Pathways to Water Sector Decarbonization, Carbon Capture and Utilization
Pathways to Water Sector Decarbonization, Carbon Capture and Utilization
This is an Open Access book distributed under the terms of the Creative Commons
Attribution-Non Commercial-No Derivatives Licence (CC BY-NC-ND 4.0), which
permits copying and redistribution in the original format for non-commercial
purposes, provided the original work is properly cited.
(http://creativecommons.org/licenses/by-nc-nd/4.0/). This does not affect the rights
licensed or assigned from any third party in this book.
Edited by
Zhiyong Jason Ren and Krishna Pagilla
Apart from any fair dealing for the purposes of research or private study, or criticism or review, as permitted under the
UK Copyright, Designs and Patents Act (1998), no part of this publication may be reproduced, stored or transmitted
in any form or by any means, without the prior permission in writing of the publisher, or, in the case of photographic
reproduction, in accordance with the terms of licenses issued by the Copyright Licensing Agency in the UK, or in
accordance with the terms of licenses issued by the appropriate reproduction rights organization outside the UK.
Enquiries concerning reproduction outside the terms stated here should be sent to IWA Publishing at the address
printed above.
The publisher makes no representation, express or implied, with regard to the accuracy of the information
contained in this book and cannot accept any legal responsibility or liability for errors or omissions that may be
made.
Disclaimer
The information provided and the opinions given in this publication are not necessarily those of IWA and should
not be acted upon without independent consideration and professional advice. IWA and the Editors and Authors
will not accept responsibility for any loss or damage suffered by any person acting or refraining from acting upon
any material contained in this publication.
This is an Open Access eBook distributed under the terms of the Creative Commons Attribution Licence
(CC BY-NC-ND 4.0), which permits copying and redistribution for non-commercial purposes with no derivatives,
provided the original work is properly cited (https:// creativecommons.org/licenses/by-nc-nd/4.0/). This does not
affect the rights licensed or assigned from any third party in this book.
Chapter 1
Toward a net zero circular water economy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
Zhiyong Jason Ren, Jerald L. Schnoor and Krishna R. Pagilla
1.1 The Water Sector and the Challenges and Opportunities on Decarbonization . . . . . . . . . . . . . . 1
1.2 Pathways Toward Water and Wastewater Decarbonization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.2.1 Decarbonization requires a better understanding of emission baseline . . . . . . . . . . . . . 4
1.2.2 Decarbonization requires a combination of approaches and collaborations
among stakeholders . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.2.3 Processes and technologies that enable energy and resource recovery . . . . . . . . . . . . . . 6
1.2.4 Processes and technologies that enable additional benefits of carbon capture
and utilization, and watershed management . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.2.5 Case studies on utility decarbonization practice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.3 The Paradigm Change for a Net Zero Circular Water Economy . . . . . . . . . . . . . . . . . . . . . . . . 12
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
Chapter 2
What can we learn from decarbonization of the energy sector? . . . . . . . . . . . . . . . . . . . . . . . . 15
A. J. Simon and Seth W. Snyder
2.1 Introduction: Energy and Water: Similarities, Differences, and a Complex Relationship . . . . . 15
2.1.1 The energy-water nexus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.1.2 Differences in scale . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.1.3 The carbon-water nexus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
Chapter 3
Greenhouse gases in the urban water cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
Andrew Shaw, Kubeshnee Chetty, Tak Fan Chan, Elena Lindsey, Anjana Kadava
and Ben Stevenson
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.1.1 Overview of the urban water cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.1.2 Definition of scope 1, 2 and 3 emissions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.1.3 Water footprint and carbon footprint . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.2 Greenhouse Gasses in the Water Cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.2.1 Scope 1 – direct emissions – from own and controlled sources . . . . . . . . . . . . . . . . . . . 37
3.2.2 Scope 2 – GHGs from energy use . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.2.3 Scope 3 – indirect emissions from other activities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.2.4 Carbon sequestration and mitigation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.3 Protocols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.3.1 International protocols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.3.2 Regional protocols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.4 Methods of GHG Quantification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.4.1 Emission factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.4.2 Direct measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.4.3 Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
3.4.4 Quantification method selection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
3.5 A Framework for Carbon Footprint Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.5.1 A roadmap to reducing carbon footprint in the water cycle . . . . . . . . . . . . . . . . . . . . . . 47
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
Chapter 4
Operational optimization and control strategies for decarbonization in WRRFs . . . . . . . . . 51
Krishna R. Pagilla
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
4.2 Optimization Strategies at the Process/Operation Level . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.2.1 Wastewater pumping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
4.2.2 Secondary treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
4.2.3 Sludge treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
Chapter 5
Energy and resource recovery using the anaerobic digestion platform . . . . . . . . . . . . . . . . . . 67
Prathap Parameswaran, Jessica A. Deaver, Sudeep C. Popat, Vikas Khanna,
Madison Kratzer and Mel Harclerode
5.1 Current State of the Art for Anaerobic Digestion in Municipal Wastewater Resource
Recovery Facilities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
5.2 Need for Sludge Pretreatment to Enhance Viability of Anaerobic Digestion . . . . . . . . . . . . . . . 70
5.3 Diversifying Portfolio of Anaerobic Digestion at Municipal Wastewater Facilities –
The Advent of Anaerobic co-digestion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
5.3.1 Theoretical basis/substrates used . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
5.3.2 Challenges of ACoD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
5.3.3 Current research on ACoD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
5.4 Enhancing the Value of the Produced Biogas Through Co-generation and Further
Purification to Natural Gas for Pipeline Delivery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
5.4.1 Biomethane for combined heat and power production . . . . . . . . . . . . . . . . . . . . . . . . . . 72
5.4.2 Biomethane for electricity generation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
5.4.3 Biomethane for upgrading and pipeline delivery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
5.4.4 Biomethane for transportation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
5.4.5 Biogas to valuable chemicals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
5.5 Altering the AD Platform for Higher Organic Carbon Product Capture Coupled with
Water Reuse and Nutrient Recovery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
5.6 Energy Management in Anaerobic Digestion for Overall Energy Neutrality or
Energy Positive Treatment – The Case for Direct Anaerobic Treatment Through AnMBRs . . . . 74
5.7 Techno-economic and Life Cycle Assessments for Shaping the Future of
Anaerobic Digestion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
5.8 Future Strategies and Roadmaps to Decarbonization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
5.9 Anaerobic Digestion Techniques for Achievement of a Circular Economy . . . . . . . . . . . . . . . . . 78
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
Chapter 6
Carbon valorization using the microbial electrochemical technology platform . . . . . . . . . . 83
Jayesh M. Sonawane, Zhiyong Jason Ren and Deepak Pant
6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
6.2 The Principles of Microbial Electrochemical Carbon Valorization . . . . . . . . . . . . . . . . . . . . . . 84
6.2.1 Biocatalytic CO2 capture and conversion to organic chemicals in MES and EF . . . . 84
6.2.2 CO2 capture and mineralization in MECC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
6.3 Valorization of Carbon Compounds by MES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
6.3.1 Methane or acetic acid production in MES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
6.3.2 Role of hydrogen in MES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
6.3.3 CO2 valorization potential from the MES platform . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
6.4 Valorization of Carbon Compounds by Electrofermentation . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
6.4.1 Mechanisms of anodic and cathodic EFs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
6.4.2 Synergy between EF and anaerobic digestion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
Chapter 7
Decarbonization potentials in nitrogen management . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
Kester McCullough, Stephanie Klaus and Charles Bott
7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
7.1.1 Carbon footprint costs of nitrogen removal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
7.2 Carbon Removal/Diversion in Primary Treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
7.3 Approaches for Carbon Efficient Nitrogen Removal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
7.3.1 Traditional nitrification/denitrification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
7.3.2 Nitrite shunt and PNA (NOB out-selection) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
7.3.3 Partial denitrification/anammox . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
7.3.4 Aeration, alkalinity, and COD requirements for mainstream nitrogen
removal technologies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
7.3.5 Process control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
7.4 Shortcut Nitrogen Removal Implementations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
7.4.1 Sidestream treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
7.4.2 Mainstream PNA/nitrite shunt . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
7.4.3 Mainstream PdNA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
7.4.4 Partial denitrification/anammox (PdNA) case study . . . . . . . . . . . . . . . . . . . . . . . . . . 108
7.5 Conclusions and Future Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
Chapter 8
Decarbonization potentials in phosphorus management in the water sector . . . . . . . . . . . 115
Annalisa Onnis-Hayden, Dongqi Wang, Ali Akbari, Mi Nguyen and April Z. Gu
8.1 Overview of Global Phosphorus Consumption and Demand in Relation to Water
Sustainability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
8.1.1 P management and decarbonization potential pathways . . . . . . . . . . . . . . . . . . . . . . . . 117
8.1.2 Phosphorus management and policy: Current status and practice . . . . . . . . . . . . . . . . 118
8.2 Direct Decarbonization and Indirect Carbon Reduction Strategies from Point Source
and Non-point Sources of Phosphorus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
8.2.1 Phosphorus in agricultural waste streams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
8.2.2 Phosphorus in industrial effluent: Best management practices for decarbonization . . . .123
8.2.3 Phosphorus in domestic waste streams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
8.2.4 Phosphorus in urban runoff and best management practices for decarbonization . . . . 126
8.3 Decarbonization in Phosphorus Removal and Recovery Processes . . . . . . . . . . . . . . . . . . . . . 126
8.3.1 Carbon requirements in enhanced biological phosphorus removal processes . . . . . 126
8.3.2 Carbon footprint reducing via operational strategies for EBPR . . . . . . . . . . . . . . . . . 130
8.3.3 Carbon footprint reducing via new pathway/process for EBPR . . . . . . . . . . . . . . . . . . 131
8.3.4 Additional technologies for phosphorus removal and recovery from
wastewater streams with carbon footprint reduction potential . . . . . . . . . . . . . . . . . 132
8.4 Quantification of Decarbonization Potential from Phosphorus Removal and Recovery
Processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
8.4.1 LCA studies for the quantification of decarbonization potential for
non-point sources . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
8.4.2 LCA studies for P removal and recovery processes in WWTPs and
quantification of decarbonization potential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
8.5 Future Outlook and Research Needs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
Chapter 9
Decarbonization potentials using photobiological systems . . . . . . . . . . . . . . . . . . . . . . . . . . 143
Lara Méndez, Cristian A. Sepúlveda-Muñoz, María del Rosario Rodero,
Ignacio de Godos and Raúl Muñoz
9.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
9.2 Photosynthetic Wastewater Treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
9.2.1 Microalgae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
9.2.2 Purple phototrophic bacteria . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
9.2.3 Photobioreactors for wastewater treatment and resource recovery . . . . . . . . . . . . . . . 151
9.3 Enhanced Biogas Production from Microalgae and PPB for WWT Decarbonization . . . . . . 154
9.4 CO2 Capture and Biogas Upgrading Using Photosynthetic Systems During WWT . . . . . . . . 157
Acknowledgement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
Chapter 10
Sludge management and utilization for decarbonization . . . . . . . . . . . . . . . . . . . . . . . . . . . .171
Meltem Urgun-Demirtas, Rachel Dalke and Krishna R. Pagilla
10.1 Overview of Current Sludge Management Practices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
10.2 Sludge to Energy/Products . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
10.3 Land Application and Dedicated Landfilling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176
10.4 Reclamation of Brownfields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
10.5 Composting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
10.6 Resource Recovery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
10.7 Sludge Stabilziation for Removal of Emerging Contaminants . . . . . . . . . . . . . . . . . . . . . . . . . . 179
10.8 Centralized vs. Distributed Sludge Management Practices . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179
10.9 Environmental and Economic Life Cycle Assessment of
Sludge Management Technologies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179
10.10 Implementation: Challenges and Opportunities, Regulatory and Social Issues . . . . . . . . . . . 180
10.11 Future Strategies and Roadmaps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182
Chapter 11
Decarbonization potentials in intensified water and wastewater systems using
membrane-related technologies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
Boyan Xu, Shujuan Huang, Chuansheng Wang, Tze Chiang Albert Ng and
How Yong Ng
11.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
11.2 Membrane Strategies for Decarbonization in Wastewater Treatment and Resource
Recovery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
11.2.1 Aerobic granular sludge membrane bioreactors (AGMBRs) . . . . . . . . . . . . . . . . . . . . 189
11.2.2 Algae membrane bioreactors (A-MBRs) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
Chapter 12
Natural treatment systems and integrated watershed management
for decarbonization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207
Hannah R. Molitor and Jerald L. Schnoor
12.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207
12.2 Natural Treatment Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209
12.2.1 Constructed wetlands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 211
12.2.2 Treatment lagoons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212
12.2.3 Bioremediation and biofiltration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
12.2.4 Microalgal cultivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214
12.2.5 Land treatment systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214
12.3 Case Study: Carbon Sequestration and Agricultural Runoff Treatment through
Phytoremediation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
12.4 Case Study: Power Plant Flue Gas and Fertilizer Wastewater Treatment by Nutritious
Microalgae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218
12.4.1 Scaling-up microalgal cultivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
12.4.2 GHG and land footprints . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
12.4.3 Microalgal end products . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221
12.5 Conclusions and Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
Chapter 13
Microbial electrochemical communication in carbon and electron flow for CO2
methanation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227
Aijie Wang, Bo Wang, Zechong Guo, Weiwei Cai and Wenzong Liu
13.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227
13.2 Carbon Conversion and Electron Flow to Electromethanogenesis . . . . . . . . . . . . . . . . . . . . . 228
13.2.1 Functional communities and genes involved in carbon conversion . . . . . . . . . . . . . . 229
13.2.2 Organic conversion for CH4 production under electrochemistry regulation . . . . . . . 230
13.3 Upgraded CH4 Production from Electrochemically Enhanced Anaerpbic Digestion . . . . . . 232
13.3.1 Hytrogentrophic methanogenesis pathway in anaerobic digestion . . . . . . . . . . . . . . . 232
13.3.2 Microbial community evolution in cathode biofilm for methanation . . . . . . . . . . . . . 233
13.3.3 Microbial network of electrochemically enhanced AD . . . . . . . . . . . . . . . . . . . . . . . . 235
13.4 CO2 Methanation Driven by Solar-powered Bioelectrochemical System . . . . . . . . . . . . . . . . 237
13.4.1 Solar intermittent driven-power accelerates bioelectrochemical performances . . . . 237
13.4.2 Intermittent electro field mediates mutualistic interspecies electron transfer . . . . . . 240
Chapter 14
Thermal energy from wastewater . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
James McQuarrie
14.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
14.2 Wastewater as a Thermal Energy Source . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 251
14.3 Integration of Thermal Energy Recovery from Wastewater with Modern District
Energy Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 254
14.4 Assessing the Technical Feasibility of Thermal Energy Recovery from Wastewater . . . . . . . 256
14.5 Adoption of Thermal Energy Recovery from Wastewater . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 257
14.6 Opportunities and Barriers to Thermal Energy from Wastewater . . . . . . . . . . . . . . . . . . . . . . 259
14.6.1 Defining strategic planning . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 259
14.6.2 Demand and resource mapping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260
14.6.3 Technical feasibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 261
14.6.4 Regulatory and financing frameworks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 261
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 263
Chapter 15
Concept wastewater treatment plants in China . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 265
Jiuhui Qu, Hongqiang Ren, Hongchen Wang, Kaijun Wang, Gang Yu, Bing Ke,
Han-Qing Yu, Xingcan Zheng and Ji Li
15.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 265
15.2 The Current Challenges of China’s Urban Wastewater Treatment Facilities . . . . . . . . . . . . . . 266
15.3 Wastewater Concept Plant Provides a Vision and Example for Future Development . . . . . . 267
15.4 The Next Paradigm of Wastewater Treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 272
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 272
Chapter 16
Data science tools to enable decarbonized water and wastewater treatment systems . . . . . 275
Kathryn B. Newhart, Amanda S. Hering and Tzahi Y. Cath
16.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 275
16.2 Principle Data Science Tools . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 277
16.2.1 Data preparation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 277
16.2.2 Measuring accuracy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 279
16.2.3 Dimension reduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 280
16.2.4 Principal component analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281
16.2.5 Prediction and forecasting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 284
16.2.6 Optimization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 288
16.3 Data Science Applications to Select Treatment Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 290
16.3.1 Pump optimization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 290
16.3.2 Nitrification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 293
16.3.3 Anaerobic digestion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 294
Chapter 17
Decarbonization policies and water sector opportunities . . . . . . . . . . . . . . . . . . . . . . . . . . . 303
Jason A. Turgeon, Steven A. Conrad and Peter A. Vanrolleghem
17.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303
17.2 Core Concepts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303
17.2.1 Concept 1: In the scale of national or global policies the sector’s energy
use is relatively small, but there are other resources contained in wastewater
worth considering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303
17.2.2 Concept 2: There is no overarching policy that mandates decarbonization
in the water sector globally . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 306
17.2.3 Concept 3: Seek out co-benefits with other policy areas . . . . . . . . . . . . . . . . . . . . . . . 309
17.2.4 Concept 4: Water reuse and energy recovery may be more beneficial in a
distributed setting where the water and energy are recovered at/near the
point of generation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 312
17.2.5 Concept 5: Change is constant. We need water policy and technological
platforms that can more easily adapt to change . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 315
17.3 Conclusion and Policy Suggestions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 316
17.4 Additional Resources . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 317
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 318
Chapter 18
Outlook for the carbon-negative circular water economy . . . . . . . . . . . . . . . . . . . . . . . . . . . 321
Glen T. Daigger
18.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 321
18.2 Resource Recovery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 322
18.2.1 Historical perspective . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 322
18.2.2 Value hierarchies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 323
18.2.3 Advancing markets and products . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 325
18.3 Accelerating Transitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 326
18.4 The Path Forward . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 329
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 331
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 333
Bo Wang Gang Yu
Aarhus University, Aarhus, Denmark Tsinghua University, Beijing, China
Decarbonization is a topic and reality whose time has come. The water sector is committed to
meeting the needs of society, from water supply and sanitation through to protection of the natural
environment. Those needs now extend to contributing, where possible, to combating climate change,
and to demonstrating leadership on how to best use the world’s precious resources. This book supports
progress on both of these fronts.
Great practical progress is already being made in cities and utilities around the world with actions
aimed at achieving net zero carbon emissions. This progress combines a drive for greater efficiency in
our current approaches to managing water with the development and deployment of new processes
and technologies opening alternative approaches to resource use.
While technology breakthroughs and innovative designs can help us respond to some of these
impacts, this needs to be coupled with comprehensive system change. Water is a system of systems,
and decarbonization approaches need to be implemented and coordinated across these systems – at
the basin level, the city level and the utility level. More than simply improving the performance and
efficiency of the component parts, change is needed at a system level too. Therefore, such progress is
not something those in the water sector can deliver fully by working alone. It requires cooperation and
partnerships, in which the water sector can demonstrate leadership as a facilitator.
This book, written by experts – both water engineers and scientists – and specifically addressing
decarbonization pathways for the water sector, is much needed. It combines the foundations, evidence
and vision to support and stimulate practical progress.
This book can inspire the Global South. Here we see the biggest opportunities to reimagine
how we do water, as many of these places are starting from scratch and are on the cusp of making
huge investments in water infrastructure and services. It would be wonderful if this could be done
in a low carbon way – with energy neutral, efficient and productive use of water, maximizing the
capture of value from water and waste streams. Much as the Global South ‘leapfrogged’ fixed wires of
communications infrastructure, so they can avoid the slow, costly, mechanistic, and heavy legacy of
high carbon, centralized systems – by moving to low carbon, off-grid, distributed, flexible and circular
systems.
The scope, content and intentions of the book align with the outlook of the International Water
Association (IWA). The water sector, especially the utilities at its heart, faces multiple challenges in
meeting the current and future needs of society. IWA supports the journey ahead, providing leadership
and generating and sharing knowledge to enable action. IWA does this through initiatives such as
Climate Smart Utilities and Digital Water Program as well as others aimed at reframing thinking at
the basin and the city level, as well as promoting the adoption of nature-based solutions.
The Climate Smart Utilities initiative encourages utilities to become leaders in climate mitigation,
by providing them with approaches and tools to assess, monitor and reduce their greenhouse emissions
while enhancing their ability to adapt to climate change. Meanwhile, our Digital Water Program is
enabling utilities and its customers to transition towards a new low-carbon paradigm, where data-
driven models can help integrate and optimize smart pumps, valves, sensors and actuators in order to
maximize service levels while minimizing carbon footprint.
The latest developments in the field of decarbonization are brought together in this book thanks
to the input of an outstanding array of authors. Their expertise spans the diversity of this important
topic area, which touches all parts of the water sector, with each chapter drawing on a great depth of
expertise.
These contributions have been brought together expertly by the Editors, Zhiyong Jason Ren and
Krishna Pagilla. Their deep appreciation of this field and their awareness that now is such an important
time for these perspectives to be presented underpin the value of this book.
The book also highlights the need for leadership, and itself represents a contribution to leadership
in the sector, helping lay out a path ahead. Here, I see that the focus on decarbonization represents
an opportunity of fundamental importance – both to the sector and to the world at large. To date,
our economies have been built as high carbon economies and an expectation of a ready supply of
water. There is a window of opportunity for water to be at the heart of the low carbon economy – one
that shifts away from use of fossil fuels and at the same time recognizes our resource limitations and
adopts a circular approach. It is an opportunity for those in the sector to show leadership, and this
book provides valuable tools to do just that.
Few would argue that economic, scientific and socio-cultural advances achieved since the Industrial
Revolution have greatly enhanced the quality of life for a large number of our world’s population.
The environmental costs of these achievements, however, are now beginning to manifest in ways
which threaten those same life qualities we all have come to take for granted. The evidence is clear:
the direct impacts on our physical world from climate change are on the rise, threatening our quality
of life. Perhaps the greatest challenge climate change poses for humanity is not in discovering and
implementing solutions that slow and reverse the rate of changing climate, but in loosening our
collective grip on the comfort of our modern conveniences. This mindset hinders our progress toward
effective change.
The dynamics of a changing climate are pressing down hard on the water industry. We must think
differently about pathways to real solutions, the lifespan of solutions, and the resiliency of solutions.
This thinking requires intensifying our efforts in identifying and accelerating innovative water
treatment and distribution and wastewater collections and treatment technologies, coupled with
tactical applied research, and forging strategic relationships with organizations, including regulators,
to partner with us in implementing solutions. This book provided exactly these pathways.
Climate change is pushing us to put carbon management front and center of every solution approach
in the water space (i.e., stepping down our carbon footprints to net-neutral and ultimately to net-zero
operation). Though the water sector may need solutions oriented toward adaptation for the short term
to bolster resiliency, emphasis needs to be on outcomes that result in long-term mitigation measures
that ensure a sustainable future of our water environment—and the whole planet. All this means
recognizing water’s role in the management of all primary resources that support energy, agriculture,
mineral mining, manufacturing, and construction economy sectors. It’s also about reducing and
capturing wastes for product reuse, remanufacturing and recycling, reducing operational carbon
emissions and offsetting emissions through sequestration. These actions outline the circular economy,
the core of a sustainable future.
In many ways, the global water sector is the epitome of an ultimate circular economy model. It is
no secret that the changing climate significantly impacts the entire hydrologic cycle, from rainfall
to drought, from declining aquifers to rising sea levels, from soil erosion to declining water quality,
and more. The first and foremost mission of the water industry is to supply, treat and distribute safe,
potable water to protect public health and sustain life itself, but in the context of climate change,
water is a “product” that every living thing depends on for survival, so the water industry is uniquely
positioned to effectively enact circular economic principles and lead by example on mitigating climate
change through decarbonization.
Our time to implement circularity in the water industry to abate the impacts of climate change is
running short. A business-as-usual mindset is unacceptable. The status quo must be shattered. The
thought-provoking ideas and potential pathways contained in this book provide a framework for
which circularity within the water industry can be a reality. The authors and the editors have provided
the start and a route map well. The industry must act now.
Preface
The water sector is in the middle of a paradigm shift from focusing on treatment and meeting
discharge permit limits to integrated operation that also enables a circular water economy via water
reuse, resource recovery, and system level planning and operation. While the sector has gone through
different stages of such revolution, from improving energy efficiency to recovering renewable energy
and resources, when it comes to the next step of achieving carbon neutral or negative emissions, it
falls behind other infrastructure sectors such as energy and transportation. Decarbonization refers
to the reduction of carbon footprint of an industry and in the long run creates a circular economy
with integrated solutions for carbon management. The water sector carries tremendous potential to
decarbonize, from technological advancements to operational optimization, to policy and behavioral
changes.
This book aims to fill an important gap for different stakeholders to gain knowledge and skills in
this area and equip the water community to further decarbonize the industry and build a low carbon
society and economy. The book goes beyond technology overviews; rather it aims to provide a system
level blueprint or pathways for decarbonization, carbon capture, and utilization in the water sector.
We hope that this book will become an inspiration to develop practices and solutions that will drive
innovation in the water sector for decarbonization. Here is a snapshot of what you will find in the
book.
The first section of the book lays out a framework on the state-of-the-art in water sector carbon
footprints. Chapter 1 provides an overview of the challenges and opportunities on water sector
decarbonization, and it summarizes the needs and approaches to achieve the net zero carbon goals.
Chapter 2 offers a comprehensive review on the pathways other infrastructure sectors (e.g., energy,
transportation) explored and identifies the synergies and examples for the water sector to consider.
Chapter 3 discusses the different scopes of greenhouse gas (GHG) emissions associated with the urban
water cycles and provides an overview of the carbon footprint accounting methods and protocols.
The second section of the book provides reviews and details on different processes and technologies
that enable decarbonization, carbon capture, and utilization. The experts in each respective field offer
deep insights on how the approach has been used to increase energy efficiency, reduce carbon footprint,
recover resources, and capture and valorize GHG while maintaining treatment goals. Chapter 4 starts
© 2022 The Editors. This is an Open Access eBook distributed under the terms of the Creative Commons Attribution Licence
(CC BY-NC-ND 4.0), which permits copying and redistribution for non-commercial purposes with no derivatives, provided the
original work is properly cited (https://creativecommons.org/licenses/by-nc-nd/4.0/). This does not affect the rights licensed or
assigned from any third party in this book. The chapter is from the book Pathways to Water Sector Decarbonization, Carbon Capture
and Utilization, Zhiyong Jason Ren and Krishna Pagilla (Eds.)
with the easily implemented methods associated with operations at pumping, preliminary, primary,
secondary, advanced, and sludge treatment level within a water resource recovery facility (WRRF).
Chapter 5 focuses on the energy and resource recovery from the commonly used anaerobic digestion
(AD) platform, which also includes emerging processes such as anaerobic membrane bioreactors
(AnMBR) and thermal hydrolysis. Chapter 6 explores opportunities for renewable energy production
and carbon valorization using the new microbial electrochemical technology (MET) platform. Chapter
7 and Chapter 8 investigate the critical considerations and tremendous potentials on decarbonization
during nitrogen and phosphorus removal and recovery processes, respectively. Chapter 9 describes
the increasing popular photobiological systems using microalgae and cyanobacteria for CO2 capture
and conversion. Chapter 10 focuses on sludge management and utilization using AD, compositing,
incineration, as well as emerging processes like hydrothermal liquefaction (HTL). Chapter 11
discusses several novel membrane technologies that enable process intensification with reduced
energy consumption, and it provides an overview of system integration and optimization. Chapter 12
analyzes the advantages and challenges of natural treatment systems and their potential for integrated
watershed management and decarbonization in the context of One Water. Chapter 13 discusses
the fundamental carbon and electron flows occurring in anaerobic bioconversion systems for CO2
capture and conversion to value-added organic chemicals. Lastly, Chapter 14 assesses the potential in
recovering the abundant low quality thermal energy from wastewater and its feasibility of integration
with district heating.
The third section of the book covers the broader prospects in water sector decarbonization in the
context of policy making, intelligent water systems, as well as case studies. Chapter 15 describes
several “wastewater concept plants” designed and built in China in recent years as examples for
the next generation of WRRFs for integrated waste management and resource recovery. Chapter 16
introduces the modern data science tools including statistical and machine learning methods that can
be used for decarbonization. Chapter 17 provides a critical analysis of local and national polices that
will impact the efforts and highlights the need to seek multiple benefits whenever possible. Lastly, the
concluding Chapter 18 summarizes the evolution of the missions of water management, the need of
concerted efforts to move the industry forward, and the tangible benefits such endeavors will make to
the society, the economy, and the environment.
