GENCOMM: GENerating energy secure COMMunities through Smart Renewable Hydrogen. Available: https://vb.nweurope.eu/projects/project-search/gencomm-generating-energy-securecommunities/. Accessed 4 June 2023. 32. HIPERION project overview. Available: https://hiperion-project.eu/. Accessed 10 May 2023. 33. Schrotenboer, A. H., Veenstra, A. A., uit het Broek, M. A., & Ursavas, E. (2022). A Green Hydrogen Energy System: Optimal control strategies for integrated hydrogen storage and power generation with wind energy. Renewable and Sustainable Energy Reviews, 168, 112744. 34. HEAVENN project overview. Available: https://heavenn.org/about/. Accessed 10 May 2023. 35. Hydrogen valley – New energy coalition. Available: https://www.newenergycoalition.org/en/ hydrogen-valley/. Accessed 5 May 2023. 36. Marzi, E., Morini, M., & Gambarotta, A. (2022). Analysis of the status of research and innovation actions on electrofuels under horizon 2020. Energies, 15(2), 618. 37. Khan, M. I., & Al-Ghamdi, S. G. (2023). Hydrogen economy for sustainable development in GCC countries: A SWOT analysis considering current situation, challenges, and prospects. International Journal of Hydrogen Energy, 48(28), 10315–10344. 38. Hydrogen Joint Undertaking. (2017). Fuel cell and hydrogen technology-Europe’s journey to a greener world. 39. Mission innovation hydrogen valley platform. Available: https://h2v.eu/. Accessed 5 May 2023. 40. Clean hydrogen joint undertaking. Available: https://european-union.europa.eu/institutionslaw-budget/institutions-and-bodies/search-all-eu-institutions-and-bodies/clean-hydrogen-jointundertaking_en. Accessed 10 May 2023. 41. Green hysland deployment of H2 ecosystem on the island of Mallorca. Available: https:// greenhysland.eu/. Accessed 10 May 2023. 42. Clean hydrogen partnership. Available: https://www.clean-hydrogen.europa.eu/media/news/ green-hysland-mallorca-first-green-hydrogen-project-mediterranean-country-due-get-euro pean-funding-2020-10-19_en. Accessed 10 May 2023. 43. GREEN HYSLAND project. Available: https://greenhysland.eu/about-green-hysland/. Accessed 10 May 2023. 44. H2FUTURE – A European flagship project. Available: https://www.h2future-project.eu/. Accessed 25 Apr 2023. 45. Tang, J., Chu, M.-S., Li, F., Feng, C., Liu, Z.-G., & Zhou, Y.-S. (2020). Development and progress on hydrogen metallurgy. International Journal of Minerals, Metallurgy and Materials, 27, 713–723. 46. Eder, W. (2017). Environment–Climate–Energy: Quo Vadis, Industry? BHM Berg-und Hüttenmännische Monatshefte, 11(162), 494–497. 47. Sasiain Conde, A., Rechberger, K., Spanlang, A., Wolfmeir, H., & Harris, C. (2021). Decarbonization of the steel industry. A techno-economic analysis. Matériaux & Techniques, 109. 48. Biebuyck, B. (2017). Green hydrogen production through electrolysis. Available: https:// hybalance.eu/wp-content/uploads/2016/03/FCH-JU_HyBalance-inauguration.pdf 49. Liao, M., Liu, C., & Qing, Z. (2020). A recent overview of power-to-gas projects. In 2020 IEEE 4th Conference on Energy Internet and Energy System Integration (EI2), IEEE, pp. 2282–2286. 50. Zong Y., et al. (2020). Identifying the system-related conditions and consequences of power-tox solutions for a high renewables penetration in denmark. In 2020 IEEE 4th Conference on Energy Internet and Energy System Integration (EI2), IEEE, pp. 955–960.

320

H. Biabani et al.

51. Zemite, L., et al. (2023). A comprehensive overview of the Europen and Baltic landscape for hydrogen applications and innovations. Latvian Journal of Physics and Technical Sciences, 60(3), 33–53. 52. HyBalance – Green energy project Denmark. Available: https://hybalance.eu/. Accessed 25 Apr 2023. 53. Hydrogen Valley. Available: https://brintbranchen.dk/en/hydrogen-valley-makes-hydrogendreams-come-true/. Accessed 25 Apr 2023.

Chapter 13

Hydrogen-Combined Smart Electrical Power Systems: An Overview of United States Projects Ashkan Safari , Mohammad Mohsen Hayati and Morteza Nazari-Heris

13.1

,

Introduction

The rapid acceleration of the global energy transition towards a low-carbon future [1, 2] has amplified the importance of GH2 technologies as a critical component of sustainable energy systems. GH2, produced through electrolysis powered by renewable sources, offers immense potential as an energy carrier and a means to decarbonize various sectors. In the context of smart electrical power systems, hydrogen plays a pivotal role in enhancing grid flexibility, resilience, and the integration of renewable energy sources. In recent years, careful examination of the existing infrastructure for hydrogen supply has been undertaken to assess the advancements made in realizing a hydrogen-based society [3]. The transition towards hydrogen, an abundant and widely distributed resource, has the potential to reshape landscape technologies and forms, altering the dynamics of electrical energy production and consumption. This shift may lead to a transformation where current importers become future exporters. Hence, the implications of hydrogen are vast and significant. Sectors solely focused on this production will lag in the pursuit of tomorrow’s prized hydrogen resources [4, 5]. The 2000 World Energy Assessment conducted by multiple UN agencies and the World Energy Council highlights hydrogen as a strategically significant energy carrier. The widespread adoption of hydrogen as a

A. Safari · M. M. Hayati (✉) Faculty of Electrical and Computer Engineering, Smart Energy Systems Lab, University of Tabriz, Tabriz, Iran e-mail: [email protected]; [email protected] M. Nazari-Heris College of Engineering, Lawrence Technological University, Southfield, MI, USA e-mail: [email protected] © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 V. Vahidinasab et al. (eds.), Green Hydrogen in Power Systems, Green Energy and Technology, https://doi.org/10.1007/978-3-031-52429-5_13

321

322

A. Safari et al.

replacement for fossil fuels has the potential to achieve substantial reductions in carbon emissions, preventing a doubling of pre-industrial levels of atmospheric carbon dioxide (Co2) [6, 7]. Such a doubling is projected to lead to significant and potentially irreversible ecological and economic disruptions. Moreover, hydrogen offers a pathway to transition from finite, non-renewable fossil fuel reserves to an inexhaustible supply of renewable energy sources, thus playing a critical role in the decarbonization of the global energy system necessary to mitigate the consequences of climate change [8, 9]. Additionally, hydrogen fuel cells hold promise in addressing global energy disparities and inequities [10, 11]. This chapter provides a comprehensive review of key initiatives and projects driving the adoption and advancement of GH2 technologies within smart grids. It explores the involvement of prominent institutions and organizations, such as the National Renewable Energy Laboratory (NREL), in pushing the boundaries of GH2 research and development. NREL’s contribution to hydrogen production, storage, fuel cells, and integrated energy systems will be examined to shed light on the cutting-edge advancements in this field. The Western Green Hydrogen Initiative is another crucial endeavor that will be discussed, focusing on its objectives of developing a regional GH2 economy in the western United States. Through collaborative efforts among stakeholders, this initiative aims to accelerate the deployment of GH2 technologies, fostering innovation and economic growth. Furthermore, the Gulf Coast Clean Energy Application Center offers a unique perspective on utilizing existing industrial infrastructure along the Gulf Coast for GH2 development. This initiative is driving the region’s transformation into a hub for GH2 production and utilization by identifying opportunities, supporting project development, and facilitating collaboration among industry stakeholders. The Hawaii Hydrogen Initiative presents a case study of Hawaii’s ambitious renewable energy goals and the integral role of GH2 in supporting the integration of intermittent renewable sources. Demonstrating various projects, research initiatives, and collaborations within the energy sector, Hawaii exemplifies the potential of GH2 as a key enabler of a sustainable and resilient energy system. The California Hydrogen Infrastructure Project, backed by the California Energy Commission, focuses on the development of a robust hydrogen infrastructure, including the construction of refueling stations and the deployment of fuel cell electric vehicles. This project aims to promote the widespread adoption of hydrogen as a clean fuel for transportation, further supporting the state’s ambitious climate and sustainability targets. Additionally, this chapter examines groundbreaking projects such as the Los Angeles Green Hydrogen Power-to-Gas Project, the Advanced Clean Energy Storage (ACES) Project, the Wind-to-Hydrogen Project in Hawaii, the Utah Advanced Clean Energy Storage (UACES) Project, the New York Offshore Wind-to-Hydrogen Project, and renewable hydrogen production efforts in the Pacific Northwest. These projects highlight the diverse applications of GH2, ranging from power generation and transportation to industrial processes, and demonstrate the pivotal role of hydrogen in achieving renewable energy goals and decarbonization targets.

13

Hydrogen-Combined Smart Electrical Power Systems: An Overview of. . .

323

Fig. 13.1 Overview of U.S. hydrogen projects described in the chapter

Fig. 13.2 U.S. renewable sources for electricity generation

Finally, the Texas Gulf Coast Hydrogen Hub showcases the establishment of a hydrogen hub that leverages the industrial infrastructure and abundant renewable resources of the Texas Gulf Coast region. This initiative drives the rapid growth of a hydrogen economy in the area, fostering economic development and positioning hydrogen as a key enabler of sustainable energy systems. By delving into these various initiatives and projects, this chapter provides a comprehensive overview of the advancements, challenges, and opportunities in integrating GH2 technologies within smart grids. It underscores the critical role of hydrogen as an enabler of a lowcarbon future and highlights the transformative potential of these developments in shaping the energy landscape. Some U.S. electrical energy generation using renewable resources is also denoted in Fig. 13.1 and 13.2.

324

13.2 13.2.1

A. Safari et al.

National Renewable Energy Laboratory (NREL) Hydrogen Technologies Description of NREL’s Involvement in Advancing Green Hydrogen Technologies

The NREL is at the forefront of advancing GH2 technologies. NREL is a renowned research institution that explores and develops sustainable energy solutions. In line with global efforts to transition to a low-carbon economy and promote a sustainable future, NREL plays a pivotal role in advancing the science, technology, and implementation of GH2 as a clean energy carrier. Hydrogen, as an energy carrier, has gained significant attention due to its potential to address global energy challenges and reduce greenhouse gas emissions. GH2, in particular, refers to hydrogen produced from renewable sources through electrolysis, such as wind, solar, or hydropower. This process involves splitting water molecules into hydrogen and oxygen using electricity derived from renewable sources. Also, NREL plays a crucial role in advancing GH2 technologies through research, development, and deployment efforts. The laboratory focuses on several key areas to drive the progress of these technologies. In other words, with backing from the U.S. Department of Energy (DOE), NREL works on developing comprehensive storage solutions, focusing on aspects such as material properties, system configurations, and interface requirements and analyzing the entire process from well to wheel. NREL also provides technical expertise to the DOE’s Hydrogen Storage Engineering Center of Excellence. Presently, NREL’s activities involve assessing the storage characteristics of innovative materials and contributing to achieving the Energy Department’s storage targets. NREL leads the Hydrogen Storage Characterization and Optimization Effort (HySCORE), a collaborative project involving the Pacific Northwest National Laboratory and the National Institute of Standards and Technology. The main aim of HySCORE is to advance the core capabilities of hydrogen storage for the Fuel Cell Technologies Office while validating claims, concepts, and theories related to hydrogen storage materials [12].