This book can be a reference book and textbook for undergraduate and graduate students,
researchers, practitioners, consultants, and policy makers, and it will provide practical guidance for
stakeholders to analyze and implement decarbonization measures in their professions. The goal is to
provide pathways for decarbonization from various perspectives. We are confident that the readers of
this book will be inspired to seek innovation in water management while achieving decarbonization.
We want to thank the authors and contributors for their time in the writing of this book and their
dedications in advancing the understanding and contributing to the grand mission of water sector
decarbonization. Their expertise and knowledge have been generously shared in this book. We also
want to thank The International Water Association, Knowledge Unlatched, Princeton University, and
University of Nevada Reno for their generous support to make this book accessible to many readers.
Krishna R Pagilla,
University of Nevada, Reno
Chapter 1
Toward a net zero circular water
economy
Zhiyong Jason Ren1*, Jerald L. Schnoor2 and Krishna R. Pagilla3
1Department of Civil and Environmental Engineering & Andlinger Center for Energy and the Environment, Princeton University,
Princeton, NJ, USA
2Department of Civil and Environmental Engineering, University of Iowa, Iowa City, IA, USA
3Department of Civil and Environmental Engineering, University of Nevada Reno, Reno, NV, USA
*Correspondence: [email protected]
Figure 1.1 The OneWater framework for the urban water cycle (image adopted from jacobs.com, 2020).
are being considered for both water extraction and supply. This further adds to the energy intensity of
water treatment and wastewater reclamation. Water, stormwater, wastewater collection, transport, and
treatment use tremendous amounts of concrete, metal, and plastic materials, all of which are associated
with non-renewable materials and energy sources or with intensive energy footprints.
Water and wastewater utilities typically spend 10–35% of their operational costs on energy, which
is mostly generated from fossil fuel sources (IKI, 2020). This can account for as much as 40% of
municipal energy use in some cities, and it mirrors the GHG emissions from the water sector. In
water treatment facilities and supply systems, both scope 2 and scope 3 emissions are more significant
than scope 1 (direct) emissions, while the direct non-biogenic GHG emissions within a wastewater
treatment facility are scope 1 emissions. In addition, emissions include those related to imported
electrical and thermal energy (scope 2) and other indirect emissions associated with the production
and transportation of chemicals and fuels, waste disposal, as well as contracted services in both
water and wastewater facilities (scope 3) (UNEP, 2017). However, different from other infrastructure
sectors like energy and transportation where the primary GHG source is fossil-based CO2, the direct
release of CO2 from wastewater via organic degradation is largely considered carbon neutral due to its
biogenic nature, despite evidence of some carbon being of fossil origin (Griffith et al., 2009). Instead,
the non-CO2 direct emissions (primarily CH4, N2O) from collection systems and treatment facilities
are of significant concern because such GHGs are many times (28–298×) stronger in global warming
potential (GWP) than CO2 over 100 years (scope 1) (Figure 1.3) (Lu et al., 2018). Currently, wastewater
Figure 1.2 Global energy uses by different water-related activities (IEA, 2017).
Figure 1.3 Global non-CO2 emissions by sector and source (2015) (USEPA, 2019a, 2019b).
accounts for ∼5% of the global total non-CO2 GHG emissions, but the impacts are expected to be
much higher with CH4 emission control becoming a top priority in the next decade. Methane has >80
times the warming power of CO2 over the first 20 years, and it only lasts 12–15 years in the atmosphere
(Saunois et al., 2020), so cutting methane emissions is the fastest opportunity to slow the rate of global
warming and get to a net-zero emissions scenario by the mid-century.
In order to meet the 1.5°C global warming target in the Paris Agreement, many countries, cities,
and industries have made commitments to move to net zero emissions by 2030–2060. That means
GHG emissions would have to be dramatically reduced, and any remaining emissions will need to
be balanced by absorbing an equivalent amount from the atmosphere (negative emissions or offsets).
While many infrastructure sectors such as energy, transportation, and building systems have been
extensively studied with regards to decarbonization pathways, the water sector is lagging behind.
The water sector has not been considered as a carbon intensive industry, and because most water and
wastewater utilities are heavily regulated public entities, hence, they lack the power to control prices
that allow them to rationalize investments for long-term benefits. The traditional public perception
is that water and wastewater service is a ‘human right’, meaning price is more of a function of cost,
instead of value as in other sectors. Due to increasing water scarcity and environmental pollution
in different parts of the world, such a ‘water should be free’ concept is being challenged, and many
opportunities have emerged to overcome such hurdles by developing win-win solutions such as
generating ‘green’ revenues via energy and resource recovery, developing new policies on carbon
credits, and transforming empirical practice to data-driven decision making that improves efficiency
and reduces cost.
Figure 1.4 The reference baseline emissions of the UK water sector in 2018–2019 (Water UK, 2020).
While scope 2 emissions associated with grid electricity used in water/wastewater treatment and
conveyance is relatively straightforward, larger uncertainties exist in estimating scope 1 direct
emissions from treatment processes and conveyance systems. Generalized emission factor (EF) based
estimates oversimplify the situation and do not reflect the reality of differences in actual emissions
between various plants and processes, while localized flux chamber methods generate significant
variations, sometimes as much as 3–4 orders of magnitude different (Delre et al., 2017; Vasilaki et al.,
2019). Further research is required to establish a better scientific basis for the sector specific emission
factors. Chapter 3 discusses the different scopes of GHG emissions associated with the urban water
cycle. It also provides a comprehensive framework for carrying out a carbon footprint assessment
along with an overview of available and relevant protocols and methods for assessing GHG emissions
for the water sector.
intensive instruments such as blowers, pumps, and boilers with more energy efficient ones, as well
as renewable energy recovery. For example, an increasingly common practice is to produce biogas
from anaerobic digestion, and then using the biogas for heat, electricity or cogeneration of heat and
electricity in a combined heat and power (CHP) station. In fact, the water sector is expanding to add
waste materials such as food waste and other organic wastes as feedstocks for anaerobic digestion
to produce biogas that can enable the facilities to be energy neutral or energy positive. Resource
recovery from wastewater in terms of energy, nutrients (N and P) and water has resulted in renaming
of wastewater treatment facilities into water resource recovery facilities (WRRFs). The embedded
value of these recovered resources provides offsets for carbon emissions and hence can also contribute
to net zero outcomes. Furthermore, new and innovative solutions such as carbon financing and real-
time energy and chemical audits have also begun to make an impact.
In addition to treatment operations, the conveyance and storage of water and wastewater
requires attention to reduce material consumption, energy demand, and construction of unnecessary
infrastructure. Smart water meter installation to improve leak detection and reduce water use would
also contribute to sustainable water management and lifestyle changes. Wastewater collection systems
that minimize methane emissions due to anaerobic or septic conditions through both design and
operational strategies are needed and are at embryonic stage and growing. Strategic evaluations of
water and wastewater facilities siting to minimize conveyance energy needs should be coupled with
centralized versus decentralized treatment facilities scenarios. The best decarbonization pathways
are possible and provide maximum benefits when water is seen as OneWater in the urban watershed.
The water sector is heavily driven by a wide variety of policies. These policies not only include
regulatory mandates on water quality and public health protection, but also include directives
from different government agencies to ensure water affordability, social equity, ecological diversity,
and infrastructure resiliency. Heavy investments in both financial and human capitals have gone
into building and governing these systems as stove-piped services. There is no overarching policy
that mandates decarbonization in the water sector globally, but seeking out co-benefits with other
policy areas such as recovering local energy and resources, building multi-functional facilities, and
protecting community assets would provide more feasible solutions. Continued progress will require
strategic planning that begins at the local level and grows to global initiatives. This is especially
the case with decarbonization. Otherwise, local action may focus almost solely on infrastructure
resilience while the primary overlying issue of climate change requires decarbonization at the local
level. In the last few decades, water and wastewater utilities have increasingly embraced the role
and the value contribution to rate payers through performance-based operations. These efforts help
reduce resource utilization and recover resources that hold local value, and such value-proposition
provides utilities with better financial health and community support in making decisions that lead
toward decarbonization. Chapter 17 discusses several concepts in policy making that impact efforts
to decarbonize the water sector and highlights the need to seek multiple benefits whenever possible.
1.2.3 Processes and technologies that enable energy and resource recovery
Wastewater contains a significant amount of heat, chemical, and hydraulic energy that in total is
estimated to be many times that required to treat the wastewater. Therefore, it is absolutely feasible to
make WRRFs energy neutral or even positive. Figure 1.5 depicts a generalized view of energy flows at a
WRRF. Technologies enable utilities to minimize the resources utilized to treat wastewater via process
intensification, low energy treatment, and reduced chemical use. Leading utilities also maximize the
extent of resources recovered through the treatment process via biosolids land application, nutrient
recovery, biogas utilization, and water reclamation. A variety of technologies have been developed
to make such operations a reality, and these technologies include but are not limited to anaerobic
digestion, microbial electrochemistry, photobiological systems, advanced nitrogen and phosphorus
management, process intensification using membranes and other technologies, as well as heat/
pressure recovery and processes that significantly increase energy efficiency.
Figure 1.5 A generalized view of energy flows at a water resource recovery plant (WRRF) (revised from WEF (2019)).
The ‘low-hanging fruit’ for utility decarbonization is operations optimization. For an existing facility,
the carbon footprint lies in operational activities and, thus, the decarbonization potential as well.
Chapter 4 discusses the current opportunities for decarbonization at pumping, preliminary, primary,
secondary, advanced, and sludge treatment level within a WRRF. For example, influent wastewater
pumping has tremendous potential for reducing energy use, and data driven strategies using fuzzy
logic, data mining, and bench-marking provide good tools to reduce specific energy and to improve
energy savings (Torregrossa et al., 2017). Similarly, aeration is the single largest source of energy use in
most plants, and alternative diffusers and control systems taking advantage of the newer developments
in membranes and online sensing have been utilized to make aeration energy efficient. At the whole
plant level, different strategies can be employed in addition to individual unit level optimizations. For
example, plant level benchmarking with comparable facilities to identify opportunities, or optimization
of plant capacity utilization by arranging peak flow and load management, can be utilized. As
chemicals use make up a significant portion of a WRRF’s carbon footprint, reduction in chemical use
through operational optimization is highly feasible for decarbonization and also results in operational
costs reduction. Personnel training is another key area that can have a significant impact. With clear
demonstration of cost savings and performance enhancement through best practices, operators will
make direct contributions to decarbonization.
Anaerobic digestion (AD) is a model technology used in WRRFs to break down and stabilize
wastewater sludge and to generate biogas and nutrient-rich effluent and biosolids. AD has been a
central part of energy and resource recovery in the wastewater industry, and a suite of new processes
and technologies have been developed to enhance sludge conversion, improve biogas production,
upgrade biogas to higher-value products, and increase biosolid quality and applicability. Chapter 5
summarizes the current knowledge of the AD platform for energy and resource recovery, including the
emerging trend of sludge pre-treatment using thermal, chemical, mechanical, and electrical methods
to increase sludge degradability, sludge co-digestion with food waste and other organic waste for
improved biogas yields and biosolids quality. It also discusses newly developed processes such as
anaerobic membrane bioreactors (AnMBR), thermal hydrolysis, and volatile organic acids production
using the AD platform. Chapter 10 furthers the discussion of AD in the broader perspective of
sludge management, and it also discusses other practices including land application, composting,
1.2.4 Processes and technologies that enable additional benefits of carbon capture and
utilization, and watershed management
Tremendous progress on decarbonization has been made by increasing energy efficiency and recovering
renewable energy, but these methods only reduce fossil fuel consumption and its associated carbon
emissions. Considering the vast amount of wastewater generated each year (∼1000 km3 per year
worldwide) and its positive correlation with population and industrial activities, wastewater treatment
may even become carbon negative by capturing external CO2 and CH4 sources and converting them
into value-added products. Because such practice can occur within existing wastewater infrastructure
during treatment, no additional land or transportation would be required for such operations.
Natural treatment systems (NTS) utilize and enhance natural processes involving vegetation, soil and
water, and the associated microbial ecosystems, and they are effective in advanced treatment, emerging
contaminant removal, stormwater management, biomass production, recreational and educational
services, and overall integrated watershed/sewershed management. NTS requires little mechanical or
technological input to function, making them less chemical or energy intensive. Furthermore, these
phyto- and microbial-based systems provide a wide range of carbon capture profiles depending on the
level of treatment, seasonal variation, and system variation. Chapter 12 describes natural treatment
technologies, their advantages and disadvantages, and their potential to decarbonize the water sector
when incorporated in integrated watershed management. Under such practice, nutrients, energy, and
water are recovered from wastewater, and GHG emissions are mitigated. By reclaiming and reusing the
water for non-potable purposes such as irrigation/fertigation, aquifer recharge, graywater applications,
or even for direct potable reuse, source water withdrawal can be reduced, aquifers are replenished, and a
harmony between water demand and supply can be achieved. The chapter also provides two engineering
case studies on the benefits of phytoremediation for carbon sequestration and agriculture runoff treatment,
as well as using microalgae for combined power plant flue gas and fertilizer wastewater treatment.
Phototrophic microorganisms like microalgae or purple photosynthetic bacteria present a unique
pathway for decarbonization, as they fix CO2 during autotrophic growth while assimilating nutrients (N
and P) in wastewater. Therefore, phototropic treatment systems are complementary to carbon-focused
treatment processes like AD or microbial electrochemistry. Chapter 9 describes the decarbonization
potentials using photobiological treatment systems. For example, microalgae have been widely studied
for CO2 capture and utilization, including extensive research on their use in large-scale (>5000
acre) cultivation systems to produce feedstocks for biofuels and bioproducts. When they are used in
wastewater treatment, they are operated at an accelerated rate compared to terrestrial plants, and
they can be integrated with AD to provide additional substrate and to condition biosolids, capture
and utilize CO2 and upgrade biogas into biomethane, and even use H 2S as an electron donor. Recent
advances in photobioreactor design have boosted the biodegradation potential of photobiological-
based systems, while lowering their energy demand, and they are critically discussed in this chapter.
Another promising technology platform for simultaneous wastewater treatment, resource recovery,
and carbon capture and utilization is microbial electrochemical technology (MET). MET offers an
extremely flexible platform for both oxidation and reduction reaction-oriented processes. The MET
systems share one common principle in the anode chamber, in which biodegradable substrates are
oxidized and generate electrical current. The current can be captured directly for electricity generation
(microbial fuel cells, MFCs) or used to produce H 2 and other value-added chemicals (microbial
electrolysis cells, MECs). In addition, such electrons from organic waste carbon can also be used in
the cathode chamber to reduce CO2 and generate organic or inorganic compounds, achieving double
benefits of carbon capture and valorization. Chapter 6 presents the principles and popular MET
carbon capture processes and discusses the variety of products and systems that have been developed.
Microbial electrosynthesis converts CO2 into organic compounds such as carboxylic acids and CH4,
while microbial electrolytic carbon capture mineralizes CO2 into carbonate products. In addition,
electro-fermentation (EF) uses electrochemistry to influence microbial metabolism and regulate
fermentation pathways to valorize organic waste carbon to higher-value products, and many consider
it to be an electrochemically enhanced AD system. A unique feature of MET is the complementary
nature with anaerobic fermentation or digestion, in which synergistic interspecies electron transfer
can occur between microbes to facilitate electro-methanogenesis or electro-acetogenesis. Chapter 13
focuses on the fundamental carbon and electron flows in such anaerobic bioconversion systems for
CO2 capture and conversion to value-added organic chemicals.
While transforming wastewater treatment to carbon-neutral/negative or even revenue-positive takes
concerted efforts from stake holders, studies have demonstrated the potential benefits of implementing
new processes and technologies. Figure 1.6 demonstrates a hypothetical process combination of microbial
carbon capture cell (MECC) and microalgae reactor to replace the traditional anaerobic/anoxic/aerobic
activated sludge process. The MECC specializes on organic carbon removal, while microalgae is
effective in nutrient removal. Moreover, they both demonstrated excellent carbon capture and utilization
(CCU) capability, with MECC converting CO2 into carbonate mineral accompanied with high rate H2
production, and microalgae captures CO2 as biomass, which subsequently can be converted to biofuels
or biochar (Lu et al., 2018). Preliminary quantitative analyses using the US and China as examples
showed that instead of being a net emitter of GHGs, a net of up to 112 (median; 5th–9th percentile
range of 84–145; USA) and 75 (57–97; China) MtCO2e can be captured and converted to valued-added
products. Among these negative emissions, approximately 41–56 and 47–58% are attributed to CCU
during organic and nutrient removal, respectively; and –2–2% is credited to avoided consumption of
fossil energy during CCU (avoided aeration, etc.; the negative value stems from uncertainty analysis). In
terms of economic benefits, while the proposed system is likely to have even greater Capex and Opex
costs than conventional processes, the recovered mineral and biofuel products may create 8.7 (6.9–10.9)
and 5.6 (4.4–6.9) billion dollars in value per year for the US and China, respectively. Additionally, carbon
capture credits in the two countries could also mobilize 4.5 (3.3–6.2) and 1.0 (0.7–1.5) billion dollars for
US and China wastewater industries, correspondingly. These estimates demonstrate that the wastewater
industry can become a significant contributor of negative carbon emissions, though significant technology
development and testing are needed since neither process has been demonstrated in full scale.
Water utilities collect and store massive amounts of data to provide a reliable and efficient service.
The data collected not only include water quantity and quality data in every treatment facility but also
Figure 1.6 Preliminary estimates of carbon capture and utilization benefits from an example integrated
MECC + microalgae process compared with conventional activated sludge process: (a) Conventional anaerobic/
anoxic/aerobic activated sludge process for simultaneous carbon and nutrient removal; (b) Integrated MECC and
microalgae cultivation for carbon and nutrient removal with resource recovery and CCU; (c) The preliminary estimates
of CCU potential and economic impacts when the conventional process is compared with the MECC + algae process.
consists of real-time data from flow monitors, and rain and stream gauges across the watershed, as
well as data at endpoints and in water/sewer pipelines. Data has become an essential asset for utilities
and will play increasingly critical roles for utilities of the future. Accordingly, Chapter 16 introduces
the modern data-driven modeling (DDM) including statistical and machine learning methods and
uses specific examples to demonstrate how these tools can be used within a larger decarbonization
strategy. The chapter explains data preparation, common DDM methods, and metrics for comparing
different models. It also analyzes unit processes and how data-driven process optimization can become
the proverbial ‘low risk, high return’ approach for carbon and cost reduction.
Figure 1.7 The design view and aerial view (insert) of the Yixing Concept Wastewater Resource Recovery Factory
in Jiangsu, China (Qu et al., 2022). The factory started operation in October 2021.
1.3 THE PARADIGM CHANGE FOR A NET ZERO CIRCULAR WATER ECONOMY
‘Water management is a path, not a destination.’ The same philosophy needs to apply to decarbonization
of the water. It is how we pursue decarbonization, identify multiple pathways, try and refine them,
and scale them to the entire water sector that will lead to positive outcomes. Former IWA President,
Prof. Glen Daigger summarizes a water professional’s core mission in the concluding Chapter 18. The
evolution in the governance and infrastructure of a water utility reflects the progression of investments
to address the most pressing needs for its community. The mission of water management has grown
through time, from the initial focus on reliable water supply and prevention of the spread of diseases
to include multifaceted objectives on water resource recovery, water/sewershed management, public
health protection, and sustainable development to not only consider economics but also environmental
and societal impacts. The momentum for change is accelerating as we transition from the current
linear economy to a circular one, and the water sector can and should provide leadership in providing
essential public services.
The topics in this book articulate many opportunities currently available to the water sector to
decarbonize and transition into a circular water economy. Numerous innovators and early adopters
are investigating options, conducting trials, and executing projects to increase energy efficiency,
reduce carbon footprints, and recover resources from the OneWater cycle. There is no good or bad
product, but there can be a right or wrong product for a specific utility. Resources need to be recovered
and carbon footprints need to be reduced, but we also need to recognize that unless the products
have sufficient market demand, a valid value proposition, and tangible benefits to the society from
decarbonization, any changes made will not be sustainable in the long term.
There is no single solution that achieves net zero on its own, so it is imperative that all stakeholders
work together and collectively transform how the water sector plans, invests, and operates with
balanced near- and long-term goals in mind. The pathways laid out in the following chapters are
developed by visioning the possible net zero futures for the sector as well as individual utilities,
and they provide critical insights for the industry to move towards a circular economy that ensures
sustainability for future generations.
REFERENCES
Delre A., Mønster J. and Scheutz C. (2017). Greenhouse gas emission quantification from wastewater treatment
plants, using a tracer gas dispersion method. Science of the Total Environment, 605–606, 258–268, https://
doi.org/10.1016/j.scitotenv.2017.06.177
Griffith D. R., Barnes R. T. and Raymond P. A. (2009). Inputs of fossil carbon from wastewater treatment plants
to US rivers and oceans. Environmental Science & Technology, 43(15), 5647–5651, https://doi.org/10.1021/
es9004043
IEA (2017). Water-Energy Nexus. IEA, Paris. Available at: https://www.iea.org/reports/water-energy-nexus
(accessed 18 November 2021).
IKI (2020). Water Companies on The Way to CO2 Neutrality. Available at: https://www.international-climate-
initiative.com/en/news/article/watercompanies_on_the_way_to_co2_neutrality (accessed 18 October 2021).
Jacobs (2020). One Water. Available at: https://www.jacobs.com/solutions/water/one-water (accessed 18 November
2021).
Lu L., Guest J. S., Peters C. A., Zhu X., Rau G. H. and Ren Z. J. (2018). Wastewater treatment for carbon capture
and utilization. Nature Sustainability, 1(12), 750–758, https://doi.org/10.1038/s41893-018-0187-9
Qu J., Wang H., Wang K., Yu G., Ke B., Yu H. Q. and Gong H. (2019). Municipal wastewater treatment in China:
development history and future perspectives. Frontiers of Environmental Science & Engineering, 13(6), 1–7.
Qu J., Ren H., Wang H., Wang K., Yu G., Ke B. and Li J. (2022). China launched the first wastewater resource recovery
factory in Yixing. Frontiers of Environmental Science & Engineering, 16(1), 1–2, https://doi.org/10.1007/
s11783-021-1429-z
Saunois M., Stavert A. R., Poulter B., Bousquet P., Canadell J. G., Jackson R. B. and Zhuang Q. (2020). The
global methane budget 2000–2017. Earth System Science Data, 12(3), 1561–1623, https://doi.org/10.5194/
essd-12-1561-2020
Sedlak D. (2014). Water 4.0. Yale University Press, New Haven, CT, USA.
Torregrossa D., Hansen H., Hernandez-Sancho F., Cornelissen A., Schutz G. and Leopold U. (2017). A data-
driven methodology to support pump performance analysis and energy efficiency optimization in waste
water treatment plants. Applied Energy, 208, 1430–1440, https://doi.org/10.1016/j.apenergy.2017.09.012
UNEP (2017). The Emissions Gap Report 2017: A UN Environment Synthesis Report. UN Environment, Kenya.
USEPA (2019a). Global Non-CO2 Greenhouse Gas Emission Projections & Mitigation 2015–2050. USEPA, USA.
USEPA (2019b). Sources of Greenhouse Gas Emissions. Available at: https://www.epa.gov/ghgemissions/sources-
greenhouse-gas-emissions (accessed 18 October 2021).
Vasilaki V., Massara T. M., Stanchev P., Fatone F. and Katsou E. (2019). A decade of nitrous oxide (N2O) monitoring
in full-scale wastewater treatment processes: a critical review. Water Research, 161, 392–412, https://doi.
org/10.1016/j.watres.2019.04.022
Water Environment Federation (2019). ReNEW water project: resource recovery to fuel and grow a circular economy.
WEF, USA.
Water UK (2020). Net Zero 2030 Routemap. Available at: https://www.water.org.uk/routemap2030 (accessed 18
October 2021).
Wett B., Buchauer K. and Fimml C. (2007). Energy self-sufficiency as a feasible concept for wastewater treatment
systems. 4th IWA Leading-edge Conference and Exhibition on Water and Wastewater Technologies, IWA,
United Kingdom.
Chapter 2
What can we learn from
decarbonization of the energy
sector?
A. J. Simon1* and Seth W. Snyder2,3
1Lawrence Livermore National Laboratory, Livermore, CA, USA
2Idaho National Laboratory, Idaho Falls, ID, USA
3Northwestern University, Evanston, IL, USA
*Correspondence: [email protected]
© 2022 The Editors. This is an Open Access eBook distributed under the terms of the Creative Commons Attribution Licence
(CC BY-NC-ND 4.0), which permits copying and redistribution for non-commercial purposes with no derivatives, provided the
original work is properly cited (https://creativecommons.org/licenses/by-nc-nd/4.0/). This does not affect the rights licensed or
assigned from any third party in this book. The chapter is from the book Pathways to Water Sector Decarbonization, Carbon Capture
and Utilization, Zhiyong Jason Ren and Krishna Pagilla (Eds.)
• Anaerobic digestion of wastewater fractionates the embedded carbon into methane and CO2.
Methane has a radiative forcing factor 28 times greater than CO2 over 100 years (and 86 times
greater over 20 years), resulting in substantial short-term climate impacts if the methane is not
captured and combusted (Roy et al., 2015).
• Innovations at the unit process level are the most straightforward way to manage the carbon
intensity of the water sector.
• Structural change in the water economy (more efficient water use, potable and non-potable
reuse, etc.) may also affect the sector’s carbon intensity.
Table 2.1 Annual flows of widely used material commodities in the US.
of electricity consumed by water and wastewater treatment plants in the US (Greenberg et al., 2017).
It is likely that additional GHG emissions are attributable to the water and wastewater treatment
sector from onsite natural gas combustion, however no data could be found to quantify this emissions
source. Offsite manufacturing of chemicals used for water and wastewater treatment have also been
hypothesized to contribute significantly to the sector’s life cycle GHG footprint, but estimates of this
quantity in the literature vary widely (Kyung et al., 2015; Szulc et al., 2021).
Combustion of fossil fuel across the entire energy sector in the US emits 5300 MMT of CO2.
These two statistics are not directly comparable; the 61 MMT CO2-e associated with the water sector
accounts for CO2 and other GHGs from scopes 1, 2, and 3 for a specific sector while the 5300 MMT
CO2 in the energy sector accounts for only fossil-fuel derived CO2. Additionally, scope 2 emissions
from water treatment (∼23 MMT) are included in the 5300 MMT of fossil fuel-derived emissions from
the energy sector. However, the vast disparity in scale between these two figures demonstrates that
despite managing a far larger quantity of material, the water sector manages a far smaller quantity
of carbon. This overlap between water sector GHG emissions and energy sector GHG emissions is a
telltale of the Energy-Water Nexus described above.
the 12 years between 2005 and 2017, carbon intensity declined from 56.6 to 49.9 MMT/EJ, a rate of
1% per year (EIA, 2021) (each of these statistics takes the five-year average carbon intensity of energy
around the reported year to smooth out noise in the statistics – looking at individual years, it appears
that the 2005–2017 trend continued through at least 2019 and was likely accelerating). This ten-fold
increase in the pace of decarbonization is due to the following changes in the energy system (listed in
order of size of carbon intensity impact):
In addition to the decrease in the carbon intensity of delivered energy, the energy intensity of the
US economy has declined. In real GDP terms (all values quoted in 2012 dollars), the US consumed
13.1 exajoules per trillion dollars (EJ/$T) of economic activity in 1978, 7.1 EJ/$T in 2005 and 5.7
EJ/$T in 2017 (US Bureau of Economic Analysis, 2021). The decline rate of the energy intensity of the
economy has been a steady 1.6% per year over that entire time frame. The decreasing energy intensity
of the overall economy is due to the following factors:
• structural changes in the economy that favor lower energy intensity commercial activity such
as financial and computing/data-driven services over higher energy intensity industrial activity
such as iron and steel making;
• improvements in energy efficiency that deliver equivalent economic service for smaller energy
inputs such as:
{ improved heavy- and light-duty vehicle fuel efficiency;
{ improved insulation in residential and commercial buildings;
{ efficient devices and appliances such as LED lighting.
Figure 2.1 shows these trends graphically. Despite nearly 300% growth in real GDP (2012 dollars)
from ∼5.6 $T in 1975 to 19.1 $T in 2019, energy use grew only 40% over that period due to decreased
energy intensity, and carbon emissions have begun to decline from their 2005 peak due to both
decreased energy and carbon intensity.