13.2.2

Overview of Projects Related to Hydrogen Production, Storage, Fuel Cells, and Integrated Energy Systems in NREL

Hydrogen Production NREL actively investigates and develops innovative technologies for the efficient and sustainable production of GH2 and developing novel catalysts. This involves research into both electrochemical and thermochemical pathways. NREL’s scientists explore advanced electrolysis technologies, such as proton exchange membrane (PEM) and solid oxide electrolysis cells (SOEC), to

13

Hydrogen-Combined Smart Electrical Power Systems: An Overview of. . .

325

produce hydrogen from water using renewable electricity. Additionally, NREL investigates thermochemical processes, such as solar-driven water splitting and biomass gasification, to enable high-efficiency hydrogen production while minimizing carbon emissions. Nevertheless, renewable energy sources inherently exhibit variability, necessitating energy storage or a combination of systems to accommodate fluctuations on a daily and seasonal basis. One approach involves generating hydrogen through electrolysis, using an electric current to split water molecules and subsequently utilizing this hydrogen in fuel cells to generate electricity during periods of low power production or high demand. Another application involves utilizing hydrogen in fuel cell vehicles. Also, by conducting screenings to identify naturally existing organisms that exhibit greater tolerance to oxygen, NREL scientists are engineering new genetic variations of these organisms that can maintain hydrogen production even in the presence of oxygen. Another approach involves developing a novel system that utilizes a metabolic switch, specifically sulfur deprivation, to alternate algal cells between the phases of photosynthetic growth and hydrogen production. NREL researchers are engaged in the development of pre-treatment methods aimed at converting lignocellulosic biomass into feedstocks that are rich in sugars. These sugars can be directly utilized in fermentation processes to generate hydrogen, ethanol, and valuable chemicals. Additionally, scientists are actively investigating a group of Clostridium bacteria that have the ability to directly ferment hemicellulose into hydrogen. Hydrogen can be generated through the thermal decomposition (pyrolysis) or conversion (gasification) of biomass resources, such as agricultural residues like peanut shells, consumer wastes like plastics and waste grease, or dedicated energy crops. Pyrolysis of biomass yields a liquid byproduct known as bio-oil, which contains a diverse range of components that can be separated into valuable chemicals and fuels, including hydrogen. Presently, NREL researchers are primarily focused on hydrogen production through the catalytic reforming of bio-oil derived from biomass pyrolysis. Their specific research endeavors involve the reforming of pyrolysis streams and the development and testing of catalysts that can be easily circulated as fluids [13, 14]. Hydrogen Storage and Distribution Efficient and safe hydrogen storage is crucial to its widespread adoption. NREL conducts comprehensive research to develop advanced hydrogen storage technologies. This includes exploring novel materials, such as metal hydrides, chemical hydrides, and advanced adsorbents, which can store high-capacity hydrogen and enable its release when needed. NREL also investigates the feasibility of utilizing underground geological formations for large-scale hydrogen storage, along with assessing the necessary infrastructure requirements. Also, NREL explores solid-state hydrogen storage technologies, such as metal-organic frameworks (MOFs) and complex hydrides. Additionally, NREL investigates the optimization of hydrogen transportation infrastructure, including pipelines and hydrogen refueling stations, to support the growth of a hydrogen economy. Hydrogen Distribution Infrastructure To enable the integration and widespread use of GH2, NREL investigates the development of a robust hydrogen distribution

326

A. Safari et al.

and infrastructure network. This includes analyzing the optimal layout of hydrogen refueling stations for transportation applications and evaluating the design and implementation of hydrogen pipelines for industrial usage. NREL’s research in this area aims to address technical challenges, optimize efficiency, and ensure the seamless integration of hydrogen into existing energy systems. Hydrogen Utilization NREL is deeply engaged in research and development related to hydrogen utilization, with a particular focus on fuel cell technologies. Fuel cells efficiently convert hydrogen into electricity, offering a range of applications in transportation, stationary power generation, and portable devices. NREL works on improving fuel cell performance, durability, and cost-effectiveness through advancements in catalysts, membrane materials, and system design. Furthermore, NREL explores the integration of fuel cells with renewable energy sources, such as solar and wind, to enable enhanced system performance and grid integration. NREL conducts unbiased and reliable assessments of emerging fuel cell technologies, emphasizing their performance, durability, and cost. As the demand for fuel cells increases, American manufacturers are actively working on developing these technologies for various applications. Through the National Fuel Cell Technology Evaluation Center (NFCTEC), participating fuel cell developers securely share pricing details and raw test data pertaining to the operation, maintenance, and safety of their fuel cell products with NREL. Projects funded by the U.S. Department of Energy aim to demonstrate the potential of fuel cell technology and facilitate the expansion of fuel cell markets. These initiatives play a vital role in transforming the market and promoting the growth of fuel cell technologies. Renewable Energy Integration The Advanced Research on Integrated Energy Systems (ARIES) serves as a research platform capable of addressing the intricacies of the contemporary energy system. It conducts comprehensive research to facilitate the advancement of revolutionary energy technologies. NREL explores integrating hydrogen production through electrolysis with the existing electricity grid, known as a power-to-gas system. These projects investigate the potential for utilizing excess renewable energy to produce hydrogen, which can be stored and used for various applications, including power generation, transportation, and industrial processes. The laboratory investigates methods for effectively coupling hydrogen production with variable renewable energy sources, such as wind and solar power. This includes developing strategies for grid integration, energy storage, and utilizing excess renewable energy for hydrogen production during periods of low demand. Energy Storage and Power Electronics The combination of large-scale storage technologies like batteries paired with thermal or hydrogen-based systems will play a crucial role in validating energy system models and controls. As these storage technologies transition from the laboratory to the multimegawatt level, ARIES will assist in proactively addressing performance and interface challenges associated with scaling. The ongoing advancement of power electronics is revolutionizing power system operation. ARIES is dedicated to addressing the inherent disparities between power electronic-based equipment and traditional devices, as well as the

13

Hydrogen-Combined Smart Electrical Power Systems: An Overview of. . .

327

barriers that must be overcome to enable greater integration of renewable energy generation. Through the integration of emerging power electronic technologies and system architectures, ARIES aims to facilitate the development of a future grid with resilient and flexible operation. Hybrid Energy Systems and the Advancement of Future Energy Infrastructure As the energy landscape evolves to encompass numerous distributed energy assets, the ARIES research platform stands out for its ability to replicate the diverse time scales, physical scales, and technologies found in these hybrid energy systems. ARIES creates a simulated environment that closely resembles the real world, incorporating high-precision, physics-based, real-time models that facilitate the connection between actual hardware devices and millions of simulated devices. This field of research aims to enhance the fundamental science behind optimizing and controlling large-scale energy systems in real time. By fostering innovation, ARIES contributes to the development of cutting-edge solutions for next-generation energy infrastructure. The future energy infrastructure research area focuses on the transmission and distribution networks that support various advanced fuel types and infrastructures, serving the power, transportation, buildings, and industrial sectors. ARIES enables the testing of grid designs that range from microgrids to high-voltage direct current transmission grids, as well as the evaluation of management and control systems that effectively integrate power delivery across diverse fuel and technology types. ARIES plays a crucial role in addressing the security vulnerabilities that arise when disparate hardware and software components are integrated. The platform leverages visualization, monitoring, and data processing capabilities to ensure ARIES research assets’ security and interconnectedness. Through the creation of a digital twin of research hardware clusters, ARIES is capable of simulating and identifying potential attacks on evolving communication and control systems, thereby mitigating vulnerabilities in energy systems as a whole. System Analysis and Techno-economic Studies NREL conducts techno-economic analysis to assess the viability and potential impacts of green hydrogen technologies. This involves evaluating the costs, performance, and environmental benefits of different hydrogen production pathways, storage options, and end-use applications. By providing comprehensive analyses, NREL assists policymakers, industry stakeholders, and investors in making informed decisions and shaping favorable policies for deploying green hydrogen technologies. Accordingly, the various scenarios for producing hydrogen and electricity are presented in Fig. 13.3. Accordingly, Fig. 13.3 illustrates different scenarios for hydrogen and renewable electricity generation according to the NREL framework.

328

A. Safari et al.

Fig. 13.3 Various scenarios for producing renewable hydrogen and electricity

13.3 Western Green Hydrogen Initiative 13.3.1 Explanation of the Initiative’s Goals to Develop a Regional GH2 Economy in the Western United States The Western Green Hydrogen Initiative, known as WGHI, is a collaborative effort between public and private entities to promote and expedite the implementation of green hydrogen infrastructure in the western region of the United States. The WGHI

13

Hydrogen-Combined Smart Electrical Power Systems: An Overview of. . .

329

operates as a steering committee, facilitating the development of a comprehensive regional strategy for green hydrogen. The primary objective of this strategy is to establish a sizable and long-lasting reserve of renewable energy storage through the utilization of GH2 technology. By leveraging the WGHI, economies and the environment in the Western region can reap the benefits of this initiative. The specific goals of the WGHI are: Scaling Up Green Hydrogen Production, Establishing a Robust Hydrogen Infrastructure, Fostering Market Development and Adoption. Therefore, WGHI aims to expedite the establishment of GH2 production facilities throughout the Western region of the United States. This endeavor involves utilizing the abundant renewable energy resources available in the area, such as solar, wind, and hydropower, to power the electrolysis process that generates hydrogen by splitting water molecules. A key objective of this ambitious strategy, focused on enhancing competitiveness in energy markets, is to construct hydrogen refueling stations to support fuel cell electric vehicles (FCEVs), integrate hydrogen storage facilities, and develop pipelines for efficient hydrogen transportation. Furthermore, the WGHI is dedicated to cultivating the market for GH2 and creating favorable investment conditions by offering financial incentives for GH2 projects. Additionally, the initiative aims to facilitate public-private partnerships to promote the widespread adoption of GH2 solutions.

13.3.2

Collaborative Efforts Among Stakeholders to Accelerate GH2 Deployment

WGHI is a collaborative effort led by the National Association of State Energy Officials (NASEO), the Western Interstate Energy Board (WIEB), and the GH2 Coalition (GHC), with support from Mitsubishi Power. It is an organization driven by the participating states, with four leading states serving as chairs and members of the Executive Committee responsible for the development and approval of work processes and deliverables. The founding organizations, namely the GH2 Coalition, the Western Interstate Energy Board, and the National Association of State Energy Officials support the states by facilitating discussions, gathering and disseminating information, conducting modeling and reporting, and assisting in the formulation of state-level roadmaps for GH2 as an integral part of the energy and economic planning. The WGHI comprises representatives from eleven western states and two Canadian provinces, all interconnected within the Western Interconnection power grid. Additionally, it includes the participation of additional states such as Florida, Ohio, and Louisiana. The WGHI also ensures the involvement of industry partners, non-profit stakeholders, and the Department of Energy to promote a comprehensive and inclusive approach toward advancing the GH2 agenda. These collaborative efforts pave the way for a cleaner, more sustainable future, driving the transition towards a carbon-neutral energy system.