There are multiple interrelated factors behind these trends including energy policy that incentivizes
sustainable energy use, the cost savings due to energy efficiency in many applications, innovation in
energy technology that improves efficiency and reduces emissions, and consumer preference for more
sustainable solutions. The remainder of this chapter examines some of these factors and provides
examples.
These trends are likely to accelerate into the future. In addition to continued expansion of the
lower-emitting and higher efficiency technologies listed above, the following trends and technologies
are beginning to roll out at scale in US energy markets. Their impact on overall energy consumption
and emissions, while not yet significant, will likely become visible in economy-wide statistics by 2025:
• remote working options (reduced local commuting and long-distance business travel, permanent
change initiated by 2020 pandemic);
• growth in online ordering or ‘e-commerce’ grew steadily over a decade and grew rapidly during
the pandemic, dramatically reducing the number of short-distance trips (DOE, 2020);
• electric vehicles (passenger cars and delivery vans);
• high efficiency electric heating (heat pumps).
Further drastic improvements in the energy efficiency and carbon intensity of economic activity are
possible with energy technologies that are technically feasible and have been demonstrated at scale,
but have not yet achieved performance and/or cost parity with competing technologies. Demand for
Figure 2.1 Carbon intensity of energy, energy intensity of the economy, carbon intensity of the economy, and
economy-wide carbon emissions relative to 2012 for the United States.
these technologies may increase substantially if certain policies are put in place or if a price on carbon
emissions is enacted:
• electricity generation with carbon capture and sequestration;
• hydrogen as a transportation fuel, heating fuel, or chemical process input;
• small modular nuclear reactors;
• biomass-derived energy (ethanol, other liquids, biogas, hydrogen or electricity) with carbon capture
and sequestration of process emissions.
Europe and some Asian economies have lower energy intensity than the US with similarly advanced
economies. While China lags the US and Europe in energy intensity, it is improving far more rapidly.
Economic energy intensity is far higher in the developing world, but total and per capita energy use is
dwarfed by the advanced economies.
Figure 2.2 A framework to qualitatively assess the drivers of sustainability in energy and water systems, with the
least meritorious on the left and most meritorious on the right.
practices (right half). The effectiveness of any specific energy/water intervention depends on both
the state of technology and of existing infrastructure. Stressors, behavior changes, technology
advancements, and the evolution of energy/water systems do not evolve linearly along this spectrum.
The histories of modernization and decarbonization include iterative loops and multi-step hops.
Examples for each of these categories are given in Table 2.2, illustrating the broad applicability of
this framework. Lessons learned from both successful and unsuccessful efforts to decarbonize energy
may be extended to inform the decarbonization of the water industry.
Table 2.2 Historical examples from each part of the sustainability framework that have alleviated the impacts of
energy and water use, demonstrating the adverse consequences of deprivation and benefits of innovation.
Table 2.2 Historical examples from each part of the sustainability framework that have alleviated the
impacts of energy and water use, demonstrating the adverse consequences of deprivation and benefits
of innovation (Continued).
Landfill gas recovery: Interception and recovery of methane from organic decay in landfills avoids high GWP
emissions. Conversion of landfill gas to electricity has a small side benefit of avoiding some fossil fuel use
Water Mitigation – Investments to reduce the emissions associated with wastewater treatment
Anaerobic digestion: Organic material is digested in a bio-reactor, producing streams of biogas and less
energetic sludge. The cost of installing and operating AD results in a useful energy product which may offset
fossil fuel use and lower environmental impacts from organic discharge
Energy Substitution – Use of an alternate resource with potential cost or reliability impacts
Cleaner fuels: Natural gas replaces coal in electricity generation, resulting in lower CO2 emissions per unit
electricity. When the price of natural gas per unit energy became comparable to coal, there was little reason
not to switch
Renewable electricity: Electricity from solar panels and wind turbines can displace fossil-fired generation.
Solar energy trades higher capital cost and inherent intermittency for zero fuel cost and no carbon
emissions
Water Substitution – Alternate resource or technology which may drive system reconfiguration
Non-potable reuse: Purple pipe systems deliver tertiary-treated wastewater to irrigation and some industrial/
cooling applications. This reduces the demand for freshwater supply, and may offset the energy used for
pumping and treatment of potable water
Ultraviolet disinfection: UV light can be substituted for the chlorine that is used to kill pathogens in water
supply or recycled wastewater. UV does not require chemical delivery or dosing equipment and avoids the
creation of disinfection byproducts. However, UV reactors incur significant upfront cost and ongoing energy
costs.
Energy Efficiency – Technological increase in services provided from equivalent resources
Aerodynamics: Refinements in vehicle shape reduce drag, enabling cars and trucks to go substantially farther
on the same amount of fuel with the same weight and volume of passenger/cargo capacity
Heat recuperation: Power generation and many industrial processes transfer thermal energy from exhaust to
intake. This process requires costly heat exchange equipment and results in substantial energy savings and
therefore emissions reduction
Variable speed drives: Novel electronics enable the motors that drive compressors and pumps to operate at
lower speeds (and therefore power) without loss in efficiency. This saves substantial energy during periods
when aeration or pumping needs are low
Water Efficiency – Increase in the benefits per unit water delivered
Drip irrigation: Replacing broadcast with drip irrigation drastically reduces the water lost to evaporation and
percolation, thereby reducing pumping and treatment requirements (and associated energy and emissions) for
water supply. However, drip systems are more costly to install and maintain
Energy Innovation – Delivers a better energy service for fewer resources consumed
LED lighting: New diode and phosphor materials enable drastically lower energy use at the same (or better)
quality and intensity of light, with longer life and lower heat generation
Hybrid and electric vehicles: Higher energy density and more durable batteries enable regenerative braking
and electric fueling, thereby increasing efficiency and reducing emissions. Electric cars are quieter and
eliminate local pollution. They accelerate and handle better than comparable conventional vehicles, and can
sometimes be fueled at home
Water Innovation – Delivers better treatment for less energy/material input
Membrane aerobic bioreactors: New materials and tube configurations enable oxygen delivery for organic
deconstruction in wastewater at much lower pumping energy than traditional aeration, thereby reducing
emissions
Water and energy investments exhibit a wide range of capital turnover rates. Capital turnover
for both water and energy equipment tends to be fastest in the residential and commercial end-use
sectors. Turnover of transportation equipment is slower. The large capital intensity of equipment in the
industrial and utility sectors tends to drive the slowest turnover rates. Energy and water distribution
and collection infrastructure is also designed for long service life and therefore very slow to change.
became prevalent. The US has retired almost half of its coal power capacity over the past decade,
declining from a more than 60% of total capacity to about 20%. Some utilities are looking to leverage
coal power plant infrastructure and retrofit coal plants with cleaner energy sources. As with coal,
operational nuclear power plants have almost all completed depreciated. With carbon-free emissions,
there are strong incentives to maintain the nuclear fleet. The challenge is to remain profitable in a
market where nuclear plants operate with constant output while demand and pricing are dynamic
due to growth in wind and solar generation. There has not been a new plant commissioned since the
partial meltdown of the Three Mile Island plant in 1979. There is one nuclear power facility under
construction in Georgia.
In comparison to the 50+ year capital turnover for thermoelectric power plants, renewable plants
tend to be both more distributed and have higher capital turnover rates. Wind turbines and solar
photovoltaic (PV) facilities are projected to have a lifetime of about 20 years. This is based on facilities
commissioned in the 1970s and 1980s that reached the end of their useful life in the 1990s and early
2000s. New wind turbines have nameplate capacities of 1–3 MW (rather than 10 s to 100 s of MW for
gas/thermal turbines), and new wind farms have total capacities of 10 s to 100 s of MW. Individual
solar panels have nameplate capacities in the 100 s of watts, making solar plant design and installation
extremely modular. With the ability to add capacity incrementally, wind and solar generation have
been growing steadily, and this trend is expected to continue. In comparison to wind and solar, capital
turnover in hydropower can be extremely long. Century-old hydropower plants are still in operation.
Dam-based hydropower plants can significantly disrupt wildlife, for example fish spawning. Recent
investments in hydropower have replaced dams with ‘run of the river’ systems to address society
demand and environmental regulations.
2.4.4 Integration
Capital turnover in the energy sector may present some unique challenges. For example, decisions
on vehicle electrification are expected to have a strong impact on both liquid fuels production and
power generation. A large increase in electricity demand for vehicles may trigger a new wave of capital
expenditures in the electric sector, and/or a major change in the operation of existing generation
and transmission assets. Similarly, a large drop in liquid fuel consumption will cause significant
disruptions to gasoline and ethanol markets (see section 2.5.3). There are not similar ‘fuel switching’
capital replacement options for water consumers.
The transition to energy efficient lighting was not without challenges though. Compact fluorescent
lamps (CFLs), an earlier generation of energy efficient lighting technology, failed to attain consumer
acceptance. CFLs were nearly as efficient as LEDs, saving 70–80% of lighting electricity over their
incandescent predecessors. However, CFLs were disliked by consumers because the quality of the
light they produced was inferior. CFL light had a high color temperature (bluish tint) and many
users perceived a flickering nature to it. CFLs were advertised as having much longer lives than
incandescents, but they burned out earlier than predicted. The failure of CFLs in the marketplace is
proof that consumers may be unwilling to trade quality of service for energy savings, even if there is
a comparable quantity of service and modest cost savings over the long term. Policy initiatives that
supported the transition to more efficient lighting (efficiency standards and incandescent ‘bulb bans’)
were met with fierce opposition when the only viable alternative to incandescent bulbs was CFL.
Government research institutions and private industry committed significant resources to developing
LED technology. Some of these investments were based on evidence that LED would ultimately be
a better technology. Some were responsive to consumer demand created by the policy incentives
described above. Today, it is nearly unthinkable to purchase incandescent lighting in the residential
sector for anything but the most niche applications. Manufacturing know-how has advanced so that
LED bulbs can be produced to meet the demands of almost any application and form factor.
of the new vehicle market. The largest global market for light-duty vehicles is China, and China has
the largest EV fleet. While the upfront costs for EVs are higher than ICEs, the total cost of ownership
when considering the cost of fuel, repairs, and vehicle life make EVs lower in cost than ICEs. At
current US energy prices and typical vehicle energy efficiencies, the fuel cost for electric vehicles is
much lower than for gasoline-powered cars. In comparison to ICEs, EVs have fewer moving parts,
generate less heat, and do not need to replace lubricating oils, cooling fluids, and brake pads as often.
Therefore, repairs are less frequent, and except for the replacement of the battery (∼10 years), vehicle
life is significantly longer, and maintenance is significantly cheaper.
However, as markets expand for light-duty EVs, it is becoming clear that the access to vehicle
charging will be a limiting factor on widespread EV adoption. With upfront costs of EVs higher than
ICE vehicles, most early adopters have been affluent buyers with ready access to overnight charging
in private garages. Less affluent drivers who live in urban and suburban rental units will not have
the same opportunity. Similarly, public charging infrastructure is being deployed in urban areas and
along high-use transportation corridors, so rural users are disadvantaged. Finally, taxi and delivery
drivers will have substantially different charging needs than the owners of vehicles whose use is
purely personal.
There are important lessons for the water industry in the adoption of electric vehicles. Transportation
is a contributor to carbon emissions in the US and light-duty EVs account for greater than 60% of
liquid fuel use. The world cannot achieve any meaningful decarbonization goals without transforming
the light-duty fleet. Note that electrifying the fleet will only achieve the decarbonization targets if the
vehicles are charged with carbon-free electricity. Similarly, electrifying energy input or unit operations
in the water and wastewater sectors will only be effective if the grid is decarbonized. Furthermore,
success requires investments in both fundamental science and engineering as well as underlying
infrastructure. Nascent science and technologies can grow into major business opportunities.
For example, Tesla has become the world’s most valuable vehicle manufacturer since before the
pandemic. However, certain performance targets must be met before a low-carbon technology will
be adopted at scale. In the case of electric vehicles, the performance target was the energy density
of the battery, and the market needed to wait for lithium-ion chemistry to be sufficiently advanced
(reliable, manufacturable) to be adopted. The water industry must identify performance targets for
decarbonized systems and seek investment in technologies that can reach those targets.
With the early mandates for corn starch ethanol, investors were incentivized to increase the size of
conventional biorefineries and production outpaced the targeted EISA volumes. Within a few years,
corn ethanol utilized 40% of corn crop production, largely achieving one of the original goals of
EPACT in supporting rural economic development. Ethanol rapidly achieved ∼10% volumetric blend
of the gasoline supply, extending the liquid fuel supply as the mandates targeted.
With cellulosic biofuels, progress was slower. Originally, the limiting technical factor was considered
enzymes to breakdown recalcitrant cellulosic into fermentable sugars. Cellulose is a structural polymer
composed of sugars monomers that are difficult to depolymerize. In comparison starch is a nutrient
source composed of readily digestible sugar polymers. As the science of cellulosic enzymes advanced,
other technical challenges were identified in the cellulosic biofuel process. When pioneer cellulosic
biorefineries were constructed, initial estimates were that they would have about twice the capital cost
per unit of product volume. With lower overall productivity, cellulosic biorefineries capital costs grew to
five- to ten-fold in comparison to mature starch ethanol biorefineries. This resulted in commercialization
delays and unique challenges to the cellulosic industry. Fourteen years after EISA created mandates and
incentives for cellulosic biofuels, the industry has yet to substantially impact decarbonization.
Two distinct challenges rapidly developed for the cellulosic and overall biofuel markets. The first
challenge is that the market is structurally limited in size. As corn starch ethanol production grew
rapidly, the US soon produced enough fuel to achieve 10% volume of the entire gasoline market. At
that time, most vehicles and most fuel infrastructure were limited to a 10% ethanol (E10) blend due to
materials compatibility. The conventional technology, which is relatively ineffective at decarbonization,
had exceeded the ability of the market to consume it. Flexible fuel vehicles (FFVs), which are capable of
using fuel with up to 85% ethanol fuel, were proposed as a solution. The manufacturing cost differential
is only about $100. The vehicle manufacturers received credits for the vehicles as if the vehicle always
used 85% ethanol (E85) fuel to meet fleet-wide corporate average fuel economy (CAFE) standards.
The fuels market did not have any incentive to market E85 fuels so FFVs continued to operate on
conventional E10 fuel. Therefore, FFVs resulted in only incidental increase in ethanol usage, and
therefore minimal impact on GHG reductions. Sixteen years after EPACT, corn ethanol accounts for
about 10% of the gasoline market and each gallon reduces GHGs by about 20%. Ethanol, therefore,
results in about ∼2% reduction in GHGs. EISA 2007 and cellulosic biofuels have had essentially no
additional impact on GHG emissions.
The second challenge in the biofuels market, and notably the cellulosic biofuel, is a significant
warning for the water sector. Realizing that cellulosic production was a nascent industry in 2007,
EISA created a regulatory requirement that the US EPA monitor cellulosic biofuel production capacity
on an annual basis. The EISA mandate for cellulosic fuels is adjusted annually to avoid fuel blenders
being mandated to use cellulosic biofuels that do not exist. Since the first blending requirements,
manufacturing capacity has lagged blend mandates, so EPA adjusted the volumetric requirements. A
cellulosic biorefinery is a complex operation and requires several years to construct and deploy. One
of the authors of this chapter interviewed project investment banks and described the EISA mandates,
EPA regulatory role, and the time and cost to build (Blazy et al., 2015). It was uniformly considered
a poor investment decision. Therefore, few cellulosic biorefinery projects have even been launched,
and there is little-to-no success in the industry. The water sector should learn that mandates without
consideration of markets, economics, fundamental science, state of technology, and the full scope of
the mandates could lead to failed investments and little progress in achieving decarbonization goals.
release no carbon emissions and have no fuel requirements. The need for fuel creates supply chain
risks and also adds fuel price volatility risk to the total cost. Together, renewables are, at the time this
book is published, the second largest generator of power in the US after natural gas, catching up to
nuclear and surpassing coal.
The growth in wind and solar demonstrate a technology ‘learning curve’ that water decarbonization
may emulate.
Driven by incentives including both investment tax credits (ITCs) and production tax credits
(PTCs), wind capacity has exhibited the largest overall increase in capacity of any type of generation.
This rapid build-out has catalyzed ‘learning-by-doing,’ and the modular nature of wind power allows
for continuous innovation in the design, manufacturing, and construction of turbines. Focused
R&D in wind has resulted in only incremental improvements, but for wind technology, incremental
improvement delivers outsized gains in performance. For example, the generation capacity of a turbine
scales as the square of the length of its blades. Therefore, small increases in blade length enabled by
novel designs and materials have resulted in a non-linear increase in turbine capacity. Similarly, taller
towers enable turbines to access more reliable wind resources. Higher reliability translates to a more
valuable electricity resource, in addition to the bulk increase in kWh generated. Ultimately, these
improvements increase land use efficiency.
There has been enormous research investment in new PV solar materials, however, no new
materials have been deployed at scale. Rather, the dramatic drop in prices (and commensurate rapid
build-out of PV-based generation) has been driven largely by reductions in manufacturing costs of
conventional solar materials (polysilicon, and to a lesser extent cadmium telluride) and installation
costs. China has driven the reduction in manufacturing costs. The ITC accelerated the domestic
market for installation, and again, learning-by-doing drove down installation costs. Each new PV
installation enabled incremental innovations in racking, interconnection, and construction logistics
for ground-mount and roof-mount systems. As this book is published, solar PV offers the lowest cost
of power costs in sunny regions such as the US southwest. However, because generation peaks during
the middle of the day and demand peaks at other times, the value of additional solar installation is
beginning to decline in areas with substantial solar penetration (Bolinger et al., 2021).
The challenge to both wind and solar is intermittency. The ultimate solution is to link intermittent
renewable power production to energy storage. Energy storage includes batteries, supercapacitors,
pumped hydropower, other mechanical systems, or even thermal systems. The value of grid-scale
energy storage is less than battery electric vehicles, so grid storage is learning and adapting from EVs.
Renewable power generation provides an important template for decarbonizing the water sector.
Significant advancements in technology were not required for transition in the market. Rather, policy
incentives drove the economics enough to foster capacity expansion. Increased installations drove
manufacturing and infrastructure support down a learning curve to further incentivize deployment.
For several years, incentives made renewable power the lowest cost pathway to increase capacity. As the
manufacturing and installation processes matured, the ITCs were no longer required. As capacity grew
to where it was disrupting grid resilience, storage and other mechanisms are developing as solutions.
Incentives for storage and now driving storage capacity growth. It is largely expected that renewable
power plus grid-scale storage will offer a cost-competitive and reliable carbon-free power sector.
ACKNOWLEDGEMENTS
This work was supported by the US Department of Energy through contract DE-AC07-05ID14517
(Idaho National Laboratory). This work was performed under the auspices of the US Department
of Energy by Lawrence Livermore National Laboratory under Contract DE-AC52-07NA27344.
This document was prepared as an account of work sponsored by an agency of the United States
government. Neither the United States government nor Lawrence Livermore National Security, LLC,
nor any of their employees makes any warranty, expressed or implied, or assumes any legal liability or
responsibility for the accuracy, completeness, or usefulness of any information, apparatus, product, or
process disclosed, or represents that its use would not infringe privately owned rights. Reference herein
to any specific commercial product, process, or service by trade name, trademark, manufacturer, or
otherwise does not necessarily constitute or imply its endorsement, recommendation, or favoring
by the United States government or Lawrence Livermore National Security, LLC. The views and
opinions of authors expressed herein do not necessarily state or reflect those of the United States
government or Lawrence Livermore National Security, LLC, and shall not be used for advertising or
product endorsement purposes.
REFERENCES
Blazy D., Pearlson M. N., Miller B. and Bartlett R. E. (2015). A Monte Carlo-Based Methodology for Valuing
Refineries Producing Aviation Biofuel, Chapter 15. Royal Society of Chemistry, Cambridge UK, pp. 336–351.
https://doi.org/10.1039/9781782622444-00336
Bolinger M., Seel J., Warner C. and Robson D. (2021). Utility Scale Solar 2021 Technical Brief. Lawrence Berkeley
National Lab, Berkeley, CA. https://emp.lbl.gov/sites/default/files/utility-scale_solar_2021_technical_brief.
pdf (accessed 2 November 2021)
DHS (2021). National Critical Functions. U.S. Department of Homeland Security, Washington, DC. https://www.
dhs.gov/cisa/national-critical-functions (accessed 30 October 2021)
Dieter C. A., Maupin M. A., Caldwell R. R., Harris M. A., Ivahnenko T. I., Lovelace J. K., Barber N. L. and Linsey
K. S. (2018). Estimated use of Water in the United States in 2015. In USGS (Circular 1441). U.S. Geological
Survey, Reston VA. https://doi.org/10.3133/cir1441
DOE (2014). The Water-Energy Nexus: Challenges and Opportunities. U.S. Department of Energy, Washington
DC. https://www.energy.gov/sites/prod/files/2014/07/f17/Water Energy Nexus Executive Summary July
2014.pdf (accessed 20 April 2020)
DOE (2020). Moving Goods in A SMART Mobility System. U.S. Department of Energy, Washington DC. https://
www.energy.gov/eere/vehicles/downloads/smart-webinar-6-moving-goods-smart-mobility-system (accessed
30 October 2021)
EIA (2021). Monthly Energy Review. U.S. Department of Energy, Washington DC. http://www.eia.gov/totalenergy/
data/monthly no (accessed 24 November 2020)
EPA (2007). Energy Independence and Security Act. Public Law 110-140. U.S. Environmetnal Protection Agency,
Washington DC. https://www.epa.gov/laws-regulations/summary-energy-independence-and-security-act
(accessed 11 January 2022)
EPA (2021). Inventory of U.S. Greenhouse Gas Emissions and Sinks 1990–2019 (No. EPA430-R-21-005). U.S.
Environmental Protection Agency, Washington DC. https://www.epa.gov/ghgemissions/inventory-us-green
house-gas-emissions-and-sinks-1990-2019 (accessed 30 October 2021)
EPRI (2013). Electricity Use and Management in the Municipal Water Supply and Wastewater Industries (No.
3002001433). Electric Power Research Institute, Palo Alto CA. https://publicdownload.epri.com/PublicDown
load.svc/product=000000003002001433/type=Product (accessed 14 January 2020)
Gleick P. H. (1994). Water and energy. Annual Review of Energy and the Environment, 19(1), 267–299. https://
doi.org/10.1146/annurev.eg.19.110194.001411
Greenberg H. R., Simon A. J., Singer S. L. and Shuster E. P. (2017). Development of Energy-Water Nexus State-
Level Hybrid Sankey Diagrams for 2010. In LLNL (LLNL-TR-669059). Lawrence Livermore National Lab,
Livermore CA. https://flowcharts.llnl.gov/report (accessed 7 January 2020)
Grubert E. and Sanders K. T. (2018). Water use in the United States energy system: A national assessment and
unit process inventory of water consumption and withdrawals (research article). Environmental Science and
Technology, 52(11), 6695–6703. https://doi.org/10.1021/acs.est.8b00139
Hermann W. (2006). Quantifying global exergy resources. Energy, 31(12), 1685–1702. https://doi.org/10.1016/j.
energy.2005.09.006
Kyung D., Kim M., Chang J. and Lee W. (2015). Estimation of greenhouse gas emissions from a hybrid wastewater
treatment plant. Journal of Cleaner Production, 95, 117–123. https://doi.org/10.1016/j.jclepro.2015.02.032
Roy M., Edwards M. R. and Trancik J. E. (2015). Methane mitigation timelines to inform energy technology evaluation.
Environmental Research Letters, 10(11), 114024. https://doi.org/10.1088/1748-9326/10/11/114024
Szulc P., Kasprzak J., Dymaczewski Z. and Kurczewski P. (2021). Life cycle assessment of municipal wastewater
treatment processes regarding energy production from the sludge line. Energies, 14(2), 356. https://doi.
org/10.3390/en14020356
US Bureau of Economic Analysis (2021). National Income and Product Accounts, Table 1.1.4 Price Indexes for
Gross Domestic Product and Table 1.1.5. Gross Domestic Product. U.S. Bureau of Economic Analysis,
Suitland MD. https://apps.bea.gov/iTable/index_nipa.cfm (accessed 30 October 2021)
USDA (2021a). CORN Quickstats: Acreage, Yield, and Price. National Agricultural Statistics Service. U.S.
Department of Agriculture, Washington DC. https://www.nass.usda.gov/Quick_Stats/Lite/result.php?B56
5104E-D426 -36E7-9481-03007890E2F2 (accessed 8 October 2021)
USDA (2021b). Wheat Sector at A Glance. USDA Economic Research Service. U.S. Department of Agriculture,
Washington DC. https://www.ers.usda.gov/topics/crops/wheat/wheat-sector-at-a-glance/ (accessed 8
October 2021)
USGS (2021). Mineral Commodity Summaries 2021. U.S. Geological Survey, Reston, VA. https://doi.org/10.3133/
mcs2021
Chapter 3
Greenhouse gases in the urban
water cycle
Andrew Shaw1*, Kubeshnee Chetty2, Tak Fan Chan2, Elena Lindsey2, Anjana Kadava1 and
Ben Stevenson3
1Black & Veatch Corporation, 920 Memorial City Way, Suite 600, Houston, TX 77024, USA
2Binnies, 60 High Street, Redhill, Surrey, RH1 1SH, UK
3Black & Veatch Ltd, Queens House, 19–29 St. Vincent Place, Glasgow, G1 2DT, UK
*Correspondence: [email protected]
3.1 INTRODUCTION
This chapter gives an overview of greenhouse gas (GHG) emissions in the context of the urban water
cycle. Starting with an overview of the urban water cycle and a definition of different equivalent
carbon emission scopes, the chapter then gives a description of the major GHGs in each part of the
urban water cycle, in order to identify opportunities for decarbonization. In the latter sections of
the chapter a framework is presented for carrying out a carbon footprint assessment along with an
overview of available and relevant protocols and methods for assessing GHG emissions.
© 2022 The Editors. This is an Open Access eBook distributed under the terms of the Creative Commons Attribution Licence
(CC BY-NC-ND 4.0), which permits copying and redistribution for non-commercial purposes with no derivatives, provided the
original work is properly cited (https://creativecommons.org/licenses/by-nc-nd/4.0/). This does not affect the rights licensed or
assigned from any third party in this book. The chapter is from the book Pathways to Water Sector Decarbonization, Carbon Capture
and Utilization, Zhiyong Jason Ren and Krishna Pagilla (Eds.)
After abstraction, water is treated to potable standards, the quality of which is dependent on
national or regional requirements. There are various treatment technologies in use today, selected
based on site specific conditions and on the treatment standards required. The advancement of
technological processes does not directly correlate to higher emissions, as modern processes seek to
reduce emissions. However, Bakhshi (2009) posits that energy use, and therefore emission release,
will increase as various water acts (e.g. the Safe Drinking Water Act) demand higher quality water.
Whether using distributed or centralized treatment plants, storage facilities, pumps, and pipes
are required to distribute the clean water to the end user. Although some systems rely solely on
gravity, most require pumping. Pumping maintains the pressure and movement of water to not only
relocate the water but to minimise corrosion and contamination of the pipe network (Klein, 2005).
This can often be an energy intensive process to achieve the required system pressure throughout the
distribution network. Due to the age of many water networks, leaks and issues with infrastructure
contribute to increased energy use within the associated systems. Moreover, energy use is primarily
driven by increased demand and urban growth.
Water is consumed by homes, businesses and industries, with further energy use within them:
treatment, circulation, heating, cooling. Surface water run-off can be considered in the ‘use’ category
although its source – precipitation – is from the natural water cycle. Surface water runoff and
stormwater management thus become processes of the urban water cycle. Wastewater from a home
or industry that does not contain fecal contamination is called grey water. Grey water can be cleaned
and used onsite, for example for cooling processes.
Wastewater collection systems often use gravity to move wastewater to a facility to be treated
typically by positioning the wastewater treatment facility ‘downstream’ of the urban area. However,
urban areas that are large or very flat will require pumping of wastewater which significantly increases
their energy use.
As the incoming water is ‘dirty’, wastewater requires more energy to treat than fresh water. The type of
waste treatment depends on the final discharge or reuse point and therefore the energy and greenhouse
gas emissions vary widely. Wastewater originates from a variety of sources, varying in contamination
from industrial processes to surface water runoff. Therefore, the treatment of wastewater requires
more consideration than water treatment. Wastewater can be cleaned sufficiently for several reuse
applications, including advanced wastewater treatment to produce potable water without returning
water to its natural source, that is potable water reuse.