330

13.3.3

A. Safari et al.

Gulf Coast Clean Energy Application Center

The Gulf Coast Clean Energy Application Center (GCCEAC) plays a crucial role within the broader Western Green Hydrogen Initiative, specifically aimed at driving the development of clean energy in the Gulf Coast region of the United States. As an integral part of the WGHI, the GCCEAC concentrates on advancing the implementation and utilization of green hydrogen technologies in the Gulf Coast area. The primary objective of the GCCEAC is to enhance market and regulatory conditions to facilitate the adoption of combined heat and power (CHP) technologies. The center has been instrumental in spearheading CHP development in Texas, Louisiana, and Oklahoma, resulting in substantial growth in the CHP market across the Gulf Coast region. These endeavors have yielded numerous benefits, including improved energy generation efficiency, reduced emissions, and increased infrastructure resilience. The Gulf Coast region has well-established markets for energy commodities, including natural gas, electricity, and petrochemical products. Integrating GH2 into these existing markets allows for seamless integration and distribution of green hydrogen. Additionally, The Gulf Coast region benefits from abundant and diverse energy inputs, including natural gas, solar resources, and wind potential. These inputs can be utilized in GH2 production processes, such as steam methane reforming or electrolysis, enabling the region to tap into its renewable and non-renewable energy sources to produce green hydrogen. By leveraging these energy inputs, the WGHI can optimize GH2 production and reduce dependency on fossil fuels [15].

13.4

Hawaii Hydrogen Initiative

The Hawaii Hydrogen Initiative (HHI) is a comprehensive and ambitious endeavor to advance the development and integration of hydrogen technologies in Hawaii. With a focus on promoting sustainability, energy independence, and carbon neutrality, the HHI seeks to leverage Hawaii’s unique characteristics and resources to establish a robust hydrogen economy. Hawaii has abundant renewable energy resources, including solar, wind, and geothermal. The HHI aims to capitalize on these resources by utilizing them to produce hydrogen through electrolysis, which splits water molecules into hydrogen and oxygen. By integrating renewable energy sources into hydrogen production, the HHI fosters the production of green hydrogen derived from renewable sources and emits no greenhouse gases during its use [16, 17]. The Hawaii Department of Business, Economic Development, and Tourism (DBEDT) initiated the Hawaii Renewable Hydrogen Program to oversee the state’s transition towards a renewable hydrogen-based economy. In support of this program, a dedicated Hydrogen Investment Capital Special Fund was established to provide initial financial support and venture capital investments for private sector and federal projects related to research, development, testing, and implementation of

13

Hydrogen-Combined Smart Electrical Power Systems: An Overview of. . .

331

the program. The Hawaii Public Utilities Commission (PUC) has been granted the authority to establish a rebate program specifically designed for renewable hydrogen fueling systems. This program offers the potential for rebates up to a maximum of $200,000, which can be utilized to install new hydrogen fueling systems or upgrade the capacity of existing ones.

13.5

California Hydrogen Infrastructure Project

The majority of hydrogen utilized in the United States is generated in close proximity to its consumption, often at large-scale industrial facilities. However, the development of a comprehensive infrastructure for distributing hydrogen to fueling stations across the country, necessary for the widespread adoption of fuel-cell electric vehicles, is still underway. The initial phase of vehicle and station deployment prioritizes the expansion of distribution networks, primarily concentrated in California’s southern and northern regions [18]. The California Hydrogen Infrastructure Project (CHIP) is a comprehensive and forward-thinking initiative to develop the necessary infrastructure to support the widespread deployment and utilization of hydrogen as a clean energy source in California. Recognizing the importance of hydrogen in achieving the state’s ambitious climate and energy goals, the project focuses on establishing a robust and interconnected network of hydrogen fueling stations across the state. CHIP aims to accelerate the construction and deployment of hydrogen fueling stations throughout California. By strategically locating these stations in key urban centers, transportation corridors, and highly populated areas, the project aims to ensure convenient access to hydrogen fuel for a growing fleet of fuel cell electric vehicles (FCEVs). This extensive network of fueling stations will support the adoption of FCEVs and encourage the use of hydrogen as a viable alternative to traditional fossil fuels. Additionally, CHIP promotes the development of advanced storage and dispensing systems to enhance the efficiency, safety, and reliability of hydrogen infrastructure [19].

13.5.1

California Energy Commission’s Support for Hydrogen Infrastructure Development

The California Energy Commission (CEC), through its support and initiatives, plays a pivotal role in fostering the growth and advancement of hydrogen infrastructure in the state, contributing to the overall transition towards a more sustainable and clean transportation system. As part of its commitment to alternative and renewable fuel technologies, the CEC administers a funding program known as the Alternative and Renewable Fuel and Vehicle Technology Program (ARFVTP). This program represents the largest initiative of its kind in the United States and includes provisions

332

A. Safari et al.

specifically aimed at advancing hydrogen-related projects. The CEC collaborates with various stakeholders to co-fund the establishment of modern hydrogen stations, which are crucial for fueling hydrogen-powered vehicles. The commission’s strategic planning for hydrogen infrastructure prioritizes early adopter cluster communities, ensuring that these regions have access to the necessary infrastructure for adopting and utilizing hydrogen as a clean energy source. The CEC’s support for hydrogen infrastructure is further strengthened by the proactive involvement of seven prominent automakers. These automakers have made commitments to roll out a specified number of vehicles, including transit buses, to sustain and promote the development of future hydrogen fueling stations. To incentivize the construction of efficient and cost-effective hydrogen stations, the CEC employs a performancebased approach. This encourages the establishment of stations with lower costs, faster installation times, and higher capacity, exceeding the state’s requirement of obtaining 33.3% of hydrogen transportation fuel from renewable sources. By doing so, the program enhances the business case for hydrogen infrastructure and optimizes the use of public funds to ensure the success of California’s hydrogen fueling network. Establishing hydrogen refueling stations and deploying fuel cell electric vehicles are pivotal measures taken to foster the adoption of hydrogen as a viable transportation fuel in California [20]. Additionally, in [21], the assessment of hydrogen energy storage and demand response systems’ potential was investigated to engage in specific wholesale electricity markets in California based on data from 2012. The utilization of hydrogen systems may offer a favorable value proposition within present market conditions. The study yielded three key observations: • The production and sale of hydrogen proved significantly more advantageous in California electricity markets than producing and storing hydrogen for subsequent electricity generation. Therefore, it is advisable for systems to prioritize hydrogen production and sales while opportunistically providing ancillary services and engaging in arbitrage. • Establishing a closer integration with electricity markets results in higher revenue generation. • Expanding storage capacity beyond diurnal shifting requirements does not enhance competitiveness in California wholesale energy markets.

13.6

Los Angeles Green Hydrogen Power-to-Gas Project

The Los Angeles Green Hydrogen Power-to-Gas Project initiative aims to produce and utilize green hydrogen as a renewable energy source within the Los Angeles region. This renewable hydrogen is generated through electrolysis, which involves separating water into hydrogen and oxygen utilizing electricity from renewable sources like solar or wind power. The primary objective of this project is to establish a power-to-gas system capable of storing surplus renewable energy generated during

13

Hydrogen-Combined Smart Electrical Power Systems: An Overview of. . .

333

peak production periods and converting it into hydrogen gas. This stored hydrogen can subsequently be employed for diverse applications, including fuel cell vehicles, energy storage, and industrial processes. By converting renewable energy into hydrogen gas, the project seeks to address the intermittent nature associated with renewable energy sources. When renewable energy generation surpasses demand, the excess energy can be utilized for hydrogen production, which can then be stored and utilized during low renewable energy production periods. Furthermore, the Los Angeles Green Hydrogen Power-to-Gas Project is an integral part of the broader efforts undertaken by the city to transition towards clean and sustainable energy sources, while also reducing greenhouse gas emissions. It aligns with the ambitious climate goals set forth by the state of California, which aim to achieve carbon neutrality by the year 2045. This project not only has the potential to contribute to the decarbonization of the transportation sector, but it also facilitates the integration of renewable energy into the existing energy infrastructure. By doing so, the project represents a significant stride towards establishing a more sustainable and resilient energy system in Los Angeles, while serving as a model for other regions interested in adopting green hydrogen technologies.

13.7

Advanced Clean Energy Storage (ACES)

The Advanced Clean Energy Storage (ACES) project, as the world’s biggest energy storage [22], represents a significant and innovative endeavor to establish a largescale green hydrogen production facility in the United States. With a primary focus on advancing clean energy storage solutions, the project seeks to harness the potential of green hydrogen for power generation and transportation sectors. In June 2022, the DoE approved a loan guarantee of $504.4 million to support the financing of the Advanced Clean Energy Storage project. This project aims to establish a clean hydrogen and energy storage facility capable of providing longterm, seasonal energy storage. Located in Delta, Utah, the facility will employ 220 megawatts of alkaline electrolysis technology and utilize two large salt caverns with a combined storage capacity of 4.5 million barrels to store clean hydrogen. The primary objective of Advanced Clean Energy Storage is to capture excess renewable energy during periods of abundance, convert it into hydrogen, and utilize it as fuel for the Intermountain Power Agency’s (IPA) IPP Renewed Project. ACES leverages a 220-megawatt electrolyte bank with intermittent renewable energy sources to produce hydrogen. The significant innovation lies in the scale of the deployed electrolyzes and the use of salt caverns for hydrogen storage. This combination enables efficient and reliable dispatchable generation of electricity in the future. The Advanced Clean Energy Storage project is anticipated to generate approximately 400 construction jobs during the implementation phase and support 25 ongoing operations jobs. One of the key climate benefits of the Advanced Clean Energy Storage project is its potential contribution to grid stabilization and the reduction of renewable energy

334

A. Safari et al.

curtailment. By utilizing hydrogen for long-term storage, the project helps address the intermittency challenges of renewable energy sources. The stored hydrogen will be utilized as fuel for a hybrid 840 MW combined cycle gas turbine (CCGT) power plant, which is intended to replace a retiring 1800 MW coal-fired power plant. The project is estimated to help prevent approximately 126,517 metric tons of carbon dioxide emissions annually. This reduction is based on a comparison of the emission profiles between the IPP turbines fueled by 100% natural gas and those fueled by a blend of 70% natural gas and 30% hydrogen. At the core of the ACES project is the construction and operation of an advanced green hydrogen production facility, employing cutting-edge electrolysis technologies. Electrolysis, utilizing renewable energy sources, enables the efficient splitting of water molecules into hydrogen and oxygen. Through the utilization of state-of-the-art electrolysis techniques, the facility aims to achieve optimal efficiency and sustainability in the production of green hydrogen. Integration with the power generation sector is a key objective within the ACES project. The produced green hydrogen assumes a pivotal role as a reliable and dispatchable fuel source for various power generation applications. By leveraging hydrogen’s inherent flexibility and scalability, the project aims to enhance grid reliability and stability, particularly when integrating intermittent renewable energy sources. This integration has the potential to mitigate the challenges associated with the intermittent nature of renewable energy, thus facilitating a smoother and more dependable transition to a renewable energy-based power system. In addition to power generation, the ACES project recognizes the viability of green hydrogen as a clean fuel for transportation. Hydrogen fuel cell vehicles (FCVs) present an emission-free mobility solution, with hydrogen serving as the fuel source and water vapor as the sole byproduct. By establishing a large-scale green hydrogen production facility, the project seeks to foster the growth of hydrogen-powered transportation, contributing to the decarbonization objectives of the transportation sector. A central aspect of the ACES project lies in its alignment with overarching environmental goals, namely the reduction of greenhouse gas emissions. Green hydrogen, produced through electrolysis utilizing renewable energy sources, exhibits a significantly lower carbon footprint compared to conventional hydrogen production methods such as steam methane reforming. The emphasis placed on large-scale green hydrogen production and utilization within the ACES project serves as a tangible contribution toward the development of a sustainable and low-carbon energy landscape. While the ACES project holds immense promise, it also faces notable challenges associated with large-scale green hydrogen production and integration. Key considerations include the availability and cost-effectiveness of renewable energy sources, the scalability of electrolysis technologies, and the development of an extensive hydrogen infrastructure. Nonetheless, the ACES project represents a pioneering effort to address these challenges head-on and unlock the vast potential of green hydrogen as a clean and versatile energy carrier. Looking ahead, the successful implementation of the ACES project holds significant implications for further advancements in green hydrogen technologies, cost reductions,

13

Hydrogen-Combined Smart Electrical Power Systems: An Overview of. . .

335

Fig. 13.4 Overall framework of ACES

and broader deployment. Beyond its immediate impact, the ACES project is an exemplary model for regions and countries aspiring to realize a sustainable and carbon-neutral energy future. Consequently, ACES is conceptualized in Fig. 13.4. According to Fig. 13.4, the invaluable insights gained from the ACES project’s endeavors in large-scale green hydrogen production and integration are poised to shape the trajectory of the clean energy landscape on a global scale.

13.8 Utah Advanced Clean Energy Storage (UACES) The Utah Advanced Clean Energy Storage (UACES) project represents a cuttingedge endeavor to seamlessly integrate wind and solar power generation with hydrogen production and storage. The primary objective of this project is to establish flexible and dispatchable energy resources that can effectively support the electrical grid, thereby contributing to grid stability, reducing curtailment of renewable energy, and enhancing overall system efficiency. The main part of the UACES project lies in the sophisticated integration of wind and solar power generation with advanced electrolysis-based hydrogen production. Leveraging Utah’s abundant renewable energy resources, the project employs stateof-the-art electrolysis technologies to convert surplus electricity from wind and solar farms into hydrogen. This hydrogen is then intelligently stored for subsequent deployment during periods of peak demand or when renewable energy generation is limited. By providing a dependable and dispatchable energy resource, the UACES project addresses the intermittency challenge associated with renewable sources, facilitating the optimal integration of renewable energy into the grid. The UACES project represents a pioneering approach in clean energy storage and grid support, offering valuable insights and advancements in hydrogen technologies. By successfully demonstrating the feasibility and effectiveness of integrating wind and solar power with hydrogen production and storage, the project sets a precedent

336

A. Safari et al.

for future endeavors in clean energy systems and sustainable grid operations. The implications of the UACES project extend beyond its immediate impact on the electrical grid. Its success can contribute to developing innovative energy storage and grid flexibility solutions, fostering the broader adoption of renewable energy sources and facilitating the transition to a low-carbon and sustainable energy landscape. Finally, the Utah Advanced Clean Energy Storage (UACES) project stands as a pioneering initiative at the intersection of wind and solar power generation, hydrogen production, and storage. With its focus on grid flexibility and dispatchability, the project holds significant promise for enhancing the stability, reliability, and efficiency of the electrical grid. Through its innovative approach and research contributions, the UACES project sets the stage for advancements in clean energy storage and grid integration. It plays a vital role in the ongoing transition towards a sustainable and decarbonized energy future.

13.9

New York Offshore Wind-to-Hydrogen Project

The New York Offshore Wind-to-Hydrogen Project is an innovative and pioneering initiative that seeks to harness offshore wind energy’s immense potential for producing green hydrogen. With a focus on establishing a sustainable and renewable pathway for hydrogen generation, this project aims to unlock various applications across power generation, transportation, and industrial processes. Central to the project is the strategic utilization of the abundant offshore wind resources along the coast of New York. By deploying offshore wind farms equipped with advanced wind turbines, the project aims to capture and harness clean and renewable energy from the wind. This energy is then channeled into electrolysis, a process that involves the splitting of water molecules into hydrogen and oxygen. The produced hydrogen is meticulously captured and stored, ensuring its availability for future utilization. The significance of the generated green hydrogen lies in its remarkable versatility and potential for diverse applications. Within the power generation domain, the project envisions the utilization of hydrogen in fuel cells, enabling electricity production with zero carbon emissions. This transformative approach promises to decarbonize the electricity sector and significantly reduce greenhouse gas emissions. Furthermore, the New York Offshore Wind-to-Hydrogen Project embraces the vision of employing green hydrogen as a fuel for various modes of transportation. Hydrogen fuel cell vehicles, with their emission of only water vapor as a byproduct, offer a compelling alternative to conventional combustion engine vehicles. By integrating green hydrogen as a transportation fuel, the project aims to curtail reliance on fossil fuels and mitigate the environmental consequences of traditional transportation systems. In addition to its impact on the energy and transportation sectors, green hydrogen holds immense potential for transforming various industrial processes. The project

13

Hydrogen-Combined Smart Electrical Power Systems: An Overview of. . .

337

foresees the replacement of fossil-based feedstocks with green hydrogen in the production of chemicals, fertilizers, and synthetic fuels. By catalyzing this transition, the project can substantially contribute to reducing carbon emissions in these sectors, fostering a more sustainable and environmentally friendly industrial landscape. The New York Offshore Wind-to-Hydrogen Project signifies a stride towards effectively harnessing offshore wind energy to produce green hydrogen. Through its diverse applications across power generation, transportation, and industrial processes, the project serves as a catalyst for driving the transition towards a cleaner and more sustainable energy system. By capitalizing on the inherent power of offshore wind and unlocking the myriad benefits of green hydrogen, this project plays a pivotal role in advancing decarbonization efforts and paving the way for a greener future.

13.10

Texas Gulf Coast Hydrogen Hub

The establishment of the Texas Gulf Coast Hydrogen Hub signifies an initiative aimed at harnessing the vast potential of the region’s abundant renewable resources and well-established industrial infrastructure to foster a dynamic and sustainable hydrogen economy. This ambitious project is focused on accelerating the growth and widespread adoption of hydrogen as a pivotal energy carrier, strongly emphasizing both environmental sustainability and economic progress. The primary objective of the Texas Gulf Coast Hydrogen Hub is to utilize the considerable renewable resources available in the region, including solar and wind energy, to facilitate the production of green hydrogen through advanced electrolysis processes. By employing electricity generated from these renewable sources, the project seeks to power sophisticated electrolyzers capable of splitting water molecules into hydrogen and oxygen. The resulting green hydrogen can then be efficiently stored, transported, and effectively utilized as a versatile and eco-friendly energy resource across a broad spectrum of applications and sectors. A significant strength of the Texas Gulf Coast Hydrogen Hub lies in the region’s well-established and resilient industrial infrastructure, comprising refineries, chemical plants, and existing hydrogen production facilities. By seamlessly integrating with this robust infrastructure, the project aims to leverage the available resources, facilities, and expertise, thereby enabling a rapid and scalable expansion of hydrogen production, distribution, and utilization. This harmonious integration presents compelling opportunities for synergy, as the generated hydrogen can be readily integrated into the region’s industrial processes, either as a substitute or complement to conventional fossil fuels, resulting in substantial reductions in carbon emissions. The establishment of the Texas Gulf Coast Hydrogen Hub is expected to yield a plethora of benefits. Firstly, significant contributions will be made towards the reduction of greenhouse gas emissions as fossil fuels are displaced by clean hydrogen in diverse industrial processes, including refining, chemical production, and manufacturing. This transition seamlessly aligns with sustainability objectives and regulatory imperatives aimed at curbing carbon emissions. Secondly, substantial economic growth

338

A. Safari et al.

and job creation within the region are anticipated as a result of the hydrogen hub. The heightened demand for hydrogen, coupled with the requisite infrastructure development, will foster fresh investments, technological advancements, and employment opportunities across the renewable energy sector, as well as in industries reliant on hydrogen as a feedstock or energy source. The Texas Gulf Coast Hydrogen Hub also holds promise in bolstering energy security and enhancing grid stability. Hydrogen, with its inherent flexibility, can serve as a versatile energy resource, offering valuable capabilities such as backup power, grid-balancing services, and long-term energy storage. Consequently, the integration of hydrogen enhances the resilience and reliability of the regional electrical grid, ensuring a stable and secure energy supply. Furthermore, the project is designed to foster robust collaboration among key stakeholders, including entities, industry players, research institutions, and community organizations. This collaborative approach ensures the alignment of efforts, facilitates knowledge sharing, and promotes coordinated actions to effectively address technical, regulatory, and market challenges associated with the widespread adoption of hydrogen technologies. The Texas Gulf Coast Hydrogen Hub represents a transformative and audacious endeavor, capitalizing on the region’s abundant renewable resources and wellestablished industrial infrastructure. By leveraging these inherent strengths, the project aims to accelerate the growth and widespread adoption of a sustainable hydrogen economy, propelling the transition towards a low-carbon future while fostering economic development and resilience. Also, intelligent control methodologies, such as metaheuristic algorithms to have offshore adjustments, quantum intelligent models, and their combination with metaverse, and optical wireless communication can be implemented to have better efficiency on the hydrogen-incorporated smart electrical power systems. The application of the metaverse and quantum technology in hydrogen-incorporated smart grids offers transformative potential for enhancing energy systems. With its immersive virtual environments, the metaverse enables real-time monitoring and optimization of grid components, facilitating efficient management and decisionmaking. On the other hand, Quantum technology utilizes smart grids with advanced computational capabilities, enabling optimal energy scheduling, resource allocation, and secure communication. Quantum sensors and communication technologies further enhance the monitoring and security of hydrogen infrastructure. By integrating the metaverse and quantum technology, hydrogen-incorporated smart grids can achieve higher efficiency, resilience, and intelligence, contributing to a sustainable and advanced energy landscape [21, 22].

13.11

Conclusion

A comprehensive overview of various projects and initiatives in the United States pertaining to green hydrogen technologies and their integration into smart electrical power systems has been provided in this chapter. The significance of green hydrogen

13

Hydrogen-Combined Smart Electrical Power Systems: An Overview of. . .

339

and its role in sustainable energy systems was initially discussed. Subsequently, the involvement of the National Renewable Energy Laboratory (NREL) in advancing green hydrogen technologies, including production, storage, fuel cells, and integrated energy systems, was highlighted. Furthermore, regional initiatives such as the Western Green Hydrogen Initiative and the Gulf Coast Clean Energy Application Center were examined, showcasing collaborative efforts among stakeholders to expedite the deployment of green hydrogen infrastructure. The commitment of states like Hawaii and California to renewable energy integration and the adoption of green hydrogen in transportation was underscored by the Hawaii Hydrogen Initiative and the California Hydrogen Infrastructure Project, respectively. The Los Angeles Green Hydrogen Power-to-Gas Project exemplified converting excess renewable electricity into hydrogen and its utilization in the natural gas pipeline system. Furthermore, the Advanced Clean Energy Storage (ACES) Project showcased the development of a large-scale green hydrogen production facility, focusing on power generation and transportation. The Hawaii Wind-to-Hydrogen Project elucidated surplus wind power’s utilization for hydrogen production, storage, and subsequent use in various applications. The integration of wind and solar power with hydrogen production and storage, as exemplified by the Utah Advanced Clean Energy Storage (UACES) Project, supported the flexibility and reliability of the electrical grid. The New York Offshore Wind-to-Hydrogen Project demonstrated the utilization of offshore wind energy for green hydrogen production, offering diverse applications in power generation, transportation, and industrial processes. Lastly, the establishment of the Texas Gulf Coast Hydrogen Hub leveraged the region’s industrial infrastructure and renewable resources to accelerate the growth of a hydrogen economy. Collectively, these projects signify significant progress in advancing green hydrogen technologies and their integration into smart electrical power systems. The integration of renewable energy sources, innovative storage solutions, and diverse applications of green hydrogen contribute to the development of a sustainable and low-carbon energy landscape. Prospects include further research, collaboration, and investment to propel the growth and adoption of green hydrogen, ultimately paving the way for a cleaner and more sustainable energy future.