Water is discharged into surface water bodies for integration back into the natural water cycle or
for reuse by the urban water cycle.
Figure 3.2 Global GHG emissions by sector in 2014, Mt CO2e (based on Rissman et al., 2020).
Figure 3.3 Sankey diagram of energy use in the whole water cycle process in Beijing in 2015 (adapted from He, 2019).
This chapter will provide an overview of scope 1, 2 and 3 emissions associated with abstraction,
water and wastewater treatment and biosolids management process. To truly achieve net zero carbon
emissions, policy makers will need to switch focus from supply and treatment to the consumer side of
the equation and target interventions at either reducing or making consumption more efficient, thereby
reducing the GHG emissions from the whole water life cycle and not just on conveyancing and treatment.
The water industry in the UK reported to have achieved 45% emission reduction of 2.4 MtCO2e
between 2011 and 2018 and consumed 6.8 TWh of electricity for a population of approximately 67
million. Most of the emissions are CO2 associated with the electricity, which is approximately 2% of
the electricity consumption in the UK to pump water to customer and wastewater treatment. The
industry has pledged to reach net zero carbon emissions by 2030 (Water UK, 2020).
bacteria produce N2O in aerobic conditions. A small percentage of N2O is also generated from onsite
grit and sludge storage tanks.
Nitrous oxide (N2O) and methane (CH4) are potent GHG gases with global warming potentials
(expressed in terms of CO2-equivalents) of 265 and 28, respectively. When emitted to the atmosphere,
they significantly contribute to climate change. Research by Oshita et al. (2014), show that most of
the methane emissions from WWTPs are closely related to processes involved in the sludge line. CH4
emissions mainly arise from the anaerobic digestion process, its associated process units such as
primary sludge thickener, the centrifuge, the exhaust gas of the cogeneration plant, the buffer tank for
the digested sludge, and the storage tank for the dewatered sludge. Daelman et al. (2012) found that
around 1% of incoming chemical oxygen demand (COD) to waste water treatment works was emitted
as methane and that the sludge process units contribute to 72% of the methane emissions from a waste
water treatment works.
Due to the apparent inconsistency between the CAW methodology for calculating process emission
and current GHG emission accounting practice, there is an uncertainty around the scale of process
emissions in the UK. This is partly due to an update to the CAW methodology which can increase
scope 1 emissions by 0.2 MtCO2e/y (∼8% of UK water sector’s net emission) and an associated increase
in baseline process emissions by ∼30%.
Fundamentally, the core focus for reduction of operational carbon emissions generally relates to
the modification or optimization of the operational conditions of the treatment plant as this is the
most economical and cost-efficient method, however this is not always possible due to the operational
limitations of the installed units. Other ways to mitigate GHG emissions also include treatment of
gaseous streams or the installation of new processes to remove both organic matter and pollutants.
Sludge management: Advanced digestion treatment methods are used to process sludge generated
onsite as well as imported from other sites to enhance the biogas (CH4) yield from the sludge treatment
process. The biogas can be used in onsite combined heat and power engines (CHP) to generate
renewable energy. Other uses for biogas also include gas to grid applications.
The stabilized and dewatered biosolids is a valuable product as fertilizer and soil conditioner for
agricultural use. There are biosolids assurance schemes, code of practices and regulations in different
countries to ensure its safe recycling.
3.2.2.1 Pumping
The energy required for abstraction depends on the source of water. Pumping is often required to lift
groundwater from the water table, while energy for abstracting surface water depends on the distance
and profile between the source and the treatment facility.
After the water has been used by the consumers the used water or wastewater is collected and,
depending on the local ground profile and the relatively location of the wastewater treatment facility,
pumped through a network or sewer for treatment to remove the pollutants. The treated wastewater is
often discharged into the local receiving water by gravity and pumped discharge is not often required
for most works.
processes chemical dosing, mixing and filtration systems consume energy. Advanced processes
such as membrane filtration, oxidation, disinfection by UV and ozonation are more energy
intensive. Desalination is highly energy intensive even with the introduction of energy efficient
reverse osmosis membrane and the move from the desalination of sea water to brackish or treated
wastewater.
A significant amount of energy is required to distribute treated water into the potable network.
However, reportedly 20% of the water put into supply in England is lost through leakage compared
with approximately 3 and 5% reported in Germany and Singapore respectively (C40 Cities, 2021).
activities both upstream and downstream of the water treatment and supply value chain. Areas that
generally apply to scope 3 emissions are as follows:
• purchased goods & services;
• fuel and energy related activities;
• transportation and distribution (both upstream and downstream);
• treatment of waste generated;
• business travel – using indirect sources such as planes, trains and so on.;
• employee commuting;
• lease/hired equipment;
• use of sold products;
• end of life treatment of sold products.
With emphasis now shifting towards full accountability and net zero emissions, more organisations
are reaching into their value chain to understand the full GHG impact of the operation. Although
the accounting and understanding of scope 3 emissions are not in water organisation’s direct control,
this will present an opportunity for the reduction of overall GHG emissions, as organisations can
influence its suppliers or streamline procurement by contracting with vendors who fully account for
their GHG emissions.
3.3 PROTOCOLS
Protocols consist of a set of standardized frameworks to estimate the GHG emissions from a process or
an activity. Protocols typically use a set of emissions factors (EF) that relate a task or a process to the
amount of GHG emitted by a similar standard process. There are specific protocols that are applicable
to a certain industry in a specific region, so selection depends on location and purpose of the project.
A protocol also provides the guidelines to define goal and boundary conditions for a project and
categorizes the activities into various scopes. These protocols are listed in the sections below.
The various updates to the original guidelines are presented Table 3.1 below.
water treatment facilities, planners and stakeholders. LGOP has been officially accepted as the
standard reporting tool for local agencies within the United States. Two versions of LGOP have been
published to date:
(1) Version 1 – Originally adopted in 2008
(2) Version 1.1 – Revised in May 2009
for an organisation to deploy a measurement campaign or to actively measure emissions, thus reducing
the resource requirements. Published EFs can be obtained through specific databases, such as the
IPCCs ‘Emission Factor Database’, or from relevant literature. Calculated or measured EFs can be
used to reflect a more representative emission for a specific process or material.
Given that the significant proportion of the water industry’s emissions come from energy and
transport, the use of EFs is an effective approach. The collection of activity data is far simpler and
cost effective than measuring emissions. Much of the data is already widely available to organisations
through electricity use, fuel purchases and vehicle movements. This extends to the quantification of
embodied carbon, whereby EFs for materials are widely available. For example, the University of Bath
has developed an open-access database to collate and report per-reviewed energy and carbon values
of various materials (Hammond & Jones, 2008).
Global average EF data is currently readily available for a range of good or services, and EFs at
a national level are becoming increasingly reported allowing for better representative data (Ercin
& Hoekstra, 2012). EFs are often regularly reviewed and updated to reflect the most accurate
emission intensity. This is particularly useful when considering emissions sources such as electricity
consumption, whereby the source of electricity can change seasonally and annually, particularly as
distribution and transmission systems decarbonize.
In the UK, water companies in England and Wales are obligated to report annual (operational)
GHG emissions via the CAW. This approach uses published emission factors and requires companies
to submit operational activity data, thus allowing for a simplified and standardised method of
estimating operational GHG emission.
The simplicity of using emission factors has its limitations. Site specific conditions are not factored
in, and therefore any use of generalized emission factors can lead to over- or under-estimates on smaller
scales. For example, N2O emissions associated with WWTPs are highly dependable on operating
conditions and therefore will vary both spatially and temporally (Law et al., 2012; Parravicini et al.,
2016). The IPCC have recently acknowledged that N2O emissions were likely being underestimated
under the 2006 guidelines and provided revised emission factors and quantification methodologies.
The revised EF is a result of further appreciation for emissions during the treatment process and
disposal of effluent to aquatic environments (IPCC, 2019). Wallace et al. (2020) reported that this
revision led to a 40-fold increase in calculated N2O emissions for a WWTP in Christchurch, New
Zealand, which substantially increased the plant’s overall carbon footprint, thus highlighting the
sensitivity in GHG reporting to the accuracy and representativeness of EFs.
The sources and sinks of methane are poorly understood and there is a growing international effort
to better understand current emission rates. The quantification of methane for emission inventories is
primarily achieved through the use of emission factors, however, there is a growing body of research
assessing the accuracy and reliability of using satellites or aircraft to quantify methane emissions
at both local and regional scales. Satellite data can be used to construct models to infer methane
emissions (e.g., Jacob et al., 2016; Turner et al., 2015). Meanwhile, methane can either be sampled (e.g.,
Schwietzke et al., 2017) or inferred using lidar (e.g., Riris et al., 2012) from aircrafts. However, much
of the focus of published research is on regional methane emissions, and any studies on identifying
individual sources tends to centre on releases from natural sources, such as wetlands, or large
anthropogenic emissions sources such as natural gas plants. As such, there appears to be no studies
dedicated to using remote sensing techniques to specifically quantify methane emissions associated
from wastewater treatment plants.
3.4.3 Models
Developing mathematical models to aid in the quantification of GHG emissions from wastewater
plants offers a solution to the use of generalised emission factors or resource-intensive measurement
campaigns. Modelling aims to identify and help better understand the various complex relationships
along the treatment process pathway, thus paving the way for solutions that can ultimately reduce the
carbon footprint (Mannina et al., 2016). The boundaries of a model can vary from the analysis of a
single process or component within a wastewater treatment plant, up to plant-wide approaches that
consider all processes and the interactions between them (Mannina et al., 2016).
Modelling approaches to estimating GHG emissions from wastewater treatment plants can be
divided into categories based on complexity. Simple comprehensive process models have been shown
to be effective in helping identify factors that influence GHG emissions, but their simplicity requires
assumptions that ultimately impact the accuracy of results (Mannina et al., 2016). Furthermore, they
are restricted to steady-state analysis, which limits the effectiveness in quantifying N2O emissions due
to their dynamic nature (Corominas et al., 2012).
Dynamic process-based models have been shown to be more effective at capturing the variability
in GHG emissions, largely a result of the dynamic nature of N2O emissions. System configuration,
operating settings and atmospheric conditions influence the release of GHGs, all of which are difficult
to account for in simple models (Corominas et al., 2012). Dynamic process-based models are complex
and require both high computational power and large amounts of data for calibration. However, when
deployed at plant-wide scales, they offer the potential to improve the description and quantification
in GHG emissions. Current knowledge of N2O formation is a clear limiting factor in the accuracy of
such models (Corominas et al., 2012; Mannina et al., 2016). The direct measurement and monitoring
of N2O emissions, alongside the collection of various other parameters such as pH and temperature,
via sites’ SCADA systems, can be used to develop and improve models to better understand dynamic
relationships within wastewater treatment processes that ultimately influence emission rates (Baresel
et al., 2016; Law et al., 2009).
Wallace et al. (2020) applied a mechanistic model of a WWTP to determine the N2O emission rate.
The calculated value was 25% lower than the N2O emission factor reported by the IPCC.
Figure 3.4 Decision framework for carbon footprint analysis (from Pagilla, 2009).
(4) Identify carbon sources and sinks – Having defined the boundaries, carbon sources and sinks
can be identified, using approaches that are acceptable for the selected protocol.
(5) Collect and assess data – Often the most labour-intensive and costly part of the project is
collecting and assessing the data needed to carry out the carbon footprint calculation. Data
gaps may be identified requiring additional measurements to be made, or alternative data to
be gathered to provide a different method for estimating emissions (e.g. developing a process
model if no direct measurements are available).
6. Select method and calculate carbon footprint – Once all available data is gathered, the carbon
footprint can be calculated using appropriate methods. In general, the preferred method is to
use direct measurements if possible, followed by estimated emissions based on other facility
data. If neither are available, then macro-scale estimates can be made based on population
equivalents.
7. Identify reduction strategies – Carbon reduction may or may not be an explicit objective of
the carbon footprint development, but regardless, having completed the calculation, this is the
ideal point at which to identify potential hot spots and opportunities for reducing them.
8. Communicate the results – Displaying results in a clear and meaningful way is an important
last step in developing the carbon footprint. In many instances the final audience for the results
may not be experts in carbon emissions and so it is important that terminology, assumptions
and the significance of the results are communicated well.
Many water and wastewater facilities also have land (e.g. for buffers) on which solar or wind turbines
can be installed for renewable energy generation.
REFERENCES
Appelbaum B. (2002). Water and sustainability (vol. 4). US electricity for consumption for water supply and
treatment systems. Urban Water, 4, 153–161, https://doi.org/10.1016/S1462-0758(02)00014-6
Bakhshi A. A. (2009). Carbon Footprint Estimation of Municipal Water Cycle. George Mason University.
Mason Archival Repository. Retrieved from Mason Archival Repository: http://mars.gmu.edu/xmlui/
handle/1920/5659 (accessed 25 February 2022)
Baresel C., Andersson S., Yang J. and Andersen M. H. (2016). Comparison of nitrous oxide (N2O) emissions calculations
at a Swedish wastewater treatment plant based on water concentrations versus off-gas concentrations.
Advances in Climate Change Research, 7(3), 185–191, https://doi.org/10.1016/j.accre.2016.09.001
C40 Cities (2021). The C40 Cities Website. Retrieved from https://www.c40.org/ (accessed 25 February 2022)
Campos J. L., Valenzuela-Heredia D., Pedrouso A., Val del Rio A., Belmonte M. and Mosquera-Corral A. (2016).
Greenhouse gases emissions from wastewater treatment plants: minimization, treatment and prevention.
Journal of Chemistry, 12, 1–12.
Corominas L., Flores-Alsina X., Snip L. and Vanrolleghem P. A. (2012). Comparison of different modeling approaches
to better evaluate greenhouse gas emissions from whole wastewater treatment plants. Biotechnology and
Bioengineering, 109, 2854–2863.
Daelman M. R., Voorthuizen E. M., Dongen U. G., Volcke E. I. and Loosdrecht M. C. (2012). Methane emission
during municipal wastewater treatment. Water Research, 46(11), 3657–3670, https://doi.org/10.1016/j.
watres.2012.04.024
Eijo-Río E., Petit-Boix A., Villalba G., Suarez-Ojeda M. A., Amores M. J., Aldea X., Rieradevall J. and Gabarell
X. (2015). Municipal sewer networks as sources of nitrous oxide, methan and hydrogen sulphide emissions:
A review and case studies. Journal of Environmental Chemical Engineering, 3(3), 2084–2094, https://doi.
org/10.1016/j.jece.2015.07.006
Ercin E. and Hoekstra A. Y. (2012). Carbon and Water Footprints: Concepts, METHODOLOGIes and Policy
Responses (World Water Assesment Programme; No. 4). United Nations Educational, Scientific and Cultural
Organization (UNESCO), 7, place de Fontenoy, 75352 Paris 07 SP, France.
Ghaemi Z. and Smith A. D. (2020). A review on the quantification of life cycle greenhouse gas emissions at urban
scale. Journal of Cleaner Production, 252, 119634, https://doi.org/10.1016/j.jclepro.2019.119634 (accessed
25 February 2022)
Gu Y., Dong Y.-N., Wang H., Keller A., Xu J., Chiramba T. and Li F. (2016). Quantification of the water, energy and
carbon footprints of wastewater treatment plants in China considering a water-energy nexus perspective.
Ecological Indicators, 60, 402–409, https://doi.org/10.1016/j.ecolind.2015.07.012
Guohua H., Yong Z., Jianhua W., Yongnan Z., Shan J., Haihong L. and Qingming W. (2019). The effects of urban
water cycle on energy consumption in Beijing, China. Journal of Geographical Sciences, 29(6), 959–970,
https://doi.org/10.1007/s11442-019-1639-5
Hammond G. P. and Jones C. I. (2008). Embodied energy and carbon in construction materials. Proceedings of the
Institution of Civil Engineers – Energy, 161(2), 87–98, https://doi.org/10.1680/ener.2008.161.2.87
Hertwich G. E. and Wood R. (2018). The growing importance of scope 3 greenhouse gas emissions from industry.
Environmental Research Letters, 13(10), 104013, https://doi.org/10.1088/1748-9326/aae19a
Hwang K.-L., Bang C.-H. and Zoh K. D. (2016). Characteristics of methane and nitrous oxide emissions
from the wastewater treatment plant. Bioresource Technology, 214, 881–884, https://doi.org/10.1016/j.
biortech.2016.05.047
IPCC (2014a). Annex I: glossary. In: IPCC, Climate Change 2014: Synthesis Report, K. J. Mach, S. Planton and C.
von Stechow (eds.), IPCC, Geneva, pp. 117–130.
IPCC (2019). Refinement to the 2006 IPCC Guidelines for National Greenhouse Gas Inventories, volume 5,
Waste. Available at: https://www.ipcc-nggip.iges.or.jp/public/2019rf/vol5.html (accessed 25 February 2022)
Jacob D. J., Turner A. J., Maasakkers J. D., Sheng J., Sun K., Liu X., Chance K., Aben I., McKeever J. and Frankenberg
C. (2016). Satellite observations of atmospheric methane and their value for quantifying methane emissions.
Atmospheric Chemistry and Physics, 16, 14371–14396. https://doi.org/10.5194/acp-16-14371-2016
Klein G. (2005). California’s Water – Energy Relationship. California Energy Commission, CA, USA.
Law Y., Ni B.-J., Lant P. and Yuan Z. (2012). N2O production rate of an enriched ammonia-oxidising bacteria
culture exponentially correlates to its ammonia oxidation rate. Water Research, 46(10), 3409–3419, https://
doi.org/10.1016/j.watres.2012.03.043
Mannina G., Ekama G., Caniani D., Cosenza A., Esposito G., Gori R., Garrido-Baserba M., Rosso D. and Olsson
G. (2016). Greenhouse gases from wastewater treatment – A review of modelling tools. Science of the Total
Environment, 551–552, 254–270, https://doi.org/10.1016/j.scitotenv.2016.01.163
Nair S., George B., Malano H. M., Arora M. and Nawarathna B. (2014). Water-energy-greenhous gas nexus of
urban water systems: review of concepts, state-of-art and methods. Resources, Conservation and Recycling,
89, 1–10, https://doi.org/10.1016/j.resconrec.2014.05.007
Oshita K., Okumura T., Takaoka M., Fujimori T., Appels L. and Dewil R. (2014). Methane and nitrous oxide
emissions following anaerobic digestion of sludge in Japanese sewage treatment facilities. Bioresource
Technology, 171, 175–181, https://doi.org/10.1016/j.biortech.2014.08.081
Pagilla K., Shaw A., Kunetz T. and Schiltz M. (2009). A systematic approach to establishing carbon footprints
for wastewater treatment plants. Proceedings of the Water Environment Federation, 2009(10), 5399–5409.
https://doi.org/10.2175/193864709793952350
Parravicini V., Svardal K. and Krampe J. (2016). Greenhouse gas emissions from wastewater treatment plants.
Energy Procedia, 97, 246–253, https://doi.org/10.1016/j.egypro.2016.10.067
Riris H., Numata K., Li S., Wu S., Ramanathan A., Dawsey M., Mao J., Kawa R. and Abshire J. B. (2012). Airborne
measurements of atmospheric methane column abundance using a pulsed integrated-path differential
absorption lidar. Applied Optics, 51, 8296–8305, https://doi.org/10.1364/AO.51.008296
Rissman J., Bataille C., Masanet E., Aden N., Morrow W. R., Zhou N. and Williams E. D. (2020). Technologies and
policies to decarbonize global industry: review and assessment of mitigation drivers through 2070. Applied
Energy, 266, 114848, https://doi.org/10.1016/j.apenergy.2020.114848
Schwietzke S., Pétron G., Conley S., Pickering C., Mielke-Maday I., Dlugokencky E. J., Tans P. P., Vaughn T., Bell
C., Zimmerle D., Wolter S., King C. W., White A. B., Coleman T., Bianco L. and Schnell R. C. (2017). Improved
mechanistic understanding of natural gas methane emissions from spatially resolved aircraft measurements.
Environmental Science & Technology, 51(12), 7286–7294, https://doi.org/10.1021/acs.est.7b01810
Thames Water (2020). Combined Report – Annual Report, Annual Performance Report and Sustainability Report
2019/20. Available at: https://www.thameswater.co.uk/media-library/home/about-us/investors/our-results/
current-reports/annual-report-2019-20.pdf (accessed 21 February 2021)
Turner A. J., Jacob D. J., Wecht K. J., Maasakkers J. D., Lundgren E., Andrews A. E., Biraud S. C., Boesch H.,
Bowman K. W., Deutscher N. M., Dubey M. K., Griffith D. W. T., Hase F., Kuze A., Notholt J., Ohyama H.,
Parker R., Payne V. H., Sussmann R., Sweeney C., Velazco V. A., Warneke T., Wennberg P. O. and Wunch D.
(2015). Estimating global and North American methane emissions with high spatial resolution using GOSAT
satellite data. Atmospheric Chemistry and Physics, 15, 7049–7069. https://doi.org/10.5194/acp-15-7049-2015
United Utilities (2020). Annual Report and Financial Statements for the Year Ended 31 March 2020. Available at:
https://www.unitedutilities.com/corporate/investors/results-and-presentations/annual-reports/ (accessed
21 February 2021)
Wallace E., Bouman R., Bourke M. and Benter-Lynch M. (2020). IPCC emissions factors – Challenging a fortyfold
jump in nitrous oxide. Proceedings of the Water New Zealand Conference, 17–19 November 2020, Hamilton,
New Zealand.
Water UK (2020). Net Zero 2030 Roadmap. Retrieved from https://www.water.org.uk/routemap2030/wp-content/
uploads/2020/11/Water-UK-Net-Zero-2030-Routemap.pdf (accessed 25 February 2022)
WRF (2014). Water Footprint: A New Concept for Sustainable Water Utilities. Water Research Foundation Project
#4378, Completed 2014. Available at: https://www.waterrf.org/research/projects/water-footprint-new-concept-
sustainable-water-utilities (accessed 25 February 2022)
Zawartka P., Burchart-Korol D. and Blaut A. (2020, April 2). Model of Carbon Footprint Assessment for the Life
Cycle of the System of Wastewater Collection, Transport and Treatment. Retrieved from Scientific Reports:
https://www.nature.com/articles/s41598-020-62798-y (accessed 25 February 2022)
Chapter 4
Operational optimization and control
strategies for decarbonization in
WRRFs
Krishna R. Pagilla*
Department of Civil and Environmental Engineering, University of Nevada, Reno, NV, USA
*Correspondence: [email protected]
4.1 INTRODUCTION
Natural water resources are stressed due to population growth and reduced and/or uneven precipitation
induced by climate change impacts. Consequently, more extreme measures are taken to meet the
water demands including energy intensive water extraction, conveyance, and treatment systems.
Likewise, the wastewater generated requires energy to collect and treat to make it suitable for reuse
and environmental discharge. As water and other resources are extracted from wastewater for reuse,
there are costs in terms of energy usage and GHG emissions from these operations. No other part of the
water infrastructure is more energy intensive per unit of water handled than water reclamation from
wastewater, and it is the sector that has the highest potential to decarbonize or become carbon neutral/
positive. The wastewater itself is a potential source of energy for recovery. The embedded energy in
wastewater includes heat, organic and inorganic reduced species that constitute chemical energy, and
potential hydraulic energy in certain situations. The case for terming wastewater as a reNEWable
resource and wastewater treatment plants as water resource recovery facilities (WRRFs) stems from
the fact that nutrients (nitrogen and phosphorus), energy, and water can be recovered while reducing
the carbon footprint of the facility and its operations. After a WRRF has been built and commissioned,
the operational aspects and capacity utilization of the facility are mainly responsible for the carbon
footprint for the entire operational life. Hence, the focus of this chapter is to explore key opportunities
for decarbonization through optimization at process level and at whole facility level. The aspects
addressed in this chapter do not consider various process options/substitutions requiring capital
infrastructure changes for water reclamation and sludge treatment and management. The emphasis
is on how a conventional (solid–liquid separation, secondary treatment for carbon removal/recovery
and nitrification only) or advanced WRRF performing C, N, and P removal/recovery processes can
be optimized for decarbonization. Further strategies to reduce chemical use, capacity utilization, and
energy recovery by operational and process optimization strategies are discussed here.
With regard to energy demand level and GHG emissions, it is known that the carbon footprint is
nearly identical to the energy footprint in most WRRFs. Figure 4.1 shows an example of an approximately
38 000 m3/day activated sludge based conventional WRRF in the US that treats predominantly domestic
wastewater and discharges the treated effluent to the environment (Pagilla et al., 2009). It can be
© 2022 The Editors. This is an Open Access eBook distributed under the terms of the Creative Commons Attribution Licence
(CC BY-NC-ND 4.0), which permits copying and redistribution for non-commercial purposes with no derivatives, provided the
original work is properly cited (https://creativecommons.org/licenses/by-nc-nd/4.0/). This does not affect the rights licensed or
assigned from any third party in this book. The chapter is from the book Pathways to Water Sector Decarbonization, Carbon Capture
and Utilization, Zhiyong Jason Ren and Krishna Pagilla (Eds.)
GHG Emissions
Plant N2O = 160 tons CO2e/year
Digesters = 930 tons CO2e/year
Effluent N 2O = 270 tons
CO2e/year
Figure 4.1 Carbon footprint of a WRRF and the external site power for operating the processes (adapted from
Pagilla et al. 2009).
seen that the external site electrical power used in terms of carbon emissions equivalent is nearly
80% of the GHG emissions from the facility, while the rest comprises of emissions from the facility,
excluding biogenic CO2. The carbon footprint and hence the energy footprint can be minimized by
energy and chemical inputs to the treatment processes, and by maximizing energy generation from the
wastewater by carbon and heat capture. Figure 4.2(a) shows a simplistic rendition of the processes and
inputs/outputs in a WRRF and Figure 4.2(b) shows the key energy inputs and outputs in a WRRF. At
the fundamental level, the energy present in the wastewater and its extraction in usable form create
opportunities to reduce carbon footprint or decarbonize the WRRFs. The energy inputs for treatment
are costs that make-up the carbon footprints. Maximizing the opportunities and minimizing the inputs
through optimization is the key to decarbonization in the WRRF.
Figure 4.2 (a) Inputs and outputs of a conventional WRRF; (b) Energy inputs and outputs of a conventional WRRF.
these process/unit levels vary from facility to facility depending on the existing infrastructure or
treatment method employed. As facilities employ more advanced treatment methods to recover water
for reuse and nutrients from wastewater and sludge, the overall WRRF becomes energy intensive and
increases its carbon footprint due to more energy intensive processes employed. There can be offsets
due to the embedded carbon value of the recovered resources. The typical operations up to secondary
treatment and the corresponding energy needs are well established. Another way to represent energy
use in a WRRF is by function or use type for existing WRRFs with typical configuration. The typical
energy use distribution based on numerous data sources in the literature in an activated sludge
process-based WRRF with sludge treatment is shown in Figure 4.3. The typical standard deviation
value for each category is approximately 20% of the respective percent energy use. The main process
areas of concern include aeration for the activated sludge process, pumping of wastewater and sludge,
sludge treatment and processing, and others (buildings and lighting, odor control, disinfection, etc.).
It can be seen that the best opportunities for energy footprint reduction and decarbonization lie in
two or three main categories including aeration for process needs. At the process level, the energy
use and carbon footprint predominantly lie in wastewater pumping, secondary treatment, and sludge
treatment and are discussed further.
Figure 4.3 Typical energy use distribution in activated sludge process-based WRRFs.
• use of variable frequency drives (VFDs) and/or smaller pumps in pumping to adapt to changing
flow rates received by the facility;
• pump maintenance to keep the performance of the pumps at optimum levels;
• pumping control including flow management through in-line equalization and/or storage;
• data driven strategies to optimize pumping for economic and energy benefits.
Use of VFDs in influent pumping instead of fixed speed pumps is a common strategy to reduce
energy use due to variable influent flows over a day, week, and month/season/year. This is particularly
true in facilities that expect wide variations in influent flow due to significant inflow and infiltration
of stormwater. Although VFDs have the ability to reduce energy consumption, excessive pump speed
control without considering the pump characteristics can be counter-productive in terms of energy
use (Kato et al., 2019). A number of factors including static head, operational range of pump rotational
speed, and number of pumps in service influence the overall energy consumption (WEF, 2009). Hence,
power consumption analysis of plant specific pumping systems and optimizing for energy used per
unit discharge (specific energy) should be the overarching strategy to reduce energy/carbon footprint
of pumping. Over 30% improvement in energy efficiency was achieved in WRRFs employing such
analysis in plants using VFDs for influent pumping (Kato et al., 2019). Similar outcomes can be
expected for other functions such as recycles and sludge pumping.
Raw or screened wastewater pumping involves fluids that have high debris content and grit, thereby
creating rapid degradation of pump performance in terms of specific energy. Both pumps and pipeline
maintenance are critical to maintain pumping efficiency in WRRFs. A pump performance monitoring
strategy such as continuous pressure measurements, periodic cleaning, and maintenance through
pigging or short duration high-speed pumping are needed during ongoing operations to maintain high
efficiency (Larsen et al., 2016).
Since wastewater pumping systems are designed for a maximum flow rate received by the facility,
there is considerable deviation from peak performance (specific energy) of the pumping system at
average and minimum flow conditions. Although flow management to create wastewater storage or
in-line equalization in the sewer network is feasible and could be energy efficient for wastewater
pumping, the unintended consequences of pipeline degradation due to deposition of sediments,
fats/oils/greases (FOG), and odor issues need to be considered. Integrated control strategies which
simultaneously control both a WRRF and sewer system are necessary without structural changes
under dry weather and wet weather flow conditions (Kroll et al., 2018). This results not only in energy
savings, but also improves the effluent quality of WRRFs.
Overall, influent wastewater pumping has tremendous potential for reducing specific energy and
hence, decarbonization of WRRFs if pumping systems have the control and sensing equipment needed.
Among numerous examples, data driven strategies and control of wastewater pumping using fuzzy
logic, data-mining, and bench-marking with the best cases seem to have much potential to reduce
specific energy and improved energy savings (Torregrossa et al., 2017). These strategies need more
full-scale experience in multiple facilities before they become routine in the broader sector of water
reclamation and sludge treatment. Benchmarking with other facilities of similar size and capacity
utilization in terms of specific energy for pumping would reveal energy use reduction and hence
decarbonization potentials.
instead of oxidizing it, but could result in challenging conditions for subsequent denitrification and
also for sludge handling. The use of a so-called high rate activated sludge system (HRAS) preceding a
conventional activated sludge process or contact stabilization (CS) process has been found to be energy
efficient in secondary treatment (Rahman et al., 2019). Furthermore, adding chemically enhanced
primary treatment to the HRAS or CS process would recover 200% more carbon from wastewater.
As novel process technologies such as Anammox become common practice for nitrogen removal, the
potential for HRAS process with low MCRT (∼0.3 days) can yield good BOD removal and thereby
allow for energy optimized WRRFs (De Graff et al., 2016). Good design of the high rate process for
optimum sludge production and processing of sludge for energy production is a critical aspect of
successful carbon capture above conventional levels. Operational stability and reliability can then
reduce the carbon footprint of secondary treatment by carbon capture rather than carbon oxidation.
Aeration is the single largest source of energy use in most WRRFs and fine bubble diffused aeration
is the predominant type used in conventional activated sludge process-based WRRFs. Alternative
diffusers and control systems taking advantage of the newer developments in membranes and online
sensing of aeration systems have been developed to make aeration energy efficient, but diffuser fouling
and air distribution to optimize energy use are critical factors in on-going energy optimization during
operations of WRRFs. A combination of diffuser scaling and backpressure can significantly increase
energy use and reduce the performance of diffused aeration systems. Maintenance and upkeep of the
aeration systems including blowers, piping, diffusers, and online sensors will enable optimum air to
be delivered for process needs and maintain energy efficiency. However, the frequency of cleaning
needed is site specific and off-gas measurements to determine oxygen transfer efficiency are critical
for each plant based on its wastewater composition and process operations (Leu et al., 2009). The key
factors of interest with regard to aeration systems evaluation and maintenance for energy efficiency
have been widely discussed in the literature (Aviles et al., 2020; Drewnowski et al., 2019) and can be
summarized as follows:
• blower maintenance, sequencing, and optimization;
• minimizing pressure losses through headers and distributor piping;
• diffuser fouling and cleaning;
• maintenance of on-line dissolved oxygen, ammonia, total organic carbon sensors that control
air supply;
• replacement of non-functioning diffusers.
The application of real-time control of aeration systems to balance between air supply and demand
as a function of loadings and variations (diurnal, seasonal, annual, wet weather vs. dry weather, etc.)
and process conditions (temperature, treatment limits, etc.) could be very successful in achieving
aeration costs reduction and carbon footprint reduction in activated sludge process secondary
treatment. This approach has been in place for a few large facilities which were focused on operational
efforts to reduce energy costs during aeration for wastewater treatment (Leu et al., 2009). However,
successful full-scale case studies with real time monitoring and control of aeration systems and
demonstrated evidence of GHG emissions reductions are scant in the literature. The future potential
lies in integrating aeration systems real time monitoring and control strategy with plant process
control systems at plant scale for overall decarbonization.
Other potential opportunities to reduce aeration energy use in the activated sludge process include
operational densification of the activated sludge and low DO operations (Arnaldos & Pagilla, 2014;
Jassby et al., 2014). It is common knowledge that bulking due to excessive filamentous bacteria
presence in activated sludge makes it less dense and reduces the treatment capacity of the activated
sludge process per unit reactor volume. Jassby et al. (2014) showed that the higher the filamentous
bacteria content, the lower is the density of activated sludge, and hence lower is the settleability.
High filamentous bacteria in activated sludge in effect disallows high biomass concentrations in the
aeration tank because of poor settleability in the secondary clarifiers. Elimination of settling problems
in the activated sludge process through various proven methods including use of biomass selectors,
feeding strategies for optimum food-to-microorganisms ratio, aeration control, and chemical control
of bulking, are imperative to reduce energy use in the activated sludge process (Jenkins et al., 2003).
Arnaldos and Pagilla (2014) demonstrated that the activated sludge process can be operated at low
DO by acclimating the biomass over an extended period of time and, thereby reduce 20% aeration
energy use and 20% improvement in oxygen mass transfer efficiency. This strategy, although employed
in some full-scale facilities based on empirical reports, is not common practice in most full-scale
facilities.
Monitoring of alternative end points of aeration systems such as N species, pH, and ORP are
also possible in full-scale decarbonization. Combined monitoring of ORP, pH, and DO using sensors
and managing aeration have been demonstrated for a long period of time (Paul et al., 1998). The
strategy has been demonstrated for aeration control even under varying loading conditions in the
activated sludge process. Similarly, real-time control of biological nitrogen removal to achieve process
performance and aeration efficiency has also been successfully demonstrated (Zanetti et al., 2012).
In fact, aeration control of the activated sludge process using ammonia as the controlling variable
can not only provide energy use reduction, but also improve N removal and biological P removal
performance. The ammonia-based aeration control using a data-driven modeling approach was found
to be more effective than the DO-based control (Newhart et al., 2020). As with real-time DO control
for aeration, the N-based control systems in full-scale facilities are scant at the present. Therefore, real-
time monitoring and control using sensors and parameters combined with data-driven modeling of
full-scale facilities provides a great opportunity to find decarbonization strategies in aeration systems.
As can be seen from Figure 4.1, N2O emissions from biological processes during nitrification-
denitrification constitute a significant portion of the overall direct GHG emissions (scope 1) in WRRFs.
The facility shown in Figure 4.1, is a nitrification facility without a denitrification requirement. It has
been clearly demonstrated that N2O emissions occur from both nitrification and denitrification steps
in N removal (Rassamee et al., 2011). This is particularly the case when DO levels are controlled
in response to ammonia levels in wastewater causing either/and/or incomplete nitrification and
denitrification evidenced by high nitrite levels. In a survey of 12 WRRFs in the US, it was found
that there is a high degree of variability in N2O emissions from biological N removal plants due to
process and operational variations (Ahn et al., 2010). The N2O emission factors determined from
these facilities ranged over two orders of magnitude. Considering only biological N removal plants
with an estimated emission factor of 7.0 g N2O/PE/year translates to a flow-based emission factor
of approximately 51 mg/m3 with US average of 378 L wastewater/PE. Overall, it can be seen from
the example in Figure 4.1 that up to 10% of the facility’s carbon footprint is due to N2O emissions
in a nitrifying activated sludge plant. Therefore, the potential to reduce those emissions lies in the
secondary treatment due to the N2O production there. Efforts should focus on complete nitrification
and denitrification in BNR processes while careful attention should be paid to nitrite formation in
short cut N pathways processes which is the main factor in N2O production.
Figure 4.4 Conceptual diagram to identify GHG emissions reductions in sludge treatment.
anaerobic digestion, and chemicals used in thickening and dewatering. An additional operation that
is integrated with sludge treatment is energy generation and use (in the form of heat and electricity)
through biogas collection, conditioning or treatment. The recovered energy and heat serve as carbon
offsets in the overall sludge treatment system.
The key opportunities for decarbonization through sludge treatment in a WRRF are:
• improved operations to generate more biogas for energy recovery;
• reduction in fugitive emissions of methane and flaring of unused biogas;
• reduction of chemical use in sludge concentration steps.
Anaerobic digestion of wastewater sludge is the energy producing operation in a WRRF and
has significant potential to decarbonize a WRRF. It is possible to reduce the energy demand of
the anaerobic digestion process itself by optimizing sludge temperature and mixing, both of which
require energy inputs. Biogas production in anaerobic digestion is dependent on temperature, and
often higher operating temperatures have a positive effect in both mesophilic or thermophilic types.
Temperature optimization can lead to higher biogas production while balancing the energy needs for
sludge heating. Effects of temperature on mesophilic anaerobic digestion of sludge from 32 to 37.5°C
showed that there is no significant difference in biogas production when the temperature was reduced
from 37.5 to 35°C; however, a further decrease in digestion temperature led to a biogas production
decrease (Andersson et al., 2020). Therefore, a good temperature sensing system in feed sludge and
digester contents, and a feedback control operations strategy to maintain the optimum temperature
under varying ambient conditions and feed sludge temperature conditions, is critical to reduce energy
demand and increase biogas production.
Mixing is another important aspect of anaerobic digestion that not only influences energy use
but also process performance and operating issues. Uniform substrate conditions in the digester are
desired, but excessive mixing leads to foaming issues which can impact operations and hence biogas
production (Pagilla et al., 1997). It was later demonstrated that in most high-rate anaerobic digesters,
excessive mixing is highly likely and can cause foaming issues which impact digestion and biogas
production (Subramanian & Pagilla, 2015a; Subramanian et al., 2015b). The natural mixing due to
biogas production and sludge recirculation through the sludge heating loop are sufficient to maintain
homogeneity in the digester and process performance.
Table 4.1 Operational parameters for optimization of anaerobic digestion for biogas production in a WRRF.
Table 4.2 Chemicals used, purpose, and GHG emission factors in a 113 000 m3/day flow rate advanced water
reclamation facility in the US.
or fields due to residual incidental or residual methane production, and raw wastewater headworks
subjected to anaerobic conditions. The key sources of methane from anaerobic digestion include the
digested sludge, digester floating cover annular space, sludge buffer tanks, and leaks from gas handling
systems. Over 70% of total methane emissions from a WRRF are due to emissions from the anaerobic
digestion complex (Daelman et al., 2012). Therefore, operational strategies such as prevention of
biogas leaks, biogas collection from digested sludge storage and buffer tanks, and returning sludge
liquor recycles to the activated sludge process to oxidize dissolved methane are likely to reduce
fugitive emission of methane in a WRRF.
Small WRRFs with low biogas production or biogas production in excess of heating needs often tend
to flare the biogas for disposal. As a WRRF size decreases, the amount of biogas produced from AD that
is flared instead of utilized increases (Shen et al., 2015). The economics of biogas cleanup for utilization
has been cited as a major barrier in small WRRFs due to the low cost of natural gas in recent years.
Another operational strategy that can make significant progress in decarbonization of WRRFs
is the reduction in the use of chemicals for sludge treatment. The main chemicals used in sludge
treatment are polymers and inorganic coagulants for sludge concentration and dewatering. A case
study of chemical uses in a WRRF and the contribution of polymer use for sludge thickening and
dewatering is nearly 10% of the GHG emissions equivalent due to all chemicals used in the WRRF
(Table 4.2) and is further discussed in the following section.
The use of benchmarking to determine relative energy use per unit wastewater treated or
per capita, and other normalized metrics, is well discussed in the literature (Longo et al., 2016).
Benchmarking itself does not improve energy use efficiency in WRRFs, but reveals factors effecting
high or low energy use in a particular plant such as plant size, wastewater strength, flow rate
and its variations, capacity utilization, reclaimed water quality, and other regulatory and plant
requirements. Benchmarking also shows how a plant is performing relative to other WRRFs of
similar type based on key performance indicators (KPIs) (Longo et al., 2016) at the whole facility
level and at individual process/operation level. The energy analysis methods should include other
KPIs such as kWh/PE, kWh/kg COD removed, kWh/kg N removed, and so on, to show nuances in
specific energy in different facilities corresponding to different inputs and outputs and functions.
For example, using energy benchmarking including chemical use, Belloir et al. (2014) showed that
two facilities that have similar treatment processes had starkly different energy use per unit volume
of wastewater treated due to process and operational differences. Similar analysis can be carried
out for a plant of interest, and then target process level or plant level strategies for decarbonization
during operations.
When KPIs are estimated in terms of energy use in a WRRF, a key consideration of importance
is the operational capacity of the facility versus its design capacity. For example, two facilities with
similar treatment processes and the same designed capacity may have different specific energy use
because of variations in capacity utilization in terms of hydraulic and mass loading rates. This also
reflects in operational costs and GHG emissions per unit flow rate treated. Although larger facilities
have the ability to take parallel units out-of-service during low flow or load conditions, this is not
always practiced because other considerations. The key considerations include availability of labor
and operational ease with which the units can be taken out of service or put back into service. Any
facility that has a design capacity well in excess of the operating capacity should conduct an analysis
of operational strategies impacting specific energy use, operating costs, and GHG emissions. A data-
driven approach based on time-variant flows and loads information to analyze capacity utilization
and its impact on energy utilization in WRRFs is needed for operational decision making (Torregrossa
et al., 2019). A possible strategy to overcome capacity under-utilization are operating a base flow or
load facility with partial capacity utilization while keeping the rest of the capacity on standby mode
for transient flow/load treatment. This strategy requires that the facility has parallel trains which can
be easily isolated and kept in standby mode. Even small WRRFs which do not have parallel units
have high specific energy use due to under-utilization of the design capacity (Foladori et al., 2015),
suggesting that novel operational and control strategies are needed for them. Certain plants have the
ability to store the influent wastewater either in reservoirs or in the sewer system to manage short
term peak flows and loads.
Chemicals used in wastewater treatment and water reclamation contribute to the indirect emissions
(scope 2) of GHGs from WRRFs. Therefore, operational efforts and strategies to minimize chemical
use during facility operations are critical for overall decarbonization of the WRRF. The chemicals
used in WRRFs at the plant scale which are significant and used in most advanced water reclamation
facilities are disinfection chemicals, coagulants and flocculants, pH control, precipitation chemicals,
and carbon augmentation chemicals. Furthermore, if facilities employ more advanced treatment
processes for P recovery, softening, biogas recovery, and water reuse, additional types or quantities of
multipurpose chemicals and additives (hydrogen peroxide, ozone, iron materials) may also be used in
a WRRF. The relative carbon footprint or indirect emissions per unit quantity is a function of supply
chain and source of the chemicals in a specific WRRF. A most recent case study to determine the
comprehensive carbon footprint of WRRFs in the Baltic Sea region is a good example of estimating
decarbonization potentials in WRRFs (Maktabifard et al., 2022). Table 4.2 shows typical chemicals
and estimated indirect emissions in a 113 000 m3/day operating flow, advanced water reclamation
facility in Reno, NV, USA. The facility includes advanced treatment for N and P control and chlorine-
based disinfection (Lacroix et al., 2020).
Table 4.3 Variability in energy footprint due to capacity utilization, level of treatment, and chemicals used in five
BNR facilities in the US.
The overall indirect emissions (scope 2) due to chemical use in this facility is about 0.24 kg CO2e/
m3 of wastewater treated. The largest contribution to the carbon footprint of this facility is the use of
methanol for tertiary denitrification which is equivalent to 0.14 kg CO2e/m3 of wastewater treated.
Therefore, the largest decarbonization potential in this facility can be realized by either finding
internal carbon sources to replace methanol or optimization of N removal in the overall facility to
minimize methanol use for denitrification. A comparison of full-scale data collected showed that
biological nutrient removal facilities that practice chemical addition such as alum and methanol
for denitrification and P removal can double the energy/carbon footprint of wastewater treatment
compared to the facilities that do not. Table 4.3 shows the role of capacity utilization and chemical use
on the energy footprint of wastewater per unit volume of wastewater treated.
A case study which investigated an alternative sludge treatment processing strategy which led to
whole plant chemical use reduction and operating cost reduction was described by Mentzer et al.
(2021). The major goal of this study was to enhance the dewaterability of final sludge being sent to a
landfill for disposal. The facility investigated bypass of thickened WAS from anaerobic digestion and
combined dewatering of digested primary sludge and un-digested thickened WAS. This operational
strategy not only reduced the overall polymer use in sludge dewatering, but either eliminated or
considerably reduced the use of other chemicals such as Mg(OH)2 (for struvite recovery), sulfuric acid
for pH adjustment, and methanol due to lower N in the recycles, and others. In fact, the outcomes were
dramatic in terms of enhanced anaerobic digestion capacity, eliminating the need to treat dewatering
centrate, and reduced operational problems such as struvite scaling, gas conditioning media fouling,
and so on.
sustainability goals in mind including potentials for decarbonization is critical. At the same time, the
early wins to set the path for future decarbonization pursuits lie in current operations. Major areas of
operations which have maximum potentials for decarbonization in WRRFs and their associated sewer
systems are as follows:
(a) Minimize dilution of wastewater and increase the energy density of wastewater through reduced
water use and inflow/infiltration into sewers.
(b) Source control of trace pollutants that have low concentration impact thresholds and are
energy intensive to treat through advanced treatment methods. These include pharmaceuticals,
forever chemicals, and anthropogenic nanomaterials.
(c) Source control of salinity to minimize energy intensive extraction of reclaimed water for reuse.
(d) Effective flow management to minimize pumping and hydraulic overloads to the facility.
(e) Carbon capture and P recovery instead of energy intensive and chemical intensive biological
and chemical processes.
(f) Anaerobic treatment systems with complete methane capture for carbon recovery.
(g) Efficient aeration systems (if cannot be replaced with anaerobic treatment options) that
supply oxygen to meet actual metabolic demands of the process instead of open tank aeration
providing less than 20% oxygen transfer efficiency.
(h) Densification of WRRF operations through process optimization and technology selection.
(i) Online sensing and feedback/feed-forward control of processes to achieve treatment goals,
energy conservation, and minimize emissions.
(j) Minimize nitrous oxide and methane emissions through operational control of processes.
(k) Minimize and eliminate chemicals use, particularly those with high life cycle GHG emissions
in WRRF operations. Use of external carbon sources such as methanol and coagulants/
flocculants/polymers present immediate decarbonization potential in existing WRRFs.
(l) Novel in-plant modifications to concentrate and dewater sludge better for further processing or
reuse.
(m) Enhance biogas production from anaerobic digestion through operational and process
modifications and optimizations.
(n) Optimizing of auxiliary facilities such as odor control, gas cleaning, and in-plant transportation.
Although the above is an extensive list of potentials and opportunities, they can be prioritized
and addressed for the overall decarbonization of the WRRFs by careful carbon footprinting
of the facility and its operations. The future lies in more WRRFs conducting operational carbon
footprint determinations at least at scope 1 level so that more full-scale data can be developed to
carry out benchmarking of WRRFs with similar process trains, capacity utilization, resource quality
requirements, and other variables of interest. This allows the water sector to develop KPIs that can be
compared across various WRRFs to set goals for decarbonization in each facility. The heterogeneity
of the carbon footprints in drinking water facilities and WRRFs is large and hence, best practices
from more full-scale facilities can be implemented at others with higher carbon footprint KPIs. For
example, the GHG emissions from energy consumption by drinking water and wastewater treatment
facilities in the US were in the range of <0.1–0.8 and <0.1–0.65 kg CO2e/m3, respectively (Zib III
et al., 2021). Such aggregate values do not reflect ground reality of the heterogeneity among facilities.
The ability to embark on large-scale water sector decarbonization is dependent on the availability of
this full-scale data and practical strategies to decarbonize at process and whole plant level.
The broader and significant opportunity and positive impact to decarbonize in WRRFs and the
water sector as a whole in existing facilities is not possible without the education and training of
facility personnel and staff to understand and implement feasible decarbonization strategies. The goal
should be to clearly demonstrate the operational costs savings, enhanced treatment performance and
efficiency, and how they can contribute to addressing the climate change effects by decarbonization
in their respective facility.
REFERENCES
Ahn J. H., Kim S., Park H., Katehis D., Pagilla K. and Chandran K. (2010). Spatial and temporal variability in
atmospheric nitrous oxide generation and emission from full-scale biological nitrogen removal and non-BNR
processes. Water Environment Research, 82(12), 2362–2372.
Andersson J., Helander-Claesson J. and Olsson J. (2020). Study of reduced process temperature for energy
optimization in mesophilic digestion: a lab to full-scale study. Applied Energy, 271, 115108. https://doi.
org/10.1016/j.apenergy.2020.115108
Arnaldos M. and Pagilla K. R. (2014). Implementation of a demand-side approach to reduce aeration requirements
of activated sludge systems: directed acclimation of biomass and its effect at the process level. Water
Research, 62, 147–155. https://doi.org/10.1016/j.watres.2014.05.040
Aviles A. B. L., Velazquez F. D. C. and del Riquelme M. L. P. (2020). Integral control system for the aeration stage
in the biological process of activated sludge and the membrane biological reactor. Sensors, 20, 4342. https://
doi.org/10.3390/s20154342
Belloir C., Stanford C. and Soares A. (2014). Energy benchmarking in wastewater treatment plants: the importance
of site operation and layout. Environmental Technology, 36, 260–269. https://doi.org/10.1080/09593330.20
14.951403
Daelman M. R. J., van Voorthuizen E. M., van Dongen U. G. J. M., Volcke E. I. P. and van Loosdrecht M. C. M.
(2012). Methane emission during municipal wastewater treatment. Water Research, 46, 3657–3670. https://
doi.org/10.1016/j.watres.2012.04.024
Dalmau J., Rodriguez-Roda I., Comas Q., Steyer J. P. and Pagilla K. R. (2010). Model development and simulation
for predicting risk of foaming in anaerobic digestion systems. Bioresource Technology, 101, 4306–4315.
https://doi.org/10.1016/j.biortech.2010.01.056
De Graff M. S., van den Brand T. P. H., Roest K., Zandvoort M. H., Duin O. and van Loosdrecht M. C. M.
(2016). Full-scale highly-loaded wastewater treatment processes (A-stage) to increase energy production
from wastewater: performance and design guidelines, Environmental Engineering Science, 33, 571–577.
https://doi.org/10.1089/ees.2016.0022
Drewnowski J., Remiszewska-Skwarek A., Duda S. and Lagod G. (2019). Aeration process in bioreactors as
the main energy consumer in a wastewater treatment plant – review of solutions and methods of process
optimization. Processes, 7, 311. https://doi.org/10.3390/pr7050311
Flores-Alsina X., Corominas L., Snip L. and Vanrolleghem P. A. (2011). Including greenhouse gas emissions
during benchmarking of wastewater treatment plant control strategies. Water Research, 45, 4700–4710.
https://doi.org/10.1016/j.watres.2011.04.040
Jassby D., Xiao Y. and Schuler A. J. (2014). Biomass density and filament length synergistically affect activated
sludge settling: systematic quantification and modeling. Water Research, 48, 457–465. https://doi.
org/10.1016/j.watres.2013.10.003
Jenkins D., Richard M. G. and Daigger G. T. (2003). Manual on the Causes and Control of Activated Sludge
Bulking, Foaming, and Other Solids Separation Problems, 3rd edn, Taylor and Francis, New York, NY.
Kato H., Fujimoto H. and Yamashina K. (2019). Operational improvement of main pumps for energy-saving in
wastewater treatment plants. Water, 11, 2438–2450. https://doi.org/10.3390/w11122438
Kroll S., Fenu A., Wambecq T., Weemases M., Impe J. V. and Willems P. (2018). Energy optimization of the
urban drainage system by integrated real-time control during wet and dry weather conditions. Urban Water
Journal, 15, 362–370. https://doi.org/10.1080/1573062X.2018.1480726
Lacroix A., Mentzer C. and Pagilla K. R. (2020). Full-scale N removal from centrate using a sidestream process
with a mainstream carbon source. Water Environment Research, 92, 1922–1934. https://doi.org/10.1002/
wer.1345
Larsen T., Arensman M. and Nerup-Jensen O. (2016). Including pressure measurements in supervision of energy
efficiency of wastewater pump systems. Journal of Hydraulic Engineering, 142(2), 04015048.
Leu S., Rosso D., Larson L. E. and Stenstrom M. (2009). Real-time aeration efficiency monitoring in the activated
sludge process and methods to reduce energy consumption and operating costs. Water Environment
Research, 81, 2471–2481. https://doi.org/10.2175/106143009X425906
Longo S., d’Antoni B. M., Bongards M., Chaparro A., Cronrath A., Fatone F., Lema J., Mauricio-Iglesias M., Soares
A. and Hospido A. (2016). Monitoring and diagnosis of energy consumption in wastewater treatment plant –
A state of the art and proposals for improvement. Applied Energy, 179, 1251–1268. https://doi.org/10.1016/j.
apenergy.2016.07.043
Lu L., Guest J., Peters C. A., Zhu X., Rau G. H. and Ren Z. J. (2018). Wastewater treatment for carbon capture and
utilization. Nature Sustainability, 1, 750–758. https://doi.org/10.1038/s41893-018-0187-9
Maktabifard M., Awaitey A., Merta E., Haimi H., Zaborowska E., Mikola A. and Makinia J. (2022). Comprehensive
evaluation of the carbon footprint components of wastewater treatment plants located in the Baltic Sea
region. STOTEN, 806, 150436.
Mannina G., Ekama G., Caniani D., Cosenza A., Esposito G., Gori R., Garrido-Baserba M., Rosso D. and Olsson
G. (2016). Greenhouse gases from wastewater treatment – a review of modelling tools. STOTEN, 551–552,
254–270.
Mentzer C., Drinkwater M. and Pagilla K. R. (2021). Investigation of direct waste activated sludge dewatering
benefits and costs in a water resource recovery facility. Water Environment Research, 93, 1–13. https://doi.
org/10.1002/wer.1651
Metcalf & Eddy, Inc. (2014). In: 2014 Wastewater Engineering: Treatment and Resource Recovery, G. Tchobanoglous,
H. D. Stensel, R. Rsuchihashi and F. Burton (eds), 5th edn, McGraw Hill, New York, NY, USA.
Newhart K. B., Marks C. A., Rauch-Williams T., Cath T. Y. and Hering A. (2020). Hybrid statistical-machine
learning ammonia forecasting in continuous activated sludge treatment for improved process control.
Journal of Water Process Engineering, 37, 101389. https://doi.org/10.1016/j.jwpe.2020.101389
Pagilla K. R., Craney K. C. and Kido W. H. (1997). Causes and effects of foaming in anaerobic sludge digesters.
Water Science and Technology, 36, 463–470. https://doi.org/10.2166/wst.1997.0624
Pagilla K., Shaw A. R., Kunetz T. and Schiltz M. (2009). A Systematic Approach to Establishing Carbon Footprints
for Wastewater Treatment Plants. Proceedings of WEFTEC. Water Environment Federation, Alexandria,
VA, USA.
Paul E., Plisson-Saune S., Mauret M. and Cantet J. (1998). Process state evaluation of alternating oxic-anoxic
activated sludge using ORP, pH, and DO. Water Science and Technology, 38, 299–306. https://doi.
org/10.2166/wst.1998.0224
Rahman A., De Clippeleir H., Thomas W., Jimenez J., Wett B., Al-Omari A., Murthy S., Riffat R. and Bott C. (2019).