References 1. Hayati, M. M., Aminlou, A., Zare, K., & Abapour, M. (2023). A two-stage stochastic optimization scheduling approach for integrating renewable energy sources and deferrable demand in the spinning reserve market. In 2023 8th International Conference on Technology and Energy Management (ICTEM) (pp. 1–7). IEEE. 2. Aminlou, A., Hayati, M., & Zare, K. (2023). ADMM-based fully decentralized Peer to Peer energy trading considering a shared CAES in a local community. Journal of Energy Management and Technology. 3. Kim, C., Cho, S. H., Cho, S. M., Na, Y., Kim, S., & Kim, D. K. (2023). Review of hydrogen infrastructure: The current status and roll-out strategy. International Journal of Hydrogen Energy, 48(5), 1701–1716.

340

A. Safari et al.

4. Ogden, J. M. (1999). Prospects for building a hydrogen energy infrastructure. Annual Review of Energy and the Environment, 24(1), 227–279. 5. Dunn, S. (2002). Hydrogen futures: Toward a sustainable energy system. International Journal of Hydrogen Energy, 27(3), 235–264. 6. Mohideen, M. M., Ramakrishna, S., Prabu, S., & Liu, Y. (2021). Advancing green energy solution with the impetus of COVID-19 pandemic. Journal of Energy Chemistry, 59, 688–705. 7. Mansour-Saatloo, A., et al. (2021). Robust decentralized optimization of Multi-Microgrids integrated with Power-to-X technologies. Applied Energy, 304, 117635. 8. Ifaei, P., Esfehankalateh, A. T., Ghobadi, F., Mohammadi-Ivatloo, B., & Yoo, C. (2023). Systematic review and cutting-edge applications of prominent heuristic optimizers in sustainable energies. Journal of Cleaner Production, 137632. 9. Sadeghian, O., Oshnoei, A., Mohammadi-Ivatloo, B., Vahidinasab, V., & Anvari-Moghaddam, A. (2022). A comprehensive review on electric vehicles smart charging: Solutions, strategies, technologies, and challenges. Journal of Energy Storage, 54, 105241. 10. U. UNDP. WEC, op. cit, note 9. 11. Lasemi, M. A., Arabkoohsar, A., Hajizadeh, A., & Mohammadi-Ivatloo, B. (2022). A comprehensive review on optimization challenges of smart energy hubs under uncertainty factors. Renewable and Sustainable Energy Reviews, 160, 112320. 12. Prabakar, K., et al. (2023). Interconnection and interoperability requirements of hydrogen assets to enable grid integration. National Renewable Energy Lab (NREL). 13. Hydrogen Production and Delivery. https://www.nrel.gov/hydrogen/hydrogen-productiondelivery.html. Accessed 1 May 2023. 14. Hydrogen & Fuel Cells. https://www.nrel.gov/hydrogen/. Accessed 1 May 2023. 15. Dillingham, G. (2013). Gulf coast clean energy application center. Houston Advanced Research Center. 16. Rocheleau, R. E., et al. (2009). Hawaii energy and environmental technologies (HEET) initiative. Hawaii University. 17. Rocheleau, R. E., et al. (2016). Hawaii energy and environmental technologies initiative 2010 (HEET10). Hawaii Natural Energy Institute Honolulu. 18. Wipke, K., Sprik, S., Kurtz, J., Ramsden, T., Ainscough, C., & Saur, G. (2011). Status of US FCEV and infrastructure learning demonstration project (presentation). National Renewable Energy Lab (NREL). 19. Heydorn, E. C. (2013). California hydrogen infrastructure project. Air Products and Chemicals. 20. Muench, T. R. (2012). California’s hydrogen infrastructure funding program. ECS Transactions, 42(1), 91. 21. Eichman, J., Townsend, A., & Melaina, M. (2016). Economic assessment of hydrogen technologies participating in California electricity markets. National Renewable Energy Lab (NREL). 22. Parnell, J. (2019). ‘World’s biggest’ energy storage project planned for Utah. https://www. forbes.com/sites/johnparnell/2019/05/30/worlds-biggest-energy-storage-project-planned-forutah/. Accessed 7 Jun 2023.

Chapter 14

An Overview of the Pilot Hydrogen Projects Maryam Shahbazitabar

14.1

and Hamdi Abdi

Status Quo, Challenges, and Outlook

The continuous growth of greenhouse gas (GHG) emissions has forced governments, companies, operators, and planners to adopt appropriate policies and decisions to deal with climate change. In this framework, increasing the renewable energy sources (RES) contribution in the final energy utilization, plays a significant role in all energy-related planning studies. Changing the current energy systems to sustainable ones with minimal pollution is achieved by using various techniques such as energy efficiency, and combining different energy systems and sectors. Hydrogen is a clean and stable energy carrier that releases only water as a byproduct with no carbon dioxide emission and simultaneously converts it into electricity. Its energy density is about twice as much as that of conventional solid fuel. It can easily be transported from the production site to the utilization/storage centers using pipeline networks or suitable transport trucks. These characteristics make hydrogen an attractive and suitable energy carrier in today’s energy systems. In this regard, the power-to-hydrogen concept for supply of storage, demand management, and the combination of various elements between different sectors has attracted considerable attention. In contrast, the system may counter with some challenges with hydrogen entering. The most important challenges related to hydrogen technology can be mentioned as follows: • Technical, infrastructural, economic, and safety issues related to hydrogen energy • Limited hydrogen storage capacities • High prices for low demand of hydrogen

M. Shahbazitabar · H. Abdi (✉) Electrical Engineering Department, Faculty of Engineering, Razi University, Kermanshah, Iran e-mail: [email protected]; [email protected] © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 V. Vahidinasab et al. (eds.), Green Hydrogen in Power Systems, Green Energy and Technology, https://doi.org/10.1007/978-3-031-52429-5_14

341

342

M. Shahbazitabar and H. Abdi

• The very high cost of developing infrastructure related to hydrogen energy applications • The complex and very expensive tools for hydrogen distribution and diffusion systems • The higher total cost of hydrogen distribution technologies compared to conventional fuels • Problems of large volumes of hydrogen as fuel • Need for a cost-effective and economical strategy for the transition to the hydrogen distribution system • Producing large amounts of carbon dioxide from the existing technologies in the field of hydrogen production • Lack of infrastructure related to the use of hydrogen • Lack of necessary rules and standards • Lack of public education • Dangers related to the use of hydrogen

14.2

Introduction

The increase in demand for electricity consumption besides the importance of concentrating on environmental pollutants, and greenhouse gas emissions reduction has caused traditional power systems that rely on fossil fuels to be pushed towards the use of clean energy in electrical energy production and make a decision to deal with climate changes [1]. Hydrogen production methods for non-carbon energy uses, including aviation, and shipping, have been investigated [2]. In [3] some very limited cases of hydrogen use in the Nexus are reported. In [4], the importance of using hydrogen fuel cells (HFC) as one of the components of the energy storage system in microgrids is emphasized. Also, ref. [5] addressed the modeling of some simple multiple inputs and multiple outputs devices including hydrogen in output. The gravitational search algorithm (GSA) is introduced to solve multi-objective energy hub economic dispatch (EHED) problems [6]. Since 2010, an increasing range of policies and laws have enhanced energy efficiency and reduced the cost of renewable energies such as wind, solar, and batteries by up to 85 percent. Also, due to the importance of environmental issues, the United Nations announced the amount of greenhouse gas emissions worldwide should be reduced to 43% by 2030 [7]. The COP21 Paris Agreement between 175 signatories is the basis of national energy policies in reducing CO2 emissions and includes the following provisions [8]: • Limiting global warming to 1.5 or 2 °C above the pre-industrial levels • Global greenhouse gas emissions peak regulation • Reduction of greenhouse gas emissions

14

An Overview of the Pilot Hydrogen Projects

343

The minimum climate goals in Europe are: • Cut in GHG emissions in comparison with 1990 level, 20% (2020), 40% (2030), and 80% (2050) • Total energy consumption from RES, 20% (2020) and 32% (2030) • Increase in energy efficiency, 20% (2020) and 27% (2030) Among the most common fuels, hydrogen which has the lowest volumetric energy density (0.01079 MJ/L in atmospheric conditions), is well able to integrate RES and a strong connection between different energy sectors and omit the climate change factors [9]. On the other hand, hydrogen is one of the most suitable and affordable alternatives for long-term storage of renewable energy, which attracts more attention as a potential source and effective energy carrier for clean and sustainable energy. Basically, hydrogen will have a very crucial role in energy transfer at the global level, and its use will lead to a noticeable GHG emission reduction in the next decades [10]. It is estimated that the production of electric energy using RES and hydrogen will cause a 60% reduction in GHG emissions [11]. On the other hand, the transportation industry plays a very important and decisive role in the consumption of various energy carriers and the emission of dangerous types of greenhouse gases. The International Energy Agency (IEA) announced that the contribution of the transportation industry to global energy consumption in 2017 was about 29% and its share in global carbon dioxide emission in 2016 was nearly 25%. It consumes nearly 30 million tons of hydrogen and needs to modernize and decarbonize to reduce the relevant emissions by 90% in 2050 [12]. The United States as a pioneer in fuel cells and hydrogen energy long-term policies, launched 5899 fuel cell vehicles in 2018 [13]. Consequently, hydrogen is known as a suitable choice for transportation, industry, and buildings because of its ability to integrate with renewable energy and omit fossil fuel uses, and carbon footprint [14]. The Fuel Cell and Hydrogen Energy Association in the United States, declared the eighth of October as the National Hydrogen and Fuel Cell Day in 2015, and also planned the hydrogen production from renewable resources by at least 10% of its by 2023, and between 20% and 40% by 2028 [15]. Due to the importance of the mentioned things, hydrogen will soon become a definite and appropriate choice in the energy sector. Therefore, all concerted efforts of governments, industries, and investors are needed for the hydrogen value chain to expand on a worldwide scale as shown in Fig. 14.1, briefly.

14.3

Definitions and Classification of Hydrogen

As the simplest element on earth, consisting of one proton and one electron, hydrogen usually does not exist by itself in nature and must be produced from containing compounds. Hydrogen can act as an energy carrier and store and deliver usable energy. Even though hydrogen is produced from non-renewable raw materials (gray hydrogen) and has significant amounts of carbon, it can become one of the cleanest and most durable fuels made from renewable sources, and even by making

344

M. Shahbazitabar and H. Abdi

Fig. 14.1 Graphical picture

significant investments in related research development, and production assets, to become an emission-neutral (green hydrogen). Of course, the dangerous characteristics that show it as an acute risk of leakage as an invisible flammable gas should not be forgotten. This object provides suitable opportunities for the decarbonization of various industrial and residential buildings as well as transportation. Of course, providing the necessary flexibility in various applications to the power grid through fuel cell technology is also a very important issue that has been favored by planners and researchers. It has been discussed about the existing gaps in hydrogen applications in industries such as aerospace, electronics, metallurgy, pharmaceuticals, etc., in [9]. In [10] the current bottlenecks and obstacles in the development and commercialization of fuel cell electric vehicles (FCEV) are discussed. In [16], various factors affecting the adoption of hydrogen in industries and industrial processes as a carbon-free source have been investigated. The following three phases of production, purification, and compression are explained in detail.