A stage and high rate contact-stabilization performance comparison for carbon and nutrient redirection from
high strength municipal wastewater. Chemical Engineering Journal, 357, 737–749. https://doi.org/10.1016/j.
cej.2018.09.206
Rajagopal R., Masse D. I. and Singh G. (2013). A critical review of inhibition of anaerobic digestion process by
excess ammonia. Bioresource Technology, 143, 632–641. https://doi.org/10.1016/j.biortech.2013.06.030
Rassamee V., Sattayatewa C., Pagilla K. and Chandran K. (2011). Effect of oxic and anoxic conditions on nitrous oxide
emissions from nitrification and denitrification processes. Biotechnology Bioengineering, 108(9), 2036–2045.
Shen Y., Linville J. L., Urgun-Demirtas M., Mintz M. M. and Snyder S. W. (2015). An overview of biogas production
and utilization at full-scale wastewater treatment plants (WWTPs) in the United States: Challenges and
opportunities towards energy-neutral WWTPs. Renewable & Sustainable Energy Review, 50, 346–362.
Subramanian B. and Pagilla K. R. (2015a). Mechanisms of foam formation in anaerobic digesters. Colloids and
Surfaces B: Biointerfaces, 126, 621–630. https://doi.org/10.1016/j.colsurfb.2014.11.032
Subramanian B., Miot A., Jones B., Klibert C. and Pagilla K. R. (2015b). A full scale study of mixing and foaming in
egg-shaped digesters, Bioresource Technology, 192, 461–470. https://doi.org/10.1016/j.biortech.2015.06.023
Torregrossa D., Hansen H., Hernandez-Sancho F., Cornelissen A., Schutz G. and Leopold U. (2017). A data-
driven methodology to support pump performance analysis and energy efficiency optimization in waste
water treatment plants. Applied Energy, 208, 1430–1440. https://doi.org/10.1016/j.apenergy.2017.09.012
Torregrossa D., Castellet-Viciano L. and Hernandez-Sancho F. (2019). A data analysis approach to evaluate the
impact of the capacity utilization on the energy consumption of wastewater treatment plants. Sustainable
Cities and Society, 45, 307–313. https://doi.org/10.1016/j.scs.2018.11.036
WEF (2009). Energy Conservation in Water and Wastewater Facilities. WEF Manual of Practice, Alexandria, VA,
USA.
Wu D., Peng X., Li L., Yang P., Peng Y., Liu H. and Wang X. (2021). Commercial biogas plants: review on
operational parameters and guide for performance optimization. Fuel, 303, 121282. https://doi.org/10.1016/j.
fuel.2021.121282
Zanetti L., Frison N., Nota E., Tomizioli M., Bolzonella D. and Fatone F. (2012). Progress in real-time control
applied to biological nitrogen removal from wastewater – a short-review. Desalination, 286, 1–7. https://doi.
org/10.1016/j.desal.2011.11.056
Zib III L., Byrne D. M., Marston L. T. and Chini C. M. (2021). Operational carbon footprint of the U.S. water
and wastewater sector’s energy consumption. Journal of Cleaner Production, 321, 128815. https://doi.
org/10.1016/j.jclepro.2021.128815
Chapter 5
Energy and resource recovery using
the anaerobic digestion platform
Prathap Parameswaran1*, Jessica A. Deaver2, Sudeep C. Popat2, Vikas Khanna3,
Madison Kratzer3 and Mel Harclerode4
1Kansas State University, Manhattan, KS 66506, USA
2Clemson University, Clemson, SC, USA
3University of Pittsburgh, Pittsburgh, PA, USA
*Correspondence: [email protected]
© 2022 The Editors. This is an Open Access eBook distributed under the terms of the Creative Commons Attribution Licence
(CC BY-NC-ND 4.0), which permits copying and redistribution for non-commercial purposes with no derivatives, provided the
original work is properly cited (https://creativecommons.org/licenses/by-nc-nd/4.0/). This does not affect the rights licensed or
assigned from any third party in this book. The chapter is from the book Pathways to Water Sector Decarbonization, Carbon Capture
and Utilization, Zhiyong Jason Ren and Krishna Pagilla (Eds.)
Digested
solids
Figure 5.1 Energy balance analysis based on actual data collected from a WRRF and adapted from Shizas and Bagley
(2004). The anaerobic digestion step is the main source of internal energy recovery from the sludge in a WRRF, as
shown in these figures. It is to be noted that the 17% of the energy requirement is for both mixing and heating
requirements of the digester, which can be optimized further to enhance the net recovery and sequestration of the
valuable energy and carbon. (a) Primary sedimentation basin – energy balance. (b) Aeration basin – energy balance.
(c) Anaerobicdigester – energy balance.
this list is limited to municipal WRRFs in the United States which accept external waste from
outside the facility to perform anaerobic co-digestion, which puts them on a positive trajectory
to achieve energy neutrality and subsequent beneficial recovery of nutrients and end use of the
produced biogas.
The objective of this chapter is to review and summarize current knowledge about sludge
management and provide a basis for future practices. The chapter focuses on how current
management practices help decarbonization, the role of sludge management strategies in achieving
the decarbonization targets of utilities, and how challenges (e.g., emerging contaminants, odors,
public scrutiny and upset) in sludge management can be addressed in meeting such targets. It should
be noted that residual digestate after AD of sludge also needs additional treatment and disposal.
New emerging concepts, namely Water-Energy Nexus, Circular Economy, and Nutrient Trading, are
important vehicles for decarbonization in shaping the future sludge management practices. These
concepts significantly help to reduce the financial burdens of sludge management on societies and
overcome ecological issues and resource scarcity. New technologies and approaches need to be
developed to extract energy and nutrients from sludge and improve process and energy efficiency.
Recovered energy and nutrients help utilities become a source of revenue generation, overcoming
their reputation as pollution mitigation entities. In turn, they will become entities contributing to
reduction in carbon emissions and achieve decarbonization of the water sector. Renewable energy
production and resource recovery are presented as areas of sludge management to close the linearity
of waste production and implement a circular economy of waste management. The chapter closes
with a section on implementation of decarbonization at the utilities and future strategies and
pathways.
East Bay Municipal Utility District 238 Energy positive operation Yes, for EBPR through No
(EBMUD), San Francisco, USA sidestream fermentation*
DC Water, USA 1450 10 MW (*67% energy Yes, for EBPR through N & P for silviculture
neutrality) sidestream fermentation*
Des Moines Metro Wastewater Reclamation 227 14 000 MWh/year NA Biosolids for land application
Authority, IA
Buffalo Sewer Authority, Buffalo, NY 465 Yes NA Biosolids for land application
Hyperion Treatment Plant, Los Angeles, CA 1135 Yes NA Class A biosolids for fertilizer
Sacramento Regional County Sanitation 567 Yes NA Biosolids for land application
District, Sacramento, CA
Stickney WRRF, Stickney, IL 2840 Yes Primary effluent Biosolids; struvite P recovery with
fermentation for EBPR Ostara®
Southeast Water Pollution Control Plant, San 212 Yes NA Biosolids for land application
Joint Water Pollution Control Plant (JWPCP), 1135 Yes Biosolids for land application or
Carson, CA composting
Point Loma Wastewater treatment plant, San 681 Yes6.4 MW through
Diego, CA cogeneration
San José-Santa Clara Regional Wastewater 416 Yes NA NA
Facility, CA
Central Wastewater Treatment Plant, 378 Yes NA Biosolids for land application
Nashville, TN
69
This information was screened from the database presented in www.resourcerecoverydata.org and was subject to further screening criteria such as published information available about
nutrient products sequestration. Detailed information foused mainly on energy self sufficiency for WRRFs with AD in the US, Canada, and Europe can be found in Shen et al. (2015).
70 Pathways to Water Sector Decarbonization, Carbon Capture and Utilization
Figure 5.2 A summary of key sludge pretreatment mechanisms under each of which there are patented and/or
commercial platforms available and being used in municipal, industrial, and agricultural wastewater installations
worldwide. The key findings summarized here are from Kim et al. (2003), Khanal et al. (2007), Rittmann et al. (2008)
and Burger and Parker (2013).
of successful co-digestion. Krupp et al. (2005) investigated the ecological and economic impacts of
sludge co-digestion with OFMSW in oversized, full-scale digesters. Life cycle assessments revealed
that compared to composting and mono-substrate digestion, ACoD was more beneficial in the climate
change category, and when applied at a larger plant its economic benefit was best. Pavan et al. (2007)
and Righi et al. (2013) also demonstrated under specific conditions at smaller, full-scale treatment
plants in rural areas, that co-digestion can be a beneficial tool for improving the economic balances of
WWTP. Important considerations include size of the digesters, volume of co-substrates for treatment,
and reduced transportation and storage time.
5.5 ALTERING THE AD PLATFORM FOR HIGHER ORGANIC CARBON PRODUCT CAPTURE
COUPLED WITH WATER REUSE AND NUTRIENT RECOVERY
Strong drivers such as global climate change, changing economic landscape, and increased demand
for chemicals and sustainably derived plastics could soon transform carbon capture from wastewater
treatment facilities to the modern wastewater resource recovery facility. Short- and medium-chain
carboxylic acids are an essential intermediate in the anaerobic food web that leads to methanogenesis,
which when controlled can accumulate to high concentrations in the bioreactor. Several strategies have
been proposed to manage arrested methanogenesis in an anaerobic digestion platform to facilitate the
hydrolysis and acidogenesis products to accumulate and be recovered through subsequent separation
techniques. Recent research has indicated that a bio-electrochemically assisted anaerobic digester
may not only serve to enhance the overall hydrolysis and fermentation rates, but also promote greater
accumulation of the higher organic acids by selective consumption of acetate by the electroactive
bacteria through thermodynamic and kinetic benefits. The H 2 rich environment in the AnMBR should
not only arrest methanogenesis but also facilitate secondary fermentation reactions leading to higher
order VFA synthesis (Bhatt et al., 2020; De Vrieze et al., 2018; Jiang et al., 2018).
Separation of VFAs from fermented broths is challenging due to low VFA concentrations in ion-rich
solutions. Consequently, separation capacity and selectivity with traditional solvents and adsorbents
are both compromised. In the recent extraction literature, ionic liquids (ILs) have been reported for
extraction of VFAs, and some have been noticed to be better than conventional solvents in terms of
extraction efficiency. ILs exist as molten salts at ambient temperature and consist entirely of ions,
usually a charge-stabilized organic cation and an inorganic or organic anion. A study concluded that
phosphonium-based ILs are better extractants than the traditional organic solvents for recovery of
short chain organic acids from aqueous dilute solutions. They succeeded to obtain higher distribution
coefficients as compared to most traditional solvents by using phosphonium-based ionic liquids for the
extraction of low concentrated lactic acid solutions, and obtained an extraction efficiency of 98.4%
with a two-step extraction (Liang et al., 2017; Oliveira et al., 2012). IL-mediated esterification for
reactive extraction of low-value VFA from dilute aqueous streams was also reported. Distillation
or evaporation will be needed as the final VFA purification step. Membrane based non-reactive or
reactive separations from integrated anaerobic digesters is an equally effective separation platform
that has been receiving increased attention (Zhu et al., 2021).
Figure 5.3 Net energy requirement comparison for: (a) conventional activated sludge; (b) AnMBR platform for methane
and nutrient capture.
decreasing fouling energy requirements even further by periodic pulse sparging at high flow rates
rather than continuous sparging; mixing energy optimization in the primary bioreactor. Anaerobic
digester energy optimizations will likely be focused on decreasing pretreatment costs, decreasing
mixing energy requirements, enhancing process based sludge dewaterability, and enhancing the
overall energy capture efficiency from the produced methane rich biogas.
5.7 TECHNO-ECONOMIC AND LIFE CYCLE ASSESSMENTS FOR SHAPING THE FUTURE
OF ANAEROBIC DIGESTION
Traditionally there has been a disconnect between actual experimental data and Life Cycle Assessment
(LCA) studies for emerging technologies, with most studies focusing on either aspect discretely, and
only a few studies attempt to bridge this gap. Green engineering principles are incorporated late in
the design/concept development process, resulting in incremental environmental improvement rather
than process pathways that minimize life cycle environmental impacts. Integrated techno-economic
and LCA platforms have the ability to proactively guide conceptual designs that are environmentally
and economically conscious and aimed at maximizing bioenergy capture and carbon recovery in a
higher value form such as carboxylates along with other valuable products like nutrient products
and water for indirect/direct potable reuse. The use of the AnMBR process for domestic wastewater
treatment presents an opportunity to mitigate environmental, social, and economic impacts currently
incurred from energy-intensive conventional aerobic activated sludge processes. A pilot-scale study
of the AnMBR and concurrrent Techno Economic Analysis (TEA) and LCA were performed to
demonstrate and validate AnMBR technology for more sustainable domestic wastewater treatment
compared to aerobic activated sludge.
The feasibility of the proposed AnMBR platform is supported through our preliminary LCA of
the AnMBR platform. Figure 5.4(a) shows the environmental impacts of various AnMBR treatment
scenarios compared to conventional treatment. The LCA indicated that the most sustainable system
was conventional treatment followed by scenario 7, a hybrid AnMBR with vacuum flash tank methane
recovery and sulfide removal operating at 15 LMH (liters of treated water per unit membrane surface
area per hour) and temperature higher than 25°C. Previous studies have performed detailed evaluations
on improving AnMBR process subcomponents to maximize energy recovery and dissolved methane
recovery. Few studies have broadly evaluated the role of chemical use, membrane fouling management,
and dissolved methane removal technologies. Figure 5.4(b) shows the potential for implementing an
AnMBR that achieves an overall environmental impact less than conventional treatment by reducing,
if not eliminating, chemical removal of sulfide and phosphorous. The global warming potential (kg CO2
equivalents) for this AnMBR configuration without chemical coagulation for chemical removal is shown
as an offset resulting from methane recovery for bioenergy use. The Sankey diagram of scenario 7 AnMBR
(Figure 5.5) shows the nutrient removal component as the major contributor to environmental impacts.
The feasibility of our studied AnMBR system is also corroborated by preliminary TEA. Preliminary
analysis show that the levelized operational cost of the hybrid AnMBR with methane and sulfide
removal without the volatile fatty acid (VFA) separation is $ 0.09/m3. On a full-scale treatment plant
with the capacity to treat 22 730 m3/day, the annual operation costs are $ 781 300 respectively.
Systems-level optimization has the potential to further reduce the operational costs for the hybrid
AnMBR design because of process improvements and associated reductions in chemical and energy
use. When compared on a construction cost basis, the AnMBR design has a higher construction cost
($ 71 582 500) compared to a conventional activated sludge design ($ 59 991 250). However, it is to be
noted that the revenue associated with value-added products that could potentially be recovered in
modified AnMBR is currently not accounted for and is expected to bring down the operational costs
significantly. Similar scenarios can be developed for the sidestream AD platform as well, although the
impacts are projected to be less impactful when compared to direct anaerobic wastewater treatment,
although there is promising potential.
Figure 5.4 Relative impacts of the AnMBR and conventional treatment scenarios at 15 LMH (F2) and >25°C (T1) with
(a) and without (b) sulfide and phosphorus removal by chemical coagulation using ferric chloride and ACH. Impacts
in b are relative to those for Scenarios 1 and 3 in a. Source of tables is Harclerode et al. (2020).
Conclusion of the TEA and LCA study determined two process subcomponents, sulfide and
phosphorus removal and sludge management, that drove chemical use and residuals generation, and
in turn the environmental and cost impacts. Furthermore, integrating primary sedimentation and a
vacuum degassing tank for dissolved methane removal maximized net energy recovery. Sulfide was
generated by anaerobic reduction of naturally occurring sulfate. Previous studies have not considered
the cost and environmental impact of sulfide generation and removal via chemical coagulation. The
TEA/LCA demonstrated that the AnMBR can have a lower environmental impact and operating cost
relative to conventional wastewater treatment if sulfide can be removed biologically rather than with
chemical coagulation.
Figure 5.5 Sankey diagrams of the global warming impact (kg CO2-equivalent/m3) assessment for scenario 7, a
hybrid AnMBR with vacuum flash tank methane recovery and sulfide removal operating at 15 LMH and temperature
higher than 25°C. Plots from Harclerode et al. (2020).
Even though the environmental impacts of chemicals in our modeled process are significant, we
expect these impacts can be reduced significantly based on process optimization studies and when
the process is implemented at scale. As such, the final integrated AnMBR platform design has the
potential to have much lower environmental impact compared to the conventional process while also
recovering value-added products.
Sound techno-economic and life cycle analyses will need to be intimately coupled to current and
future AD configurations to maximize not only carbon capture, but overall resource recovery in a
holistic fashion.
ACKNOWLEDGEMENTS
The authors would like to acknowledge help from graduate PhD student in Environmental Engineering,
Mr. Arvind Damodara Kannan, for his assistance with formatting the chapter and references. The
authors would also like to specifically acknowledge time and effort from graduate and undergraduate
students from Dr. Parameswaran’s team: Ms. Priyasha Fernando; Ms. Emily Randig; Ms. Marleigh
Hutchinson; and Mr. Mason Ericson for their help with compiling Table 5.1 from a very large US
dataset. Direct help for Figure 5.1 was obtained from Dr. Cesar Torres, Associate Professor, Chemical
Engineering at Arizona State University.
REFERENCES
Agler M. T., Spirito C. M., Usack J. G., Werner J. J. and Angenent L. T. (2012). Chain elongation with reactor
microbiomes: upgrading dilute ethanol to medium-chain carboxylates. Energy & Environmental Science,
5(8), 8189–8192. https://doi.org/10.1039/c2ee22101b
Alves H. J., Bley C., Niklevicz R. R., Frigo E. P., Frigo M. S. and Coimbra-Araujo C. H. (2013). Overview of
hydrogen production technologies from biogas and the applications in fuel cells. International Journal of
Hydrogen Energy, 38(13), 5215–5225. https://doi.org/10.1016/j.ijhydene.2013.02.057
Amha Y. M., Anwar M. Z., Brower A., Jacobsen C. S., Stadler L. B., Webster T. M. and Smith A. L. (2017). Inhibition
of anaerobic digestion processes: applications of molecular tools. Bioresource Technology, 247, 999–1014.
Augelletti R., Conti M. and Annesini M. C. (2017). Pressure swing adsorption for biogas upgrading. A new process
configuration for the separation of biomethane and carbon dioxide. Journal of Cleaner Production, 140,
1390–1398. https://doi.org/10.1016/j.jclepro.2016.10.013
Batstone D. J. and Virdis B. (2014). The role of anaerobic digestion in the emerging energy economy. Current
Opinion in Biotechnology, 27, 142–149. https://doi.org/10.1016/j.copbio.2014.01.013
Batstone D. J., Hulsen T., Mehta C. M. and Keller J. (2015). Platforms for energy and nutrient recovery from
domestic wastewater: a review. Chemosphere, 140, 2–11. https://doi.org/10.1016/j.chemosphere.2014.10.021
Battistoni P., Paci B., Fatone F. and Pavan P. (2006). Phosphorus removal from anaerobic supernatants: start-up
and steady-state conditions of a fluidized bed reactor full-scale plant. Industrial & Engineering Chemistry
Research, 45(2), 663–669. https://doi.org/10.1021/ie050796g
Bhatt A. H., Ren Z. and Tao L. (2020). Value proposition of untapped wet wastes: carboxylic acid production
through anaerobic digestion. Iscience, 23(6), 101221. https://doi.org/10.1016/j.isci.2020.101221
Burger G. and Parker W. (2013). Investigation of the impacts of thermal pretreatment on waste activated sludge
and development of a pretreatment model. Water Research, 47, 5245–5256. https://doi.org/10.1016/j.watres.
2013.0 6.005
Carlsson M., Lagerkvist A. and Morgan-Sagastume F. (2012). The effects of substrate pretreatment on anaerobic
digestion systems: a review. Waste Management, 32, 1634–1650. https://doi.org/10.1016/j.wasman.2012.04.016
Carrère H., Dumas C., Battimelli A., Batstone D. J., Delgenès J. P., Steyer J. P. and Ferrer I. (2010). Pretreatment
methods to improve sludge anaerobic degradability: a review. Journal of Hazardous Materials, 183, 1–15.
https://doi.org/10.1016/j.jhazmat.2010.06.129
Cecchi F., Traverso P. G., Mata-Alvarez J., Clancy J. and Zaror C. (1988). State of the art of R&D in the anaerobic
digestion process of municipal solid waste in Europe. Biomass, 16(4), 257–284. https://doi.org/10.1016/
0144-4565(88)90031-5
Chauzy J., Cretenot D., Bausseron A. and Gokelaere X. (2007). Thermal Hydrolysis to Increase Sludge
Biodegradability or How to Turn Mesophilic Anaerobic Digestion of Biological Sludge Into an Attractive
Process. WEFTEC 2007, 80th Annual Water Environment Federation Technical Exhibition and Conference,
13–17 October, San Diego, CA.
De Vrieze J., Arends J. B. A., Verbeeck K., Gildemyn S. and Rabaey K. (2018). Interfacing anaerobic digestion
with (bio)electrochemical systems: potentials and challenges. Water Research, 146, 244–255. https://doi.
org/10.1016/j.watres.2018.08.045
Dewil R., Appels L. and Baeyens J. (2006). Energy use of biogas hampered by the presence of siloxanes. Energy
Conversion and Management, 47(13–14), 1711–1722. https://doi.org/10.1016/j.enconman.2005.10.016
Eskicioglu C., Kennedy K. J. and Droste R. L. (2006). Characterization of soluble organic matter of waste activated
sludge before and after thermal pretreatment. Water Research, 40(20), 3725–3736. https://doi.org/10.1016/j.
watres.2006.08.017
Gonzalez A., Hendriks A. T. W. M., van Lier J. B. and de Kreuk M. (2018). Pretreatments to enhance the biodegradability
of waste activated sludge: elucidating the rate limiting step. Biotechnology Advances, 36, 1434–1469. https://doi.
org/10.1016/j.biotechadv.2018.06.001
Grosser A. and Neczaj E. (2016). Enhancement of biogas production from sewage sludge by addition of grease trap
sludge. Energy Conversion and Management, 125, 301–308. https://doi.org/10.1016/j.enconman.2016.05.089
Hagos K., Zong J., Li D., Liu C. and Lu X. (2017). Anaerobic co-digestion process for biogas production: progress,
challenges and perspectives. Renewable and Sustainable Energy Reviews, 76(March 2016), 1485–1496.
https://doi.org/10.1016/j.rser.2016.11.184
Harclerode M., Doody A., Brower A., Vila P., Ho J. and Evans P. J. (2020). Life cycle assessment and economic analysis
of anaerobic membrane bioreactor whole-plant configurations for resource recovery from domestic wastewater.
Journal of Environmental Management, 269, 110720. https://doi.org/10.1016/j.jenvman.2020.110720
Haug R. T., Stuckey D. C., Gossett J. M. and Mccarty P. L. (1978). Effect of thermal pretreatment on digestibility
and dewaterability of organic sludges. Journal of Water Pollution Control Federation, 50(1), 73–85.
Jiang Y., Lu L., Wang H., Shen R. X., Ge Z., Hou D. X., Chen X., Liang P., Huang X. and Ren Z. J. (2018).
Electrochemical control of redox potential arrests methanogenesis and regulates products in mixed culture
electro-fermentation. ACS Sustainable Chemistry & Engineering, 6(7), 8650–8658. https://doi.org/10.1021/
acssuschemeng.8b00948
Khanal S. K., Grewell D., Sung S. and Van Leeuwen J. (2007). Ultrasound applications in wastewater sludge
pretreatment: a review. Critical Reviews in Environmental Science and Technology, 37, 277–313. https://doi.
org/10.1080/10643380600860249
Ki D., Parameswaran P., Popat S. C., Rittmann B. E. and Torres C. I. (2015). Effects of pre-fermentation and
pulsed-electric-field treatment of primary sludge in microbial electrochemical cells. Bioresource Technology,
195, 83–88. https://doi.org/10.1016/j.biortech.2015.06.128
Kim J., Park C., Kim T. H., Lee M., Kim S., Kim S. W. and Lee J. (2003). Effects of various pretreatments for
enhanced anaerobic digestion with waste activated sludge. Journal of Bioscience and Bioengineering, 95(3),
271–275. https://doi.org/10.1016/S1389-1723(03)80028-2
Kim T.-H., Kim T.-H., Yu S., Nam Y. K., Choi D.-K., Lee S. R. and Lee M.-J. (2007). Solubilization of waste activated
sludge with alkaline treatment and gamma ray irradiation. Journal of Industrial and Engineering Chemistry,
13(7), 1149–1153.
Kim D., Lee K. and Park K. Y. (2015). Enhancement of biogas production from anaerobic digestion of waste
activated sludge by hydrothermal pretreatment. International Biodeterioration & Biodegradation, 101, 42–
46. https://doi.org/10.1016/j.ibiod.2015.03.025
Krupp M., Schubert J. and Widmann R. (2005). Feasibility study for co-digestion of sewage sludge with OFMSW on
two wastewater treatment plants in Germany. Waste Management, 25(4), 393–399. https://doi.org/10.1016/j.
wasman.2005.02.009
Lanzini A. and Leone P. (2010). Experimental investigation of direct internal reforming of biogas in solid oxide fuel cells.
International Journal of Hydrogen Energy, 35(6), 2463–2476. https://doi.org/10.1016/j.ijhydene.2009.12.146
Lee I. and Rittmann B. E. (2016). Using focused pulsed technology to remove siloxane from municipal sewage
sludge. Journal of Environmental Engineering, 142(1). https://doi.org/10.1061/(asce)ee.1943-7870.0000975
Lee I. S., Parameswaran P., Alder J. M. and Rittmann B. E. (2010). Feasibility of focused-pulsed treated waste
activated sludge as a supplemental electron donor for denitrification. Water Environment Research, 82(12),
2316–2324. https://doi.org/10.2175/106143010X12609736967288
Li H., Jin Y., Mahar R., Wang Z. and Nie Y. 2008 Effects and model of alkaline waste activated sludge treatment.
Bioresource Technology, 99(11), 5140–5144. https://doi.org/10.1016/j.biortech.2007.09.019
Liang L., Li C., Xu F., He Q., Yan J., Luong T. and Sun N. (2017). Conversion of cellulose rich municipal solid waste
blends using ionic liquids: feedstock convertibility and process scale-up. RSC Advances, 7(58), 36585–36593.
https://doi.org/10.1039/C7RA06701A
Lim K., Evans P. J. and Parameswaran P. (2019). Long-term performance of a pilot-scale gas-sparged anaerobic
membrane bioreactor under ambient temperatures for holistic wastewater treatment. Environmental Science
and Technology, 53(13), 7347–7354. https://doi.org/10.1021/acs.est.8b06198
Long J. H., Aziz T. N., Reyes F. L. D. L. and Ducoste J. J. (2012). Anaerobic co-digestion of fat, oil, and grease
(FOG): a review of gas production and process limitations. Process Safety and Environmental Protection,
90(3), 231–245. https://doi.org/10.1016/j.psep.2011.10.001
Mata-Alvarez J., Mace S. and Llabres P. (2000). Anaerobic digestion of organic solid wastes. An overview of
research achievements and perspectives. Bioresource Technology, 74, 3–16. https://doi.org/10.1016/S0960-
8524(00)00023-7
Mata-Alvarez J., Dosta J., Romero-Güiza M. S., Fonoll X., Peces M. and Astals S. (2014). A critical review on
anaerobic co-digestion achievements between 2010 and 2013. Renewable and Sustainable Energy Reviews,
36, 412–427. https://doi.org/10.1016/j.rser.2014.04.039
McCarty P. L., Bae J. and Kim J. 2011 Domestic wastewater treatment as a net energy producer – can this be
achieved? Environmental Science and Technology, 45(17), 7100–7106. https://doi.org/10.1021/es2014264
Nickel K. and Neis U. (2007). Ultrasonic disintegration of biosolids for improved biodegradation. Ultrasonics
Sonochemistry, 14(4), 450–455. https://doi.org/10.1016/j.ultsonch.2006.10.012
Oliveira F. S., Araújo J. M., Ferreira R., Rebelo L. P. N. and Marrucho I. M. (2012). Extraction of L-lactic, L-malic,
and succinic acids using phosphonium-based ionic liquids. Separation and Purification Technology, 85,
137–146. https://doi.org/10.1016/j.seppur.2011.10.002
Pavan P., Bolzonella D., Battistoni E. and Cecchi F. (2007). Anaerobic co-digestion of sludge with other organic
wastes in small wastewater treatment plants: an economic considerations evaluation. Water Science and
Technology, 56(10), 45–53. https://doi.org/10.2166/wst.2007.730
Pavlostathis S. G. and Giraldo-Gomez E. (1991). Kinetics of anaerobic treatment. Water Science and Technology,
24(8), 35–59. https://doi.org/10.2166/wst.1991.0217
Pickworth B., Cranshaw I., Abraham K., Coleman P., Walley P. and Solheim O. E. (2005). Large Scale Reality of
Sewage Sludge Pasteurisation and Thermal Hydrolysis. WEFTEC 2005, 78th Annual Water Environment
Federation Technical Exhibition and Conference, 29 October–2 November, Washington, DC.