14.3.1

Production

The global production of hydrogen will be nearly 100 million tons by 2020 which about 96% be extracted from fossil fuels, divided by: 48% from the cracking of fossil fuels (Steam methane reforming -SMR), 30% from alcohol cracking, and the coke oven gas share is 18% [17]. Methane reformation technology such as Auto Thermal Reformation (ATR), or Steam Methane Reforming (SMR) is a typical technology for the production of grey hydrogen. The SMR is responsible for approximately 75% of all global hydrogen production. In SMR heat and pressure are used to convert the methane in natural gas to carbon dioxide, and hydrogen [18]. The challenges and

14

An Overview of the Pilot Hydrogen Projects

345

Fig. 14.2 Common methods for production Production Methods

Electrolysis Biomass Gasification Steam Methane Reforming Photoelectrolysis Aqueous Phase Reforming Pyrolysis and Co-pyrolysis

drawbacks of gray hydrogen methods are the production of carbon dioxide and increasing global warming. For this reason, the power-to-gas approach based on water electrolysis and the use of energy obtained from renewable sources (geothermal, hydroelectric, solar, and wave) is used as an environmentally friendly method. Renewable energy integration for combined heat and power production is investigated with a focus on solar, geothermal, and minimizing curtailment in windintegrated systems [19]. A comprehensive review and classification of thermal energy storage technologies applied in the built environment is investigated in [20]. As a result of passing an electric current through water, hydrogen, and oxygen are released. This is the most suitable and effective method in hydrogen production because its carbon content is zero. Electrolyzed water accounts for a small proportion of electrolyzed water which counts about 4% because its cost is more than twice in rather of fossil fuels [21]. Hydrogen production with electrolyzed water method can compensate for the instability of renewable energy sources and develop distributed economic production by reducing the cost [22]. The electrolyzer will produce clean and green hydrogen by electrolyzing water with an electric current and dividing it into dioxygen (O2) and dihydrogen (H2) [23]. Figure 14.2 shows common production methods. The carbon emission of green hydrogen is less than 36.4 g CO2/MJ [7]. The simple definition of the process is that excess wind or solar energy is used to generate hydrogen by water electrolysis when the load is less than electricity generation. In the next step, hydrogen is stored as a compressed gas or in metal hydrides. Then the stored hydrogen can be used to generate electricity by using fuel cells if electricity generated by the solar and/or wind is lower than electricity demand. Industrial electrolyzers that are used to generate green hydrogen must meet the following significant criteria [24]: • The ability to absorb maximum additional electricity generation from renewable energy resources [25] • Broad-range applications • Short time of response Numerous advances in this field in the last decade indicate that water electrolysis is at the forefront of flexible energy storage technology on a global scale.

346

M. Shahbazitabar and H. Abdi

Hydrogen Purification Methods

Metal Hydride Separation

Pressure Swing Adsorption (PSA)

Polymer Membrane Diffusion

Cryogenic Separation

Catalytic Purification

Palladium Membrane Diffusion

Solid Polymer Electrolyte cell

Fig. 14.3 Purification methods for hydrogen production

14.3.2

Purification

In the twenty-first century, hydrogen is considered the best replacement for coal, petroleum, and natural gas because it has various clean sources with low carbon footprint, flexibility, and high efficiency. Nowadays, hydrogen fuel cell vehicles have been welcomed and gradually replaced traditional fossil fuel vehicles because of their zero-carbon emission. Crude hydrogen is produced in different methods and cannot be directly used before purifying. In reality, hydrogen purification is a vital step from production to utilization and a fuel cell can operate effectively only with a hydrogen quality guarantee [26]. Figure 14.3 shows chemical and physical purification methods.

14.3.3

Compression

Hydrogen storage is one of the most important challenges due to its lightness. There are various ways for hydrogen storage such as: stored as compressed gas, liquid at very low temperatures, and stored through physisorption and chemisorption using metal hydrides. The best and simplest way to store hydrogen is compression (mechanical or non-mechanical), which allows for a wider use of hydrogen as a sustainable and renewable fuel. Figure 14.4 depicts the compression technology of hydrogen.

14.4

Background and Motivations

Replacing fossil fuels with hydrogen will be one of the most significant future trends in efficient and clean energy generation. Hydrogen extraction from RES, on the one hand, overcomes some unreliable parameters such as low energy density and weak sustainability of RES; on the other hand, it can improve the grid connection insecurity and some disadvantages of

14

An Overview of the Pilot Hydrogen Projects

347

Reciprocating Piston Compressors

Mechanical Compression

Linear Compressors Liquid Compressors Diaphragm Compressors

Compression Technologies

Adsorption Compressors

Non-Mechanical Compression

Metal Hydride Compressors Cryogenic Compression Electrochemical Compressors

Fig. 14.4 Compression technologies of hydrogen

Fig. 14.5 Two main categories of hydrogen technologies

conventional batteries which cannot store electrical energy for a long period. The hydrogen procedure may be categorized into two main phases, as they are depicted in Fig. 14.5. Due to the aforementioned environmental and economic reasons, many major governments in the world, such as the United States, Japan, and Germany, are seeking to launch, use, and develop fuel cells and hydrogen energy. These developed countries have been able to bring the development, research, and industrialization of fuel cells and hydrogen energy to the level of national strategy. Figure 14.6 shows the first pioneers of the world in each continent. Table 14.1 depicts the hierarchical improvements of hydrogen based on year and country.

14.5 Conclusion This chapter reviewed some pilot hydrogen projects. For this purpose, the definitions and classification of Hydrogen in production, purification, and compression were detailed. Then a brief background and motivations were addressed in terms of different continents. Finally, the hierarchical improvements of the hydrogen pilot project based on year and country were presented.

348

M. Shahbazitabar and H. Abdi

America

●United States: The poineer in adopting fuel cells and hydrogen energy as an energy procedure with an investment cost of more than 1.6 billion$.

Asia

●Japan: A leader in the realization of the hydrogen energy association, with the graet number of patents in the world hydrogen and fuel cell energy and the commercialization of fuel cell vehicles and large-scale domestic cogeneration systems. ●South Korea: is focusing on the large fuel cell power plants expansion.

Australia

●Australia: Presenting 57 strategic actions in the field of research and development of the regulations of the Minister of Transport as a world leader in the production and export of hydrogen.

Europe

●European Union: Introducing hydrogen as a guarantor of security and energy transformation by spending a budget of 665 million euros in 2014-2020 and setting up about 200 hydrogen refuelling stations. Germany is at the forefront of Europe by creating suitable infrastructure for the development of fuel cell vehicles, hydrogen refueling stations and combined heat and power plant.

Africa

●South Africa: Generate low-cost green hydrogen relying the wind and solar power potential. With energy costs making up more than 60% of hydrogen generation costs, bountiful low-cost renewable resources energy is a large competitive advantage.

Fig. 14.6 The first pioneers of the world in each continent

Table 14.1 The hierarchical improvements of hydrogen based on year and country Year 2001

Country United States

2002

United States

2004

India

2009

European Union

2013

Germany

2014

United States

Contribution to hydrogen goals “Integrated Energy System Development Plan” suggestion, besides the micro and smart grid expansion proposed a clean energy supply [27]. Announced a promotion goal to produce hydrogen and develop its application in the “National Hydrogen Energy Development Roadmap” [28]. “National Hydrogen Energy Roadmap” with emphasis on policies, laws, financing, and supporting infrastructure along with research, development, and demonstration programs for the commercialization of various hydrogen energy technologies [29] Regulation of the “Renewable Energy Directive” to the policies and proceedings of using clean and renewable energy by the member countries [7, 30, 31]. Construction of the first hybrid energy plant all over the world with hydrogen energy as an intermediate energy storage. The generator is derived from the hydrogen which is obtained from electrolysis, and the produced electricity continues the electrolysis process to produce hydrogen [32]. Developing a comprehensive energy strategy to develop low-carbon technologies as a basis for the production and use of clean energy and the main role of transporting hydrogen [33]. (continued)

14

An Overview of the Pilot Hydrogen Projects

349

Table 14.1 (continued) Year 2015

Country Germany

2016

European Union Germany China

2017

Japan

Bangladesh

2018

Germany

Australia

2019

France

European Union Russia

South Korea

Japan

Asia-Pacific

Contribution to hydrogen goals Mainz Energy Project: Installation and operation of the world’s largest hydrogen production and storage station to convert renewable energy into hydrogen to effectively reduce the fluctuating nature of renewable energy systems in the grid [34]. Introduced a comprehensive policy and program for the use of clean energy to protect the environment by introducing hydrogen energy as a suitable alternative in 2050 [33]. Revised its hydrogen energy strategic plan, concentrating on the integrated expansion of the industrial chain, and energy supply [33]. Proposed decarbonization large-scale roadmap for generating, storing, transporting, and applying hydrogen, with the “Roadmap for Key Innovation Actions in the Energy Technology Revolution” [28]. The first country to introduce hydrogen as an energy carrier and create competition between hydrogen and natural gas to realize a “hydrogenbased society” [35]. Proposing a combined coastal system of wind, solar, and tidal energy to produce hydrogen in the Patenga region and use hydrogen to reduce grid distortion without producing carbon dioxide [36]. The first demonstration project for a combination of hydrogen and fuel cells with Germany’s hydrogen-powered trains between Buxtehude and Cuxhaven near 100 km length [37]. Defining a strategic plan with a project that includes the use of hydrogen in different sections such as electricity and heat production, storage, industries, and transportation for creating a sustainable economic society. Formulation and development of a hydrogen renewable energy plan and carbon-free production reforms in the industry and creation of an energy network [18]. Proposed two-level hydrogen energy development roadmap: mediumterm (2010–2020) and long-term (2020–2050) [38] Studies of hydrogen production technology in the network when renewable energy has output power below the threshold level. Fluctuation effects in hydrogen production and improving its quality and total efficiency are investigated [33]. The long-term plan for the development of the hydrogen economy calls for a 5% contribution of hydrogen in the national level of energy consumption, creating 420,000 new job opportunities, and the reduction of greenhouse gas emissions [39]. Operation of the first global hydrogen supply chain in the world on the basis of organic liquid hydrogen storage in 2020, which can deliver 210 tons of hydrogen to Japan, annually. Brunei’s wind power resources will be used for hydrogen production and exploitation of low-carbon clean hydrogen in the near future. Production of 45% of the global volume of hydrogen in 2025, will be mainly done by the People’s Republic of China, due to the increase in Europe’s need to import hydrogen. (continued)

350

M. Shahbazitabar and H. Abdi

Table 14.1 (continued) Year 2020

Country Turkey

2021

European Union

2022

South African

Contribution to hydrogen goals Organizing the first meeting of shareholders, officials, and representatives of hydrogen production and gas distribution associations by the Ministry of Energy and Natural Resources of Turkey to provide a road map for transferring hydrogen and mixing hydrogen with natural gas in the distribution network until 2030 [40]. Formulated the “2050 Energy Technology Roadmap,” taking decarbonization as a main goal, and hydrogen extraction as a basic and critical part of the energy system. Hydrogen and fuel cells will become an important driving parameter in the fundamental evolution of the next energy systems. The South African Department of Science and Innovation announced the Hydrogen Society Roadmap (HSRM), as a significant milestone in the launch of South Africa’s hydrogen economy [7, 30]