Popat S. C. and Deshusses M. A. (2008). Biological removal of siloxanes from landfill and digester gases: opportunities
and challenges. Environmental Science and Technology, 42(22), 8510–8515. https://doi.org/10.1021/es801320w
Puyol D., Batstone D. J., Hulsen T., Astals S., Peces M. and Kromer J. O. (2016). Resource recovery from wastewater
by biological technologies: opportunities, challenges, and prospects. Frontiers in Microbiology, 7, 2106.
Righi S., Oliviero L., Pedrini M., Buscaroli A. and Della Casa C. (2013). Life cycle assessment of management
systems for sewage sludge and food waste: centralized and decentralized approaches. Journal of Cleaner
Production, 44(2013), 8–17. https://doi.org/10.1016/j.jclepro.2012.12.004
Rittmann B. E., Lee H. S., Zhang H. S., Alder J., Banaszak J. E. and Lopez R. (2008a). Full-scale application of
focused-pulsed pretreatment for improving biosolids digestion and conversion to methane. Water Science
and Technology, 58(10), 1895–1901. https://doi.org/10.2166/wst.2008.547
Salerno M. B., Lee H. S., Parameswaran P. and Rittmann B. E. (2009). Using a pulsed electric field as a pretreatment
for improved biosolids digestion and methanogenesis. Water Environment Research, 81(8), 831–839. https://
doi.org/10.2175/106143009X407366
Scarlat N., Dallemand J. F. and Fahl F. (2018). Biogas: developments and perspectives in Europe. Renewable
Energy, 129, 457–472. https://doi.org/10.1016/j.renene.2018.03.006
Shen Y. W., Linville J. L., Urgun-Demirtas M., Mintz M. M. and Snyder S. W. (2015). An overview of biogas
production and utilization at full-scale wastewater treatment plants (WWTPs) in the United States:
challenges and opportunities towards energy-neutral WWTPs. Renewable and Sustainable Energy Reviews,
50, 346–362. https://doi.org/10.1016/j.rser.2015.04.129
Shizas I. and Bagley D. M. (2004). Experimental determination of energy content of unknown organics in
municipal wastewater streams. Journal of Environmental Engineering, 130(2), 45. https://doi.org/10.1061/
(ASCE)0733-9402(2004)130:2(45)
Steinbusch K. J. J., Hamelers H. V. M., Plugge C. M. and Buisman C. J. N. (2011). Biological formation of caproate
and caprylate from acetate: fuel and chemical production from low grade biomass. Energy and Environmental
Science, 4(1), 216–224. https://doi.org/10.1039/C0EE00282H
Tandukar M. and Pavlostathis S. G. (2015). Co-digestion of municipal sludge and external organic wastes for
enhanced biogas production under realistic plant constraints. Water Research, 87, 432–445. https://doi.
org/10.1016/j.watres.2015.04.031
Tyagi V. K., Fdez-Güelfo L. A., Zhou Y., Álvarez-Gallego C. J., Garcia L. I. R. and Ng W. J. (2018). Anaerobic
co-digestion of organic fraction of municipal solid waste (OFMSW): progress and challenges. Renewable
and Sustainable Energy Reviews, 93(April), 380–399. https://doi.org/10.1016/j.rser.2018.05.051
Vlyssides A. G. and Karlis P. K. (2004). Thermal-alkaline solubilization of waste activated sludge as a
pretreatment stage for anaerobic digestion. Bioresource Technology, 91(2), 201–206. https://doi.org/10.1016/
S0960-8524(03)00176-7
Wolff H. J., Nickel K., Houy A., Lunden A. and Neis U. (2007). Two Years Experience on a Large German STP
with Acoustic Disintegration of Waste Activated Sludge for Improved Anaerobic Digestion. 11th IWA World
Congress on Anaerobic Digestion, Session PP9C-Biosolids, 23–27 September, Brisbane, Australia.
Xie S., Higgins M. J., Bustamante H., Galway B. and Nghiem L. D. (2018). Current status and perspectives on
anaerobic co-digestion and associated downstream processes. Environmental Science: Water Research and
Technology, 4(11), 1759–1770. https://doi.org/10.1039/C8EW00356D
Yang Q., Wu B., Yao F., He L., Chen F., Ma Y., Shu X., Hou K., Wang D. and Li X. (2019). Biogas production
from anaerobic co-digestion of waste activated sludge: co-substrates and influencing parameters. Reviews
in Environmental Science and Biotechnology, 18(4), 771–793. https://doi.org/10.1007/s11157-019-09515-y
Zhang D., Feng Y. M., Huang H. B., Khunjar W. and Wang Z. W. (2020). Recalcitrant dissolved organic nitrogen
formation in thermal hydrolysis pretreatment of municipal sludge. Environment International, 138, 105629.
https://doi.org/10.1016/j.envint.2020.105629
Zhu X., Leininger A., Jassby D., Tsesmetzis N. and Ren Z. J. (2021). Will membranes break barriers on volatile fatty
acid recovery from anaerobic digestion? Environmental Science and Technology, 1(1), 141–153.
Chapter 6
Carbon valorization using the
microbial electrochemical
technology platform
Jayesh M. Sonawane1, Zhiyong Jason Ren2 and Deepak Pant2*
1Department of Chemistry, Alexandre-Vachon Pavilion, Laval University, 1045, avenue de la medicine, local 4064E, Quebec, Canada
G1V 0A6
2Department of Civil and Environmental Engineering & Andlinger Center for Energy and the Environment, Princeton University,
*Correspondence: [email protected]
6.1 INTRODUCTION
The world’s economic growth is currently dependent upon resources obtained from non-renewable
resources such as fossil fuels. These resources generate excessive greenhouse gases resulting in global
climate change, leading to floods and droughts, rising sea levels, and more frequent natural disasters
(Bhatia et al., 2019; Lu et al., 2018). The global CO2 emission has increased to approximately 35.5
gigatons (Gt), and the capture and utilization of CO2 have been expensive (MacDowell et al., 2017).
In the meantime, large amounts of organic carbon waste like municipal solid waste (2.01 Gt) and
wastewater (around 1000 km3) have also been a major environmental challenge with the high cost
of disposal and treatment (Kaza et al., 2018; Unesco, World Water Assessment Programme, 2012).
Alternately, these carbon-abundant waste materials (solid, liquid, and gaseous) can be reused diligently.
In that case, value-added energy and products can be generated to increase the value proposition of
the process and transform the waste valorization industry. Recently, the US Department of Energy
Bioenergy Technologies Office (BETO) reported that the US generates 50 million dry tons of organic
waste streams from food waste, manure, oils, fats, greases, and wastewater sludge. Combining the
carbon-containing gaseous waste streams, a total 2.6 quadrillion Btu of renewable energy can be
recovered (DOE 2017). Different technologies, including biochemical, photochemical, electrochemical,
and thermochemical processes, have been developed for carbon valorization, but they all have some
advantages and specific challenges (DOE 2017).
Microbial electrochemical technology (MET) is a platform technology in which electroactive
microorganisms are used to catalyze bioelectrochemical reactions to generate energy and products
from waste carbon materials (Wang & Ren, 2013; Zou & He, 2018). In this process, oxidation and
reduction reactions are separated in suitable environmental conditions for the first time. The main asset
of the process is that the electrodes can be used as either electron acceptor (anode) or electron donor
(cathode) (Jiang & Zeng, 2019; Pandey et al., 2016). Compared to other environmental technologies
© 2022 The Editors. This is an Open Access eBook distributed under the terms of the Creative Commons Attribution Licence
(CC BY-NC-ND 4.0), which permits copying and redistribution for non-commercial purposes with no derivatives, provided the
original work is properly cited (https://creativecommons.org/licenses/by-nc-nd/4.0/). This does not affect the rights licensed or
assigned from any third party in this book. The chapter is from the book Pathways to Water Sector Decarbonization, Carbon Capture
and Utilization, Zhiyong Jason Ren and Krishna Pagilla (Eds.)
with only one or two functions, the MET platform is very flexible and has discovered dozens of
functions over the years. Almost all MET reactors share one common principle in the anode, in which
biodegradable substrates, such as wastewater and food waste, are oxidized by microorganisms and
generate electrical current (Wang & Ren, 2013). The current can be captured directly for electricity
generation (microbial fuel cells, MFCs) or used to produce H 2 and other value-added chemicals
(microbial electrolysis cells, MECs) (Logan, 2008). In addition, such electrons from organic waste
carbon can also be used in the cathode chamber to reduce CO2 and generate organic or inorganic
compounds, achieving double benefits of carbon capture and valorization. Microbial electrosynthesis
(MES) and microbial electrolytic carbon capture (MECC) are two popular processes in MET that
can directly convert cathodic CO2 and anodic organic waste into products (Lu et al., 2015; Rabaey &
Rozendal, 2010), while electrofermentation (EF) is another MET process that uses electro potential to
regulate fermentation processes for different products (Nevin et al., 2010).
Figure 6.1 Schematic representation of the principle of the microbial ecosystem (MES) and electro-fermentation
(EF): (a) The reduction of CO2 by MES. (b) In anodic EF, the working electrode (anode) accepts electrons from
microbes. (c) In cathodic EF, the working electrode (cathode) provides electrons to microbes. Figure reprinted from
a previous study with permission from Jiang et al. (2019).
EF can be divided into anodic EF and cathodic EF. In an anodic EF, the electrode is an electron
acceptor receiving electrons from microbial substrate oxidization. With an EF cathode, the working
electrode is an electron donor, which microbes can use to synthesize different products depending on
the redox potential. The electron transfer between the electrodes and microbes can be bidirectional.
Still, compared with MES, syntrophic interactions between fermentative bacteria and electroactive
bacteria were found to dominate and played essential roles in EF reactors, mainly due to the need for
degradation of complex organics present in waste materials (Choi & Sang, 2016).
Figure 6.2 Schematic representation of the principle of microbial electrolytic carbon capture (MECC) systems:
(a) An integrated system with in situ CO2 capture and mineralization. (b) A system that separates acid and alkali
generations and reactions for carbonization (Lu et al. 2015; Zhu & Logan, 2014).
Figure 6.3 Historical evolution of acetic and butyric acid production by MES from CO2: (a) Maximum concentrations
achieved in catholyte. (b) production rates with respect to cathode projected surface area. Green circles (acetic
acid), red circles (butyric acid) (Prévoteau et al. 2020).
A potential application of MES is the combination with anaerobic digestion (AD), because MES
can enhance the overall organic removal while in the meantime purifying the biogas generated by
AD by converting CO2 into CH4 and therefore also increasing the overall CH4 production yield and
efficiency. Studies showed that the gas generated with such a combination had a less than 10% CO2
content. Figure 6.4 describes an assortment of food waste exhibited by solid lines and a wastewater
treatment plant demonstrated by dashed lines along with the employment of a joined AD-MES unit.
The rejected water in the wastewater plant contains a COD concentration ranging from 1000 to
8000 mg L −1. The rejected water was recycled to reduce total solids in the inlet feed in the plants for
treating food waste. This AD-MES plant could reduce ammonium, sulfide, and COD concentrations
from the rejected water. In addition, CO2 was reduced to CH4 electrochemically by using optimal
cathode potentials and pH. This was done by decreasing the pH of the rejected mass, which enabled
the dissolution of the CO2 present in the biogas. Also, MES had to be designed so that the dissolved
CO2 could efficiently react with the electrons liberated from the surface of the cathode (Nelabhotla &
Dinamarca, 2019).
Acetic acid is another popular product from MES del Pilar Anzola Rojas et al. (2021). Acetic acid
can be produced electrochemically under biologically relevant conditions at −0.28 V (Equation (6.4)),
which is slightly lower than the CH4 generation potential (−0.24 V) (Rabaey & Rozendal, 2010):
2HCO−3 + 9H+ + 8e− → CH 3COO− + 4H2O (6.4)
Figure 6.4 AD-MES integrated treatment plant setup (figure reprinted from Nelabhotla & Dinamarca (2019)).
Table 6.1 Conversion of CO2 to acetic acid using different microbial strains.
The production rate of acetic acid (around 685 gm−2 day−1) from CO2 using MES technology was recently
achieved using a newly manufactured electrode and an adaptable microbial culture (Jourdin et al., 2015).
A 3D electrophoretic deposition electrode was used as a biocathode, and a multiwalled carbon nanotubes
layer was deposited onto reticulated vitreous carbon. A slightly acidic pH (∼5.8) increased the rate of the
formation of acetate, while the effect of current was carried out separately (Batlle-Vilanova et al., 2015).
However, after the biofilm’s growth of a particular thickness, a decrease in bacterial growth was observed.
The conversion efficiencies of acetate from carbon dioxide and electrons were steady and superior for
a mixed culture system, with an average of 98 ± 4 and 100 ± 1%, respectively. A high production rate of
the compounds also depends on other factors like the hydraulic retention time. Studies also showed the
production rate could be enhanced by increasing the cell voltage or altering different parameters like
membrane and electrode resistance, concentrations, pH, and variation in the anode potential (Blanchet
et al., 2015). Some found there was a loss of biomass in the continuous mode attributed to the participation
of both suspended (planktonic) and biofilm bacteria in the reduction process of CO2. Moreover, the loss
of planktonic bacteria also reduced the production in the reactor operating in continuous mode, which
was solved by substituting biofilm-forming microorganisms in place of planktonic bacteria. CO2 was
made available to acetogenic bacteria by continuously passing CO2 through the culture medium or using
bicarbonate as feed. In terms of inoculum, many studies used anaerobic sludge, and some observed a
prominent shift toward specific microbial families such as Clostridiaceae and Pseudomonadaceae on
the electrodes’ surface (Saratale et al., 2017). Other dominant microbes reported in literature included
Sulfurospirillum, Sporomusa, Clostridium, Tissierella, Arcobacter, Ochrobactrum, Pseudomonas,
Sacharolyticum, and Desulfovibrio (Zaybak et al., 2013). Table 6.1 summarizes typical CO2 to acetic acid
MES parameters and identified microbial cultures.
many studies reported that in situ generated H 2 was the direct electron donor for CO2 reduction with a
higher organic conversion rate, though the exact mechanisms were influenced by the cathode working
potential (Blanchet et al., 2015). For example, one study evaluated the function of hydrogen in electron
transfer on the methanogenic biofilm formed on the MES cathode. By using a microsensor to detect
the in-situ hydrogen generation in conjunction with cyclic voltammetry, Cai et al. (2020) analyzed
the hydrogen evolution dynamic and confirmed the presence of hydrogen-associated electron transfer
near the cathode within a micrometer scale, and they observed colocalized community of archaea
and bacteria developed within a 58.10-µm-thick biofilm was correlated with the hydrogen gradient
detected by the microsensor.
Hydrogen is increasingly produced via electrolysis powered by low-cost green electricity generated
from MET or other renewable sources, and hydrogen serves as a reductant for different chemotropic
microbes (Jack et al., 2021). These microbes consist of reversible hydrogenase enzymes that oxidizes
molecular hydrogen for CO2 reduction to organic compounds such as methane, acetic acid, or
butyric acid (Ganigué et al., 2015). Besides electrochemically produced hydrogen at the electrodes,
hydrogen can also be produced microbially via fermentation, especially in reactors working with
mixed microbial communities. Such a source of hydrogen allows the MES to survive periods when no
electric power is supplied to the system, thus making intermittent operation possible (del Pilar Anzola
Rojas et al., 2018).
Hydrogen-mediated electron transfer has been identified as an important extracellular pathway
of sharing reducing equivalents to regulate biofilm activities in MESs and demonstrated higher
reactor performance than the direct electron transfer pathway. Direct electron transfer could
only provide low current density, but electrocatalytic hydrogen production is tunable. Hydrogen
supply from the electrode can be increased with the increase in current densities, but a high rate
of hydrogen production may not necessarily lead to high CO2 conversion by microbes due to their
slow metabolism rates compared to the abiotic hydrogen evolution. Therefore, a balance needs to
be maintained between hydrogen supply and consumption in MES reactors. Studies reported that
biofilm could be eliminated from the electrodes in conditions where vigorous hydrogen evolution
occurs at the electrode. Also, though the alkaline condition is desired for the water electrolysis,
in MES cathode, the pH needs to be maintained near neutral for biological reactions. As a result,
a hybrid MES system can be fabricated by attaching a microbial gas-liquid contactor towards the
downstream of the water electrolysis cell.
Figure 6.5 Preliminary analysis of the CO2 valorization potential by MES in terms of economics and carbon
utilization. For each circle, the Y-axis value determines the location of the center, while the X-axis value determines
the radius. The radius range has no meaning for Y-axis. (a) The product value of different compounds generated
in MES reactors, which was calculated by multiplying the unit price with the maximum production rate. (b) The
world market size of each produced compound versus their unit price, respectively. (c) CO2 conversion potential via
different products, which was calculated by multiplying the unit conversion ratio with the maximum production rate
of each compound. (d) The world market size of CO2 conversion based on the production of each compound in MES
(figure reprinted from Jiang et al. (2019).
for overall metabolic activity. As a result, if the redox potential is artificially tuned, the fermentation
pathways can be regulated to generate different ratios of products (Jiang et al., 2018).
Figure 6.6 Schematic diagram of an anaerobic digestion/fermentation system incorporated with electrodes to
enable electro-fermentation (figure reprinted from Zhao et al. (2020)).
unchanged (Gajaraj et al., 2017). It was hypothesized that the reduction of CO2 to CH4 becomes
a major pathway of methanogenesis, in which the electrons are supplied by electroactive bacteria
such as Geobacter degrading organic acids or collecting electrons from the cathodes. Moreover, IHT
can also be enhanced due to the increase in hydrogen generation on the cathodes and subsequent
utilization by hydrogenotrophic methanogens (Villano et al., 2017). For example, Liu et al. (2019)
found that by introducing carbon brush electrodes into anaerobic digestion, VFA concentration drops
faster than regular AD control, indicating an accelerated stabilization. Moreover, methane production
was increased by 26.3% when a low voltage (0.8 V) was applied, and the content of methane in the
headspace also increased by nearly 30%. Community analysis showed the electric current stimulated
the growth of hydrogenotrophic methanogens, and Geobacter occurred at the cathode with a low
abundance. However, acetotrophic Methanosaeta still made up a high portion of the archaeal
community.
While AD produces renewable biogas, it faces challenges on economic viability and environmental
concerns due to the low value of biogas and concerns about its greenhouse gas effects. Recent
developments on arrested methanogenesis allow the AD process to be rewired to suppress
methanogenesis and promote the production of short chain VFAs and alcohols, because such products
not only bring up the values by themselves, they are also chemical precursors for the production
of higher-valued chemicals such as PHBs, biofuels, medium chain fatty acids (MCFAs), and single
cell protein (SCP) (Zhu et al., 2021). By controlling the redox potential in the EF reactor, the
fermentation pathways can be influenced and subsequently regulate the product spectrum. Recent
studies confirmed that the electrochemical potential control regulated the product distribution
in anaerobic fermentation using natural microbial consortia. Jiang et al. (2019) characterized the
product spectrum under different working potentials of −1.0, −0.6, and −0.2 V (vs. Ag/AgCl), which
spans the electron flow direction from cathodic current to anodic current. It was found when a
working potential of −0.2 V was applied; the electrode potential was more positive than the open
circuit potential (−0.55 V); therefore, anodic electro-fermentation reactions occurred with electrons
flowing toward the electrode. In contrast, when the applied potentials were more negative than the
open circuit potential (−0.6 and −1.0 V), cathodic EF conditions were created, where the working
electrode became an electron source. Results showed that more negative potential led to higher CH4
accumulation, while more positive potential showed inhibited methanogenesis activity. For example,
increasing the potential from −1.0 to −0.2 V greatly reduced methanogenesis by 68% and acetic acid
generation by 58% in neutral pH. Butyric acid production increased by 25%, while propionic acid
concentrations remained stable. Laboratory studies showed a range of 61–78% in carbon recovery
and 70–87% in electron recovery could be obtained by comparing final fermentation products and the
substrate, and the spectrum of each product was tuned based on the difference of working potential.
Since EF is regarded as an electrochemically-influenced spontaneous fermentation, the contribution
of electron consumption or donation by the working electrode to the electron balances was limited,
and energy consumption was low.
Figure 6.7 (a) schematic of the mutual benefits between an MECC equipped water resource recovery facility (WRRF)
and a CO2 point emission source. (b) example co-locations of CO2 point sources and WRRFs enable complementary
CCU in Beijing, Shanghai, New York city, and Denver (Lu et al., 2015, 2018).
many of these facilities are co-located with or near major WRRFs. The wastewater facility can help
capture and sequester the CO2 emitted from the nearby point source and generate carbon credits,
and the emitter can save costs by avoiding the use of expensive and energy-intensive CCS systems yet
still meet EPA mandates on carbon pollution reduction. The calcium/magnesium-abundant fly ash
generated by the power plant may be used as a silicate supplement to facilitate CO2 mineralization. For
WRRFs that do not have a nearby point source, MECC can help capture the CO2 from either aerobic
or anaerobic treatment processes. Even air capture is feasible when the process is combined with a
pre-concentration process by commercially available ion-exchange resins (Huang et al., 2016).
Studies also performed preliminary economic analysis on MECC systems and found that the net
cost for mitigating one ton of CO2 could be $48/ton (Lu et al., 2015), which was calculated based on
a combination of CO2 capture cost (capital plus operation cost), potential cost offsets (revenue of H2
and wastewater treatment) and avoided CO2 emission through reduced fossil fuel consumption for
wastewater treatment and commercial H 2 production (nature gas reforming). This net cost is well
below the $70–270/t- CO2 estimated for coal power plant with geological storage CCS, and also below
the cost for direct air CO2 capture using chemical/thermal methods (on the order of $1000/t- CO2) or
abiotic electrolytic dissolution of silicate ($86/t- CO2) (Cornils, 2020; House et al., 2011; Rau et al.,
2013). It should be noted that though MECC has potential for significant energy savings and carbon
benefits for the wastewater industry, further work is needed to better understand the technology
barriers and to optimize system designs, operational protocols, and applications.
6.6 OUTLOOK
Microbial electrochemical technology provides a versatile platform for simultaneous waste treatment,
resource recovery, and CO2 capture and utilization. While this technology has yet been demonstrated
in full scale, this chapter aims to summarizing different processes and opportunities for water and
wastewater treatment. We recognize it is challenging to meet and balance multiple objectives and
fulfill different treatment needs while in the meantime capturing CO2 and recovering resources, so this
chapter offers several feasible approaches to make such an operation a reality. Rather than building
new and separated systems, current reactors such as aeration tanks and anaerobic digestors can be
upgraded by installing electrodes to accomplish multi-function and process intensification (ElMekawy
et al., 2016). MECC, MES, and EF are based on different mechanisms and generate different products,
so they can be retrofitted into different systems with tailored purposes. For example, EF can be used
to enhance biogas production, MES can be used to generate VFAs, while MECC will help improve the
alkalinity of the wastewater.
It is possible to realize the carbon-negative, revenue-positive wastewater treatment. Still,
technological development and implementation as well as more detailed techno-economic, life cycle,
and socioeconomic analyses, are required to understand the potential of these technologies. Chapter
1 presented a hypothetical example of MECC plus microalgae to replace traditional anaerobic-anoxic-
aerobic activated sludge system, and it shows such conversion can potentially transform wastewater
treatment to carbon capture and valorization facilities with a positive revenue flow (Lu et al., 2018).
REFERENCES
Bajracharya S., Ter Heijne A., Dominguez Benetton X., Vanbroekhoven K., Buisman C. J. N., Strik D. P. B. T. B.
and Pant D. (2015). Carbon dioxide reduction by mixed and pure cultures in microbial electrosynthesis using
an assembly of graphite felt and stainless steel as a cathode. Bioresource Technology, 195, 14–24. https://doi.
org/10.1016/j.biortech.2015.05.081
Bajracharya S., Vanbroekhoven K., Buisman C. J. N., Pant D. and Strik D. P. B. T. B. (2016). Application of
gas diffusion biocathode in microbial electrosynthesis from carbon dioxide. Environmental Science and
Pollution Research, 23, 22292–22308. https://doi.org/10.1007/s11356-016-7196-x
Bajracharya S., Vanbroekhoven K. and Buisman C. J. N. (2017). Bioelectrochemical conversion of CO2 to
chemicals: CO2 as a next generation feedstock for electricity-driven bioproduction in batch and continuous
modes. Faraday Discussions, 202, 433–449. https://doi.org/10.1039/C7FD00050B
Batlle-Vilanova P., Puig S., Gonzalez-Olmos R., Vilajeliu-Pons A., Balaguer M. D. and Colprim J. (2015).
Deciphering the electron transfer mechanisms for biogas upgrading to biomethane within a mixed culture
biocathode. RSC Advances, 5, 52243–52251. https://doi.org/10.1039/c5ra09039c
Bhatia S. K., Bhatia R. K., Jeon J. M., Kumar G. and Yang Y. H. (2019). Carbon dioxide capture and bioenergy
production using biological system – a review. Renewable and Sustainable Energy Reviews, 110, 143–158.
https://doi.org/10.1016/j.rser.2019.04.070
Bian B., Bajracharya S., Xu J., Pant D. and Saikaly P. E. (2020). Microbial electrosynthesis from CO2: challenges,
opportunities and perspectives in the context of circular bioeconomy. Bioresource Technology, 302, 122863.
https://doi.org/10.1016/j.biortech.2020.122863
Blanchet E. M., Duquenne F., Rafrafi Y., Etcheverry L., Erable B. and Bergel A. (2015). Importance of the hydrogen
route in up-scaling electrosynthesis for microbial CO2 reduction. Energy and Environmental Science, 8,
3731–3744. https://doi.org/10.1039/c5ee03088a
Blankenship R. E., Tiede D. M., Barber J., Brudvig G. W., Fleming G., Ghirardi M., Gunner M. R., Junge W.,
Kramer D. M., Melis A., Moore T. A., Moser C. C., Nocera D. G., Nozik A. J., Ort D. R., Parson W. W., Prince
R. C. and Sayre R. T. (2011). Comparing photosynthetic and photovoltaic efficiencies and recognizing the
potential for improvement. Science, 332, 805–809. https://doi.org/10.1126/science.1200165
Chandrasekhar K., Amulya K. and Venkata Mohan S. (2014). Solid phase bio-electrofermentation of food waste
to harvest value-added products associated with waste remediation. Waste Management, 45, 57–65. https://
doi.org/10.1016/j.wasman.2015.06.001
Cai W., Liu W., Wang B., Yao H., Guadie A. and Wang A. (2020). Semiquantitative detection of hydrogen-
associated or hydrogen-free electron transfer within methanogenic biofilm of microbial electrosynthesis.
Applied Environment Microbiology, 86, 1–29. https://doi.org/10.1128/aem.01056-20.
Cheng S., Xing D., Call D. F. and Logan B. E. (2009). Direct biological conversion of electrical current into
methane by electromethanogenesis. Environmental Science and Technology, 43, 3953–3958, https://doi.
org/10.1021/es803531g
Choi O. and Sang B. I. (2016). Extracellular electron transfer from cathode to microbes: application for biofuel
production. Biotechnology for Biofuels, 9, 1–14. https://doi.org/10.1186/s13068-016-0426-0
Choi O., Kim T., Woo H. M. and Um Y. (2014). Electricity-driven metabolic shift through direct electron uptake
by electroactive heterotroph Clostridium pasteurianum. Scientific Reports 4, 6961. https://doi.org/10.1038/
srep06961
Cornils B., Herrmann W. A., Xu J. H. and Zanthoff H. W. (2020). Carbon dioxide capture and storage. Catalysis
From A to Z:A Concise Encyclopedia. John Wiley & Sons. https://doi.org/10.1002/9783527809080.
cataz02826
del Pilar Anzola Rojas M., Zaiat M., Gonzalez E. R., de Wever H. and Pant D. (2018). Effect of the electric supply
interruption on a microbial electrosynthesis system converting inorganic carbon into acetate. Bioresource
Technology, 266, 203–210. https://doi.org/10.1016/j.biortech.2018.06.074
del Pilar Anzola Rojas M., Zaiat M., González E. R., de Wever H. and Pant D. (2021). Enhancing the gas–liquid
mass transfer during microbial electrosynthesis by the variation of CO2 flow rate. Process Biochemistry, 101,
50–58. https://doi.org/10.1016/j.procbio.2020.11.005
DOE (2017). Biofuels and Bioproducts from Wet and Gaseous Waste Streams: Challenges and Opportunities.