References 1. Zhang, B., et al. (2021). Progress and prospects of hydrogen production: Opportunities and challenges. Journal of Electronic Science and Technology, 19(2), 100080. 2. Irena, I. (2019). Hydrogen: A renewable energy perspective–Report prepared for the 2nd Hydrogen Energy Ministerial Meeting in Tokyo. Japan. 3. Abdi, H., Shahbazitabar, M., & Mohammadi-Ivatloo, B. (2020). Food, energy and water nexus: A brief review of definitions, research, and challenges. Inventions, 5(4), 56. 4. Abdi, H. (2022). A brief review of microgrid surveys, by focusing on energy management system. Sustainability, 15(1), 284. 5. Beigvand, S. D., Abdi, H., & La Scala, M. (2021). Multicarrier energy systems. In Handbook of energy economics and policy (pp. 433–519). Elsevier. 6. Beigvand, S. D., Abdi, H., & La Scala, M. (2017). A general model for energy hub economic dispatch. Applied Energy, 190, 1090–1111. 7. UN Climate Change Conference (COP21). (2015). Editor. France. https://unfccc.int/processand-meetings/the-paris-agreement 8. Chehade, Z., et al. (2019). Review and analysis of demonstration projects on power-to-X pathways in the world. International Journal of Hydrogen Energy, 44(51), 27637–27655. 9. Okolie, J. A., et al. (2021). Futuristic applications of hydrogen in energy, biorefining, aerospace, pharmaceuticals and metallurgy. International Journal of Hydrogen Energy, 46(13), 8885–8905. 10. Luo, Y., et al. (2021). Development and application of fuel cells in the automobile industry. Journal of Energy Storage, 42, 103124. 11. Egeland-Eriksen, T., Hajizadeh, A., & Sartori, S. (2021). Hydrogen-based systems for integration of renewable energy in power systems: Achievements and perspectives. International Journal of Hydrogen Energy, 46(63), 31963–31983. 12. Zhao, F., et al. (2020). Hydrogen fuel cell vehicle development in China: An industry chain perspective. Energy Technology, 8(11), 2000179. 13. Meng, X., et al. (2021). Status quo of China hydrogen strategy in the field of transportation and international comparisons. International Journal of Hydrogen Energy, 46(57), 28887–28899.

14

An Overview of the Pilot Hydrogen Projects

351

14. Von Colbe, J. B., et al. (2019). Application of hydrides in hydrogen storage and compression: Achievements, outlook and perspectives. International Journal of Hydrogen Energy, 44(15), 7780–7808. 15. Roos, T., & Wright, J. (2020). Powerfuels and green hydrogen (public version). The Council for Scientific and Industrial Research (CSIR). 16. Griffiths, S., et al. (2021). Industrial decarbonization via hydrogen: A critical and systematic review of developments, socio-technical systems and policy options. Energy Research & Social Science, 80, 102208. 17. Younus, T., et al. (2018). Production of hydrogen by steam methane reformation process. In E3S Web of Conferences. EDP Sciences. 18. Birol, F. (2019). The future of hydrogen: Seizing today’s opportunities. IEA report prepared for the G, 20. 19. Bagherian, M. A., & Mehranzamir, K. (2020). A comprehensive review on renewable energy integration for combined heat and power production. Energy Conversion and Management, 224, 113454. 20. Borri, E., Zsembinszki, G., & Cabeza, L. F. (2021). Recent developments of thermal energy storage applications in the built environment: A bibliometric analysis and systematic review. Applied Thermal Engineering, 189, 116666. 21. Li, Y., Chen, J., & Liu, C. (2019). Research progress in hydrogen production technology. Plating & Finishing, 41(10), 22–27. 22. Rashid, M., et al. (2015). Hydrogen production by water electrolysis: A review of alkaline water electrolysis, PEM water electrolysis and high temperature water electrolysis. International Journal of Engineering and Advanced Technology, 4, 2249–8958. 23. Bünger, U., et al. (2014). Power-to-gas (PtG) in transport status quo and perspectives for development. Report to the Federal Ministry of Transport and Digital Infrastructure (BMVI), Germany. 24. Valdés-López, V. F., et al. (2020). Carbon monoxide poisoning and mitigation strategies for polymer electrolyte membrane fuel cells–A review. Progress in Energy and Combustion Science, 79, 100842. 25. Annan-Diab, F., & Molinari, C. (2017). Interdisciplinarity: Practical approach to advancing education for sustainability and for the sustainable development goals. The International Journal of Management Education, 15(2), 73–83. 26. Du, Z., et al. (2021). A review of hydrogen purification technologies for fuel cell vehicles. Catalysts, 11(3), 393. 27. Zou, Y., et al. (2022). Research on the Technology Framework of Integrated Energy System and Recent Progress. in 2022 IEEE 6th Conference on Energy Internet and Energy System Integration (EI2). IEEE. 28. Li, Z., et al. (2020). Development of renewable energy multi-energy complementary hydrogen energy system (A case study in China): A review. Energy Exploration & Exploitation, 38(6), 2099–2127. 29. Pathak, V. V., et al. (2016). Deployment of fermentative biohydrogen production for sustainable economy in Indian scenario: Practical and policy barriers with recent progresses. Current Sustainable/Renewable Energy Reports, 3, 101–107. 30. United Nations of Climate Change. Available from: https://unfccc.int/process-and-meetings/ the-paris-agreement 31. Fleming, R. (2021). Clean or renewable–hydrogen and power-to-gas in EU energy law. Journal of Energy & Natural Resources Law, 39(1), 43–63. 32. Fischer, U. R., et al. (2016). Hydrogen hybrid power plant in Prenzlau, Brandenburg. Hydrogen Science and Engineering: Materials, Processes, Systems and Technology, 44, 1033–1052. 33. Wang, M., et al. (2019). Review of renewable energy-based hydrogen production processes for sustainable energy innovation. Global Energy Interconnection, 2(5), 436–443. 34. Xianxu, H., et al. (2016). Review on key technologies and applications of hydrogen energy storage system. Energy Storage Science and Technology, 5(2), 197.

352

M. Shahbazitabar and H. Abdi

35. D’Ambrogio, E. (2021). Japan's 2050 goal: A carbon-neutral society. 36. Basar, M. A. (2021). Prospect of harnessing solar energy, offshore wind energy and installing liquid air energy storage system to power Chittagong Port and design of a hybrid power system. World Maritime University. 37. Köhler, U. (2019). Examining exoplanetary candidates. DLR Magazine, 162, 26–27. 38. Cuevas, F., Zhang, J., & Latroche, M. (2021). The vision of France, Germany, and the European Union on future hydrogen energy research and innovation. Elsevier LTD on behalf of Chinese Academy of Engineering and Higher Education Press Limited Company 39. Lebrouhi, B., et al. (2022). Global hydrogen development – A technological and geopolitical overview. International Journal of Hydrogen Energy, 47(11), 7016–7048. 40. Uyar, T. S., Beşikci, D., & Sulukan, E. (2020). An urban techno-economic hydrogen penetration scenario analysis for Burdur, Turkey. International Journal of Hydrogen Energy, 45(50), 26545–26558.

Index

A Activated sludge, 167 ADMM optimization, 86, 89, 90, 93, 94, 99 Anaerobic digestion (AD) systems, 179, 182, 186, 188 Anion exchange membrane (AEM), 232, 233

B Biohydrogen production, 14, 115, 116, 120, 137, 156–159, 161, 163, 166–168 Biohydrogen-producing microorganisms, 159 Biomass, 7, 15, 26, 41, 42, 105, 115–123, 136, 137, 150, 156, 157, 162–168, 174–176, 179–186, 191, 193, 196, 197, 227, 233–235, 302, 311, 325 Biomass washout, 160, 162 Bioreactor designs, 156, 168 Blue hydrogen, 133, 147

C California Hydrogen Infrastructure Project (CHIP), 322, 331–332, 339 Carbon capture, 136, 145, 226 Carbon capture utilization and storage (CCUS), 145, 146 Carbon emissions, 1, 5, 6, 14, 29, 35, 59, 85, 89, 144, 149–151, 175, 207, 228, 246, 247, 250, 303, 304, 307, 309, 312, 314, 317, 322, 325, 336, 337, 345 Carbon free community, 142 Carbon neutral, 59, 115, 149, 174, 196, 206, 246, 304, 329, 335

Carbon neutrality, 145–146, 174, 238, 246–249, 256, 301, 309, 314, 330, 333 Catalytic nano-particles, 168 Cell entrapment, 156–158 Centralized GH2 power system, 70 China, 2, 3, 134, 231, 238, 246–250, 256–266, 268, 269, 275, 279, 288, 294, 295, 349 Circular economy, 149, 196 Clean and renewable energy, 336, 348 Clean energy transition, 302 Clean fuels, 10, 25, 26, 115, 121, 174, 175, 196, 301, 322, 334 Climate change, 28, 59, 65, 85, 115, 141, 149–151, 173, 174, 203, 206, 225, 245, 246, 301, 304, 309, 314, 315, 317, 322, 341–343 Climate change mitigation, 52 Continuous stirred tank reactor (CSTR), 155, 159, 160, 162, 163, 166, 167 Cost-effectiveness, 49, 50, 52, 144, 326, 334 Cost reductions, 47, 100, 220, 278, 334

D Decarbonization, 2, 26–28, 33, 34, 52, 203, 228, 234, 238, 239, 249, 250, 256, 260, 264, 271, 301, 302, 304, 307, 322, 333, 334, 337, 344, 349, 350 Decentralized energy systems, 85, 303 Decentralized GH2 power systems, 85, 86 Dibenzyltoluene (DBT), 256, 277

© The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 V. Vahidinasab et al. (eds.), Green Hydrogen in Power Systems, Green Energy and Technology, https://doi.org/10.1007/978-3-031-52429-5

353

354 E Economic analysis, 13, 14 Economic considerations, 37, 52 Economic optimization, 26 Efficient hydrogen utilization, 26, 44–46 Electrolysis, 29, 59, 85, 106, 148, 179, 204, 229, 248, 301, 321, 345 Electrolysis of water, 15, 107, 110, 126, 134, 136, 176, 230–232 Electrolyzed water, 126, 345 Electrolyzers, 7, 10, 13, 67–70, 74, 75, 78, 80, 86, 92, 93, 95, 97, 98, 100, 109–112, 114, 125, 137, 179, 180, 182, 204, 207–213, 218, 219, 221, 306, 309–312, 337, 345 Encapsulation, 156, 158 Energy, 25, 59, 85, 105, 141, 168, 173, 203, 225, 246, 301, 321, 341 Energy analysis, 12, 14, 188–190 Energy consumption, 1, 3, 15, 36, 37, 42, 45, 65, 106, 113, 114, 120, 137, 144, 176, 228, 234, 250, 254, 257–260, 263, 264, 268, 271, 273, 288–292, 343, 349 Energy density, 28, 30, 44, 105, 126, 132, 144, 175, 251, 271, 302, 341, 343, 346 Energy efficiency, 9, 10, 13–15, 26, 32, 34, 36, 62, 143, 145, 176, 179, 186, 188, 191, 246, 301, 307, 341–343 Energy flexibility, 89, 91, 100, 121, 248, 335, 345 Energy flows, 31, 90 Energy landscape, 5, 16, 41, 46, 59, 62, 85, 204, 309, 313, 316, 323, 327, 334–336, 338, 339 Energy management, 26, 31, 32, 35–37, 45, 65, 68, 94, 97 Energy marketing, 8, 37, 38, 46, 60, 62–63, 87, 88, 100, 329, 332 Energy resilience, 60 Energy sectors, 26, 28, 30, 32, 52, 62, 125, 142, 149, 301, 316, 322, 343 Energy security, 7, 8, 27, 52, 85, 184, 206, 226, 311, 338 Energy storage, 25, 59, 89, 107, 144, 175, 203, 249, 305, 325, 342, 345 Energy system integration, 27, 301, 302 Energy transfer, 238, 343 Energy transition and sectors, 142, 252 Energy transitions, 59, 60, 142, 252, 255, 307, 316, 321 Environmental concerns, 150 Environmental impacts, 4, 12, 25, 27, 32, 47, 51, 65, 88, 132, 195, 197, 250, 303

Index Environmental pollutants, 106, 342 Environmental pollution, 9, 106, 122, 228, 257 Exergoeconomic, 13, 177, 178, 188–191 Exergoenvironmental, 13, 177, 178, 188, 190 Exergy, 11–15, 177, 178, 180, 181, 185, 188–191, 193, 196 Exergy analyses, 11–14, 16, 177, 189, 190 Exergy efficiencies, 12–15, 180–182, 185, 188–191, 193

F Flexible energy carrier, 248 Fluidized bed reactor (FBR), 156, 159, 162, 165–167 Fossil fuels, 1–7, 10, 28, 29, 33, 34, 48, 65, 66, 85, 89, 100, 105–107, 121, 125, 135, 137, 141, 143–147, 149, 173, 174, 195, 197, 225, 230, 241, 246–248, 251, 256–259, 263, 271, 272, 283, 285, 288, 290, 291, 305, 306, 312, 314, 317, 322, 330, 331, 336, 337, 342–346 4E analysis, 188, 192 Freshwater, 180, 183, 186 Fuel cell, 27, 68, 86, 112, 150, 180, 204, 236, 248, 306, 322, 343 Future energy systems, 4

G GHG emissions, 28, 47, 48, 59, 85, 125, 133–135, 141, 143, 147, 149, 173, 176, 206, 245, 324, 333, 334, 336, 337, 341–343, 349 Gansu, 249, 250, 256, 260–268, 270, 273–277, 288, 291, 292, 295 Gasification, 7, 11, 26, 41, 42, 107, 115, 119, 120, 123, 133, 136–138, 179, 180, 182–185, 188, 191, 193–197, 229, 234, 325 GH2 standards and codes, 134, 149 Gray hydrogen, 133, 147, 343, 345 Gray Wolf Optimizer (GWO), 217, 218 Greenhouse Gas (GHG), 9, 28, 47, 48, 59, 85, 125, 133–136, 138, 141, 143, 144, 147, 149, 150, 176, 203, 206, 225–228, 239, 240, 245, 301, 306, 314–316, 324, 333, 334, 336, 337, 341–343, 349 Green hydrogen (GH2), 27, 59, 85, 134, 142, 194, 203, 225, 263, 301, 322, 345 Green Hydrogen Technologies, 59, 237–239, 323, 345 Grid balancing, 51, 306, 338

Index

355

Grid stability, 26, 37, 46, 47, 52, 85, 309, 335, 338 Gulf Coast Clean Energy Application Center (GCCEAC), 322, 330, 339

J Jing-Jin-Ji, 247, 250, 256–260, 264, 269, 270, 272, 273, 275, 277, 278, 282–286, 288–290, 293, 295

H Horizon Europe, 312–313 Horizon projects, 301, 316 Hydraulic retention time (HRT), 155–157, 160–166 Hydrogen, 25, 59, 85, 105, 142, 157, 174, 203, 226, 247, 301, 321, 341 Hydrogen applications, 107, 121–126, 147, 179, 325, 339, 344 Hydrogen compression, 127, 128 Hydrogen economy, 4–5, 38, 39, 44, 126, 207, 238, 314, 323, 325, 330, 337–339, 349, 350 Hydrogen energy (HE), 11, 12, 14, 46, 59, 65, 86, 89, 95, 98, 142, 194, 207, 225, 226, 234, 237, 238, 304, 309, 312, 313, 332, 341–343, 347–349 Hydrogen evolution reaction (HER), 232, 234 Hydrogen production, 26, 92, 105, 147, 157, 176, 204, 225, 248, 303, 322, 342, 343 Hydrogen production methods, 5, 35, 107, 108, 113, 114, 118, 120, 137, 175, 176, 179, 225, 232–234, 249, 303, 334, 342 Hydrogen purification, 230, 346 Hydrogen refueling station (HRS), 27, 41, 44, 236, 239, 325, 326, 329, 332 Hydrogen storage technologies, 126–132, 207, 231, 250, 325 Hydrogen tanks, 67, 68, 70, 74, 77–80, 92, 97, 98, 100, 209

L Lignocellulosic materials, 156, 158 Liquid organic hydrogen carrier (LOHC), 41, 250, 253, 254, 278–280 Low-carbon energy systems, 32, 121 Low-carbon future, 28, 41, 59, 60, 85, 321, 323, 338 Low-carbon transition, 41, 59, 150, 321, 324, 336, 338

I Immobilization matrix, 168 Immobilized cell culture system, vi, 155–168 Inner Mongolia, 248–250, 256, 260–268, 270, 273, 275–278, 284, 285, 288, 290–292, 295 Integrated energy systems, 31, 207–218, 322, 324–327, 339, 348 Integration of GH2 and Renewable energy, 203, 322, 329 Intermittent renewable energy sources, 311, 333, 334

M Mass transfer, 156, 157, 162, 165, 166 Microgrids, 1, 96, 204, 205, 207, 208, 210–212, 214, 217, 218, 220–222, 327, 342 Multi-energy system, 10, 11 Multigeneration, 14, 185, 186

N National Renewable Energy Laboratory (NREL), 116, 134, 322, 324–327, 339 Net-zero emission (NZE), 142–145 New York Offshore Wind-to-Hydrogen Project, 322, 336–337, 339

O Operation, 7, 34, 37, 44–47, 52, 68, 70–78, 110, 112, 114, 118, 124–126, 138, 155–157, 159, 160, 163, 204, 206–212, 216, 217, 220, 232, 240, 268, 314, 326, 327, 333, 334, 336, 349 Optimal design

P Packed/fixed bed reactor (PBR), 156, 159, 162, 165–167 Pearl River Deltas, 250, 256–260, 264, 272, 273, 275, 277, 278, 282, 284, 285, 287, 288, 291, 292, 294

356 Planning, 30, 50, 52, 88, 124, 143, 145, 204, 206, 207, 213–218, 220, 222, 239, 247, 248, 256, 273, 295, 329, 332, 341 Polygeneration, 13, 173–186, 196, 197 Poor settling characteristics, 160 Power, 26, 59, 85, 108, 145, 173, 203, 231, 248, 303, 321, 342, 344 Power systems, 1–3, 5, 8, 9, 11, 12, 16, 39, 59–61, 72, 85, 86, 124, 137, 203, 206, 207, 213, 214, 321, 326, 334, 338, 339, 342 P2P energy trading, 46, 60–63, 65, 70, 85–91, 94–96, 98–100 P2P energy trading market, 95 P2P market, 67–69, 79, 89–91, 95, 98–100 Proton exchange membrane (PEM), 14, 42, 110–112, 114, 125, 137, 179, 207, 232, 233, 302, 303, 307, 309–311, 314, 324 Pyrolysis, 119, 137, 229, 234, 325

R Reliability, 8, 30, 35, 37, 50, 61–63, 85, 88, 124, 125, 132, 138, 206, 214–215, 219, 222, 314, 315, 331, 334, 336, 338, 339 Renewable adoption, 62, 206, 305 Renewable energy, 5, 6, 8–11, 15, 25–30, 34, 37–40, 45, 46, 52, 60, 62, 63, 65–67, 79, 85–88, 91, 99, 100, 106–121, 124, 125, 134, 136, 137, 144, 147–150, 173, 174, 195, 203, 204, 206, 207, 226–228, 231, 232, 237–239, 241, 252, 260, 290, 302–306, 311, 315, 322, 326, 327, 329, 332–335, 338, 342, 343, 348, 349 Renewable energy integration, 16, 47, 52, 59, 60, 85, 326, 339, 345 Renewable energy resources (RER), 9, 31, 60, 141, 226, 329, 330, 335, 345 Renewable energy sources (RES), 2, 5–9, 15, 16, 25, 26, 28, 35–38, 42, 59, 65, 85, 86, 88, 89, 99, 106–108, 112, 114, 115, 121–123, 125, 126, 134, 136, 137, 141, 143–146, 148–151, 173, 174, 195, 197, 203, 206, 207, 218, 226, 227, 230, 249, 264–266, 268–270, 272, 273, 288, 290, 291, 303–306, 309, 314–316, 321, 322, 325, 326, 330, 333, 334, 336, 339, 341, 343, 345, 346

Index Renewable sources, 8, 9, 16, 26, 28, 32, 35, 42, 65, 87, 105, 106, 126, 147, 150, 175, 179, 203, 250, 257, 260, 263, 301, 304, 305, 317, 321–324, 330, 332, 335, 337, 343, 345 Research and development initiatives, 315 Resource consumption, 173

S Sector coupling, 26, 27, 30–33, 220 Smart grid control, 26, 44–46 Solar energy, 2, 8–10, 14, 43, 95, 112–114, 203, 204, 207, 218, 230, 231, 233, 266, 268, 302, 306, 309, 345 Solar power, 9, 10, 12, 41, 144, 146, 148, 203, 204, 218, 309, 315, 326, 335, 336, 339 Solar systems, v, 11 Solid oxide electrolysis (SOE), 41, 110–112, 179, 183, 232, 233, 248, 324 Solid retention time (SRT), 156, 164 Steam methane reforming (SMR), 26, 105, 136, 229, 330, 334, 344 Supply chain, 29, 47, 51, 228–230, 240, 241, 250, 251, 253, 256, 260, 264–278, 284, 285, 295, 312, 349 Sustainability, 5, 7, 25, 27, 32, 35, 47, 51, 60, 63, 179, 185, 186, 196, 197, 239, 303, 312, 313, 322, 330, 334, 337, 346 Sustainable development, 47, 247, 313, 315, 317 Sustainable energy future, 314, 317, 339 Sustainable energy systems, 32, 59, 85, 86, 146, 301, 305, 306, 314, 321, 323, 337, 339 Syngas, 42, 184–186, 194–197 System performance, 12, 14, 188, 326

T Techno-economic analysis, 51, 60, 86, 327 Technological advancements, 9, 26, 41–46, 48, 86, 168, 206, 338 Thermodynamic assessment, 15, 177–178, 186 Tibet, 250, 256, 260–264, 266–270, 273–277, 285, 292, 295 Toluene, 250, 251 Transactive energy (TE), vi, 85, 86, 88 Transition to a low-carbon future, 41, 85, 321, 338

Index U United States projects, viii, 321–339 Up-flow anaerobic sludge bioreactors (UASB), 156, 159, 162–167

W Water-gas shift reactor, 184 Western Green Hydrogen Initiative (WGHI), 322, 328–330, 339 Widespread adoption, 11, 49, 52, 87, 148, 303, 321, 322, 325, 329, 331, 337, 338 Wind power, 7, 10, 12, 39, 40, 66, 207, 230–231, 257, 263, 269, 332, 339, 349

357 X Xinjiang, 249, 250, 256, 260–268, 270, 273–278, 285, 288, 291, 295

Y Yangtze River Deltas, 250, 256–260, 269, 270, 273, 275, 277, 278, 282, 284–288, 290, 291, 293–295