Bionergy Technologies Office, Energy Efficiency and Renewable Energy, U.S. Department of Energy,
Washington, D.C. https://www.energy.gov/sites/default/files/2017/09/f36/biofuels_and_bioproducts_from_
wet_and_gaseous_waste_streams_full_report.pdf.
ElMekawy A., Hegab H. M., Mohanakrishna G., Elbaz A. F., Bulut M. and Pant D. (2016). Technological advances
in CO2 conversion electro-biorefinery: a step toward commercialization. Bioresource Technology, 215, 357–
370. https://doi.org/10.1016/j.biortech.2016.03.023
Gajaraj S., Huang Y., Zheng P. and Hu Z. (2017). Methane production improvement and associated methanogenic
assemblages in bioelectrochemically assisted anaerobic digestion. Biochemical Engineering Journal, 117,
105–112. https://doi.org/10.1016/j.bej.2016.11.003
Ganigué R., Puig S., Batlle-Vilanova P., Balaguer M. D. and Colprim J. (2015). Microbial electrosynthesis of butyrate
from carbon dioxide. Chemical Communications, 51, 3235–3238. https://doi.org/10.1039/c4cc10121a
Giddings C. G. S., Nevin K. P., Woodward T., Lovley D. R. and Butler C. S. (2015). Simplifying microbial
electrosynthesis reactor design. Front Microbiol, 6, 1–6. https://doi.org/10.3389/fmicb.2015.00468.
Gildemyn S., Verbeeck K., Slabbinck R., Andersen S. J., Prévoteau A. and Rabaey K. (2015). Integrated production,
extraction, and concentration of acetic acid from CO2 through microbial electrosynthesis. Environmental
Science and Technology Letters, 2, 325–328. https://doi.org/10.1021/acs.estlett.5b00212
House K. Z., Baclig A. C., Ranjan M., van Nierop E. A., Wilcox J. and Herzog H. J. (2011). Economic and energetic
analysis of capturing CO2 from ambient air. Proceedings of the National Academy of Sciences of the United
States of America, 108, 20428–20433. https://doi.org/10.1073/pnas.1012253108
Huang Z., Jiang D., Lu L. and Ren Z. J. (2016). Ambient CO– capture and storage in bioelectrochemically
mediated wastewater treatment. Bioresource Technology, 215, 380–385. https://doi.org/10.1016/j.
biortech.2016.03.084
Jack J., Zhu W., Avalos J., Gong J. and Ren Z. J. (2021). Anode co-valorization for scalable and sustainable
electrolysis. Green Chemistry, 23, 7917–7936. https://doi.org/10.1039/D1GC02094C
Jiang Y. and Zeng R. J. (2019). Bidirectional extracellular electron transfers of electrode-biofilm: mechanism and
application. Bioresource Technology, 271, 439–448. https://doi.org/10.1016/j.biortech.2018.09.133
Jiang Y., Lu L., Wang H., Shen R., Ge Z., Hou D., Chen X., Liang P., Huang X. and Ren Z. J. (2018). Electrochemical
control of redox potential arrests methanogenesis and regulates products in mixed culture electro-
fermentation. ACS Sustainable Chemistry & Engineering, 6(7), 8650–8658. https://doi.org/10.1021/
acssuschemeng.8b00948
Jiang Y., May H. D., Lu L., Liang P., Huang X. and Ren Z. J. (2019). Carbon dioxide and organic waste valorization by
microbial electrosynthesis and electro-fermentation. Water Research, 149, 42–55. https://doi.org/10.1016/j.
watres.2018.10.092
Jourdin L., Grieger T., Monetti J., Flexer V., Freguia S., Lu Y., Chen J., Romano M., Wallace G. G. and Keller
J. (2015). High acetic acid production rate obtained by microbial electrosynthesis from carbon dioxide.
Environmental Science and Technology, 49, 13566–13574. https://doi.org/10.1021/acs.est.5b03821
Jourdin L., Freguia S., Flexer V. and Keller J. (2016a). Bringing high-rate, CO2–based microbial electrosynthesis
closer to practical implementation through improved electrode design and operating conditions.
Environmental Science and Technology, 50, 1982–1989. https://doi.org/10.1021/acs.est.5b04431
Jourdin L., Lu Y., Flexer V., Keller J. and Freguia S. (2016b). Biologically induced hydrogen production drives
high rate/high efficiency microbial electrosynthesis of acetate from carbon dioxide. ChemElectroChem, 3,
581–591. https://doi.org/10.1002/celc.201500530
Kaza S., Yao L., Bhada-Tata P. and Van Woerden F. (2018). What A Waste 2.0: A Global Snapshot of Solid
Waste Management to 2050. Urban Development. World Bank, Washington, DC. © World Bank.
https://openknowledge.worldbank.org/handle/10986/30317 License: CC BY 3.0 IGO. https://doi.
org/10.1596/978-1-4648-1329-0
Lee K. Y., Ng T. W., Li G., An T., Kwan K. K., Chan K. M., Huang G., Yip H. Y. and Wong P. K. (2015). Simultaneous
nutrient removal, optimised CO2 mitigation and biofuel feedstock production by Chlorogonium sp. grown
in secondary treated non-sterile saline sewage effluent. Journal of hazardous materials, 297, pp. 241–250.
Liu S., Deng Z., Li H. and Feng K. (2019). Contribution of electrodes and electric current to process stability and
methane production during the electro-fermentation of food waste. Bioresource Technology, 288, 121536.
https://doi.org/10.1016/j.biortech.2019.121536
Logan B. E. (2008). Microbial Fuel Cells. Wiley Blackwell, Hoboken, New Jersey. https://doi.org/10.1002/9780470
258590
Logan B. E., Rossi R., Ragab A. and Saikaly P. E. (2019). Electroactive microorganisms in bioelectrochemical
systems. Nature Reviews Microbiology, 17, 307–319. https://doi.org/10.1038/s41579-019-0173-x
Lovley D. R. (2011). Powering microbes with electricity: direct electron transfer from electrodes to microbes.
Environmental Microbiology Reports, 3, 27–35, https://doi.org/10.1111/j.1758-2229.2010.00211.x
Lovley D. R. and Nevin K. P. (2011). A shift in the current: new applications and concepts for microbe-electrode
electron exchange. Current Opinion in Biotechnology, 3, 441–448. https://doi.org/10.1016/j.copbio.2011.01.009
Lu L., Huang Z., Rau G. H. and Ren Z. J. (2015). Microbial electrolytic carbon capture for carbon negative and
energy positive wastewater treatment. Environmental Science and Technology, 49, 8193–8201. https://doi.
org/10.1021/acs.est.5b00875
Lu L., Fang Y., Huang Z., Huang Y. and Ren Z. J. (2016). Self-sustaining carbon capture and mineralization
via electrolytic carbonation of coal fly ash. Chemical Engineering Journal, 306, 330–335. https://doi.
org/10.1016/j.cej.2016.07.060
Lu L., Guest J. S., Peters C. A., Zhu X., Rau G. H. and Ren Z. J. (2018). Wastewater treatment for carbon capture
and utilization. Nature Sustainability, 1, 750–758. https://doi.org/10.1038/s41893-018-0187-9
MacDowell N., Fennell P. S., Shah N. and Maitland G. C. (2017). The role of CO2 capture and utilization in
mitigating climate change. Nature Climate Change, 7, 243–249. https://doi.org/10.1038/nclimate3231
Mateos R., Escapa A., San-Martín M. I., de Wever H., Sotres A. and Pant D. (2020). Long-term open circuit
microbial electrosynthesis system promotes methanogenesis. Journal of Energy Chemistry, 41, 3–6. https://
doi.org/10.1016/J.JECHEM.2019.04.020
Mikkelsen M., Jørgensen M. and Krebs F. C. (2010). The teraton challenge. A review of fixation and transformation
of carbon dioxide. Energy and Environmental Science, 3, 43–81. https://doi.org/10.1039/b912904a
Mohanakrishna G., Seelam J. S., Vanbroekhoven K. and Pant D. (2015). An enriched electroactive homoacetogenic
biocathode for the microbial electrosynthesis of acetate through carbon dioxide reduction. Faraday Discuss,
183, 445–462. https://doi.org/10.1039/c5fd00041f.
Moscoviz R., Toledo-Alarcón J., Trably E. and Bernet N. (2016). Electro-fermentation: how to drive fermentation
using electrochemical systems. Trends in Biotechnology, 34, 11, 856–865. https://doi.org/10.1016/j.
tibtech.2016.04.009
Nie H., Zhang T., Cui M., Lu H., Lovley D. R. and Russel T. P. (2013). Improved cathode for high efficient
microbial-catalyzed reduction in microbial electrosynthesis cells. Phys. Chem. Chem. Phys., 15, 14290–
14294. https://doi.org/10.1039/c3cp52697f.
Nelabhotla A. B. T. and Dinamarca C. (2019). Bioelectrochemical CO2 reduction to methane: MES integration in
biogas production processes. Applied Sciences (Switzerland), 9, 16–18. https://doi.org/10.3390/app9061056
Nelabhotla A. B. T., Pant D. and Dinamarca C. (2021). In: Power-to-gas for methanation. Emerging Technologies
and Biological Systems for Biogas Upgrading, Aryal N., Ottosen L. D. M., Kofoed M. V. W. and Pant D. (eds),
London, UK, pp. 187–221. https://doi.org/10.1016/B978-0-12-822808-1.00008-8
Nevin K. P., Woodard T. L., Franks A. E., Summers Z. M. and Lovley D. R. (2010). Microbial electrosynthesis:
feeding microbes electricity to convert carbon dioxide and water to multicarbon extracellular organic
compounds. mBio, 1, e00103–10. https://doi.org/10.1128/mBio.00103-10
Nevin K. P., Hensley S. A., Franks A. E., Summers Z. M., Ou J., Woodard T. L., Snoeyenbos-West O. L. and
Lovley D. R. (2011). Electrosynthesis of organic compounds from carbon dioxide is catalyzed by a diversity
of acetogenic microorganisms. Applied and Environmental Microbiology, 77, 2882–2886. https://doi.
org/10.1128/AEM.02642-10
Nikhil G. N., Venkata Subhash G., Yeruva D. K. and Venkata Mohan S. (2015). Synergistic yield of dual energy
forms through biocatalyzed electrofermentation of waste: stoichiometric analysis of electron and carbon
distribution. Energy, 88, 281–291. https://doi.org/10.1016/j.energy.2015.05.043
Nishio K., Kimoto Y., Song J., Konno T., Ishihara K., Kato S., Hashimoto K. and Nakanishi S. (2013). Extracellular
electron transfer enhances polyhydroxybutyrate productivity in Ralstonia eutropha. Environmental Science
and Technology Letters, 1, 40–43. https://doi.org/10.1021/ez400085b
Pandey P., Shinde V. N., Deopurkar R. L., Kale S. P., Patil S. A. and Pant D. (2016). Recent advances in the use of
different substrates in microbial fuel cells toward wastewater treatment and simultaneous energy recovery.
Applied Energy, 168, 706–723. https://doi.org/10.1016/j.apenergy.2016.01.056
Prévoteau A., Carvajal-Arroyo J. M., Ganigué R. and Rabaey K. (2020). Microbial electrosynthesis from CO2:
forever a promise? Current Opinion in Biotechnology, 62, 48–57. https://doi.org/10.1016/j.copbio.2019.08.014
Rabaey K. and Rozendal R. A. (2010). Microbial electrosynthesis – revisiting the electrical route for microbial
production. Nature Reviews Microbiology, 8, 706–716. https://doi.org/10.1038/nrmicro2422
Rau G. H., Carroll S. A., Bourcier W. L., Singleton M. J., Smith M. M. and Aines R. D. (2013). Direct electrolytic
dissolution of silicate minerals for air CO2 mitigation and carbon-negative H2 production [WWW document].
Proceedings of the National Academy of Sciences of the United States of America, 110(25), 10095–10100.
https://doi.org/10.1073/pnas.1222358110
Russell J., Van Ballegooy S., Torvelainen E. and Gulley R. (2015). Consideration of ground variability over an area
of geological similarity as part of liquefaction assessment for foundation design. mBio, 1, 1–8. https://doi.
org/10.1128/mBio.00190-10.Editor
Saratale G. D., Saratale R. G., Shahid M. K., Zhen G., Kumar G., Shin H. S., Choi Y. G. and Kim S. H. (2017).
A comprehensive overview on electro-active biofilms, role of exo-electrogens and their microbial niches in
microbial fuel cells (MFCs). Chemosphere, 178, 534–547. https://doi.org/10.1016/j.chemosphere.2017.03.066
Sherrard J. H. (1976). Destruction of alkalinity in aerobic biological wastewater treatment. Journal of the Water
Pollution Control Federation, 48, 1834–1839.
Unesco, World Water Assessment Programme (2012). Taylor & Francis Group. https://doi.org/10.4324/9781849773355
Villano M., Paiano P., Palma E., Miccheli A. and Majone M. (2017). Electrochemically driven fermentation of
organic substrates with undefined mixed microbial cultures. ChemSusChem, 10, 3091–3097. https://doi.
org/10.1002/cssc.201700360
Wang H. and Ren Z. J. (2013). A comprehensive review of microbial electrochemical systems as a platform
technology. Biotechnology Advances, 31, 1796–1807. https://doi.org/10.1016/j.biotechadv.2013.10.001
Wett B., Eladawy A. and Becker W. (2004). Carbonate addition – an effective remedy against poor activated
sludge settling properties and alkalinity conditions in small wastewater treatment plants. Water Science and
Technology, 48(11), 411–417. https://doi.org/10.2166/wst.2004.0889
Xafenias N., Anunobi M. O. and Mapelli V. (2015). Electrochemical startup increases 1,3-propanediol titers in
mixed-culture glycerol fermentations. Process Biochemistry, 50(10), 1499–1508. https://doi.org/10.1016/j.
procbio.2015.06.020
Zaybak Z., Pisciotta J. M., Tokash J. C. and Logan B. E. (2013). Enhanced start-up of anaerobic facultatively
autotrophic biocathodes in bioelectrochemical systems. Journal of Biotechnology, 168, 478–485. https://doi.
org/10.1016/j.jbiotec.2013.10.001
Zhang T., Nie H., Bain T. S., Lu H., Cui M., Snoeyenbos-West O. L., Franks A. E., Nevin K. P., Russell T. P. and
Lovley D. R. (2013). Improved cathode materials for microbial electrosynthesis. Energy and Environmental
Science, 6, 217–224. https://doi.org/10.1039/c2ee23350a
Zhao Z., Li Y., Zhang Y. and Lovley D. R. (2020). Sparking anaerobic digestion: promoting direct interspecies electron
transfer to enhance methane production. Iscience, 23(12), 101794. https://doi.org/10.1016/j.isci.2020.101794
Zhou M., Chen J., Freguia S., Rabaey K. and Keller J. (2013). Carbon and electron fluxes during the electricity
driven 1,3-propanediol biosynthesis from glycerol. Environmental Science and Technology, 47, 11199–11205.
https://doi.org/10.1021/es402132r
Zhu X. and Logan B. E. (2014). Microbial electrolysis desalination and chemical-production cell for CO2
sequestration. Bioresource Technology, 159, 24–29. https://doi.org/10.1016/j.biortech.2014.02.062
Zhu X., Hatzell M. C. and Logan B. E. (2014). Microbial reverse-electrodialysis electrolysis and chemical-
production cell for H 2 production and CO2 sequestration. Environmental Science and Technology Letters, 1,
231–235. https://doi.org/10.1021/ez500073q
Zhu X., Leininger A., Jassby D., Tsesmetzis N. and Ren Z. J. (2021). Will membranes break barriers on volatile
fatty acid recovery from anaerobic digestion. ACS ES&T Engineering, 1(1), 141–153. https://doi.org/10.1021/
acsestengg.0c00081
Zou S. and He Z. (2018). Efficiently ‘pumping out’ value-added resources from wastewater by bioelectrochemical
systems: A review from energy perspectives. Water Research, 131, 62–73. https://doi.org/10.1016/j.
watres.2017.12.026
Chapter 7
Decarbonization potentials in
nitrogen management
Kester McCullough1,2*, Stephanie Klaus2 and Charles Bott2
1Cornell University, Ithaca, NY, USA
2Hampton Roads Sanitation District, Virginia Beach, VA, USA
*Correspondence: [email protected]
7.1 INTRODUCTION
Nitrogen management is a critical function of the water sector and is necessary to protect receiving
water bodies. This chapter explores the potential for decarbonization of nitrogen removal processes
within the water resource recovery facility (WRRF). Nitrogen is present in wastewater primarily in
organic forms and as ammonia, and most of the organic nitrogen is hydrolyzed to ammonia in the
treatment process. Ammonia is removed from wastewater by conversion to nitrogen gas through
oxidation (nitrification) and reduction (denitrification), or through anaerobic ammonia oxidation
(anammox). The equations governing these nitrogen transformations can be written as follows:
Autotrophic nitrification:
Nitritation, ammonia oxidizing bacteria (AOB) (yield = 0.15 gVSS/gNH4 + − N):
1.019NH4+ + 1.41O2 + 0.074CO2 + 0.019HCO3−
(7.1)
→ 0.019C 5H7O2N + NO2− + 2H+ + 0.981H2O
Heterotrophic denitrification:
Denitratation (yield = 0.36 gVSS/gCOD, methanol):
NO3− + 0.521CH 3OH + 0.056NH4+ + 0.056HCO3−
(7.3)
→ 0.056C 5H7O2N + NO2− + +0.2
296CO2 + 0.985H2O
© 2022 The Editors. This is an Open Access eBook distributed under the terms of the Creative Commons Attribution Licence
(CC BY-NC-ND 4.0), which permits copying and redistribution for non-commercial purposes with no derivatives, provided the
original work is properly cited (https://creativecommons.org/licenses/by-nc-nd/4.0/). This does not affect the rights licensed or
assigned from any third party in this book. The chapter is from the book Pathways to Water Sector Decarbonization, Carbon Capture
and Utilization, Zhiyong Jason Ren and Krishna Pagilla (Eds.)
Figure 7.1 The nitrogen cycle. Abbreviations: dissimilatory nitrate reduction to ammonium (DNRA); ammonia
oxidizing organisms (AOO) which includes bacteria (AOB) and archaea (AOA); nitrite oxidizing bacteria (NOB);
ordinary heterotrophic organisms (OHO).
These biological processes are the focus of this chapter (Figure 7.1). The artificial fixation of
ammonia from nitrogen gas via the Haber–Bosch process is driven by the need for ammonia fertilizer.
Research is currently at lab and pilot-scale to recover ammonia directly from wastewater as a
potentially more sustainable alternative (Beckinghausen et al., 2020; Ye et al., 2018). However, no
technology currently exists that can compete economically with biological nitrification/denitrification
at full-scale facilities (Winkler & Straka, 2019). WRRFs must consider the management of influent
(mainstream) nitrogen and nitrogen generated from waste sludges within the facility (sidestream);
approaches to decarbonization are specific to each of these contexts.
This introduction provides a broad overview of the carbon costs and decarbonization potentials
associated with nitrogen removal in WRRFs, followed by a detailed review and quantitative comparison
of technologies and process configurations for both sidestream and mainstream contexts. A review
of current research on the topic as well as a case study of a full-scale WRRF implementing these
strategies is provided. Finally, paths forward for the industry and the future outlook are considered.
Figure 7.2 Conventional WRRF layout with typical inputs and outputs shown in blue text and locations of potential
shortcut nitrogen removal improvements labeled with Green circles.
7.1.1.1 Aeration
In both mainstream and sidestream treatment, the energy required for aeration (to provide oxygen for
nitrification) generates a significant capital and carbon cost. Reducing the aeration energy required
for nitrogen removal involves shortcutting the nitrification/denitrification process and/or utilizing
anammox (a collection of strategies broadly referred to as shortcut nitrogen removal). Novel aeration
strategies, such as intermittent aeration or low dissolved oxygen, have the potential to decrease
both aeration energy demands and improve denitrification through simultaneous nitrification-
denitrification or dedicated anoxic periods. Optimized aeration controls, such as ammonia-based
aeration control (ABAC) or ammonia vs. NOx (AvN) control, also allow for more efficient aeration
and decreased energy usage.
7.1.1.2 Alkalinity
Aerobic nitrification consumes alkalinity, which must be managed to maintain pH and alkalinity
concentrations conducive to the growth of the organisms responsible for these processes. Denitrification
allows recovery of a fraction of the alkalinity consumed, but alkalinity supplementation is often
required, especially with low alkalinity wastewaters. The manufacture, transportation, storage, and
use of chemicals for alkalinity adjustment thus contributes to the carbon cost of nitrogen removal.
Shortcut nitrogen strategies generally decrease alkalinity demands.
strategies such as modifying the aeration strategy and minimizing nitrite concentrations without
impacting nitrogen removal performance (Duan et al., 2020; Pijuan et al., 2014). Mitigation strategies
for N2O in anoxic process include minimizing nitrite concentrations which can be controlled via
external carbon source dosing (Du et al., 2016; Song et al., 2015). It is important to mitigate N2O
production so that an increase in greenhouse gas production does not offset the carbon savings of
shortcut nitrogen removal (Chen et al., 2020).
depending on influent C:N, post-anoxic denitrification with external carbon addition is normally
required. Exceptions can be seen a relatively small plants without primary treatment and with long
‘extended aeration’ SRTs that maximize endogenous denitrification (e.g., some oxidation ditch processes).
Near complete N removal can be implemented both in the step-feed approach to which supplemental
carbon is added to the last anoxic zone, and combined with pre-anoxic denitrification in processes
such as a 4- or 5-stage Bardenpho, or UCT, VIP, or Johannesburg processes to which a second anoxic
zone is added. Because of the carbon demand for traditional nitrification/denitrification processes, it
is desirable to explore shortcut nitrogen removal technologies.
as a two-step process (van Kessel et al., 2015). The term Comammox (complete ammonia oxidation)
was coined to describe the process (van Kessel et al., 2015). The discovery of this pathway may help
to explain why NOB out-selection is so difficult (Daims et al., 2016).
7.3.4 Aeration, alkalinity, and COD requirements for mainstream nitrogen removal
technologies
The mainstream nitrogen removal processes presented thus far (nitrification/denitrification, nitrite
shunt, PNA, and PdNA) present varied opportunities for improved resource efficiency. As mentioned
before, it is often assumed that shortcut nitrogen removal (nitrite shunt, PNA, and PdNA) will
increase process efficiency, but this is not always the case for mainstream treatment. These potential
efficiency improvements are contingent on how influent COD is managed in an upstream process
(carbon diversion) and how the COD is utilized in the aerobic/nitrifying process.
Daigger (2014) demonstrated that, although PNA appears to require the least oxygen for nitrogen
removal, these benefits disappear when influent COD is used for complete TIN reduction in nitrification/
denitrification and nitrite shunt processes, as the additional oxygen required for nitrification is
‘recovered’. McCullough et al. (2022) expanded this analysis to include alkalinity and supplemental
COD requirements and the PdNA process. This analysis, which incorporated yield and assimilation,
also included cases where complete TIN removal was not achieved with influent COD, which more
accurately reflects most WRRFs. These results are summarized in Table 7.1. Aeration, alkalinity, and
supplemental COD requirements for each nitrogen removal process are a function of how efficiently
Table 7.1 Aeration, alkalinity, and supplemental COD requirements for complete nitrogen removal in nitrification/
denitrification, nitrite shunt, PdNA, and PNA processesa.
influent COD is used for TIN removal, and process requirements are similar with complete (100%)
TIN removal with influent COD. When complete TIN removal cannot be achieved with influent COD,
the varied process efficiencies are significant. Thus, decarbonization can be achieved by increasing
the efficiency with which influent COD is used for TIN removal and by transitioning to more efficient
nitrogen removal processes. PNA is the most efficient of these processes, but also the most difficult
to implement due to NOB out-selection. PdNA provides comparable resource efficiency increases but
does not require NOB out-selection.
Figure 7.3 Maximum allowable upstream COD removal for complete nitrogen removal in conventional and shortcut
nitrogen removal processes, shown as a function of how efficiently influent COD is used for nitrogen reduction
in the treatment process. As influent COD is used more efficiently for nitrogen reduction, greater amounts can
be removed in the upstream COD diversion process. Shortcut nitrogen removal processes allow for greater COD
diversion/capture without compromising nitrogen removal.
removal in the A-Stage process can exceed 10%, and this nitrogen ultimately ends in sludge streams
where it can be managed more efficiently in sidestream processes.
Upstream carbon diversion processes also provide the possibility for redirection carbon in the
treatment process to where it can be used most beneficially. Fermentation of captured COD can
produce an effluent rich in volatile fatty acids (VFA), which can be used for mainstream or sidestream
biological phosphorus removal or denitrification.
Shortcut nitrogen removal allows for upstream carbon diversion, providing maximal potential
for decarbonization as the aeration, alkalinity, and carbon requirements for nitrogen removal can
be drastically reduced, the plant capacity can be increased through SRT reduction, and diverted
carbon can be used to generate electricity, heat, or VFA to use as a carbon source elsewhere in the
process. Maximizing the efficient use of influent COD for denitrification further decreases the carbon
requirements of the process and allows for more carbon redirection. If carbon redirection cannot be
implemented, shortcut nitrogen removal processes still provide increased plant capacity and reduce the
exogenous carbon costs of nitrogen removal, but the full benefits of these processes will not be realized.
heterotrophic denitrification using NO3− or NO2− without the addition of supplemental carbon. This
can be achieved with either continuous or intermittent aeration. Another option for AvN is the slope-
intercept control concept using the following equation: NH4+ = slope*NOx + intercept where the slope
controls the NH4+/NOx ratio and the intercept controls the ammonia effluent limit. When implemented
for mainstream deammonification, the slope will be higher as more nitrogen is removed through the
anammox pathway. Controlling the ratio of NH4+ to NOx is required in the first stage of two-stage PNA
and PdNA systems in order to meet the proper stoichiometry for anammox downstream.
Figure 7.4 The Hampton Roads Sanitation District (HRSD) York River Treatment Plant (YRTP) in Seaford, VA.
to the denitrification filters by approximately 25% (Nifong et al., 2013). AvN control was achieved
manually through a combination of aeration and step feed control. After secondary clarification and
intermediate pumping are deep-bed filters configured for sensor-driven methanol feed control. The
filters were sized for an expansion of the plant and are rated for 114 000 m3/day, a nitrate load of
0.4 kgN/m3/day and 5.9 m/hr. At current flows (2021), the filter is underloaded at an average flow
rate of 2.4 m/hr. Ammonia removed anoxically in the denitrification filter meant that less ammonia
needed to be oxidized upstream in the BNR process. So, when 50% of the aeration volume was turned
into anoxic zones with step feed to better utilize influent carbon for denitrification, the PdNA filters
were able to make up for this loss of aerobic SRT (Table 7.2). The transition from full nitrification
to two-pass step feed resulted in a methanol saving of approximately 60%. The transition from full-
denitrification to PdNA, which soon followed the implementation of two-pass step feed, resulted in
an additional 50% methanol saving (Fofana et al. in progress). This saving in units of methanol added
per nitrogen removed is shown in Table 7.2.
Table 7.2 YRTP capacity increase and methanol reduction due to PdNA process implementation.
processes. Whole plant level optimization is needed to coordinate C, N, P removal and recovery in
order to accomplish decarbonization of the sector.
While N recovery through urine separation may be cost-effective and practical, the recovery
of N-based fertilizers from wastewater remains challenging, as the cost of biological N removal is
declining with shortcut opportunities, and emerging technologies that have the potential to recover
N are currently not cost-competitive with Haber–Bosch production of N fertilizers. N recovery is
generally limited to the land application of biosolids and biosolids products and the limited amount of
N that is precipitated with struvite recovery.
One of the great challenges in preparing a summary of advances in N removal technology is that
each facility is very different from a treatment technology standpoint. Other relevant differences
include TN and NH4 + limits, energy costs, chemical costs, and ‘available’ capacity. Decarbonization
often depends on the starting point or the baseline plant condition, and this cannot be generalized or
ignored.
For example, a small plant with stringent TN limits, a five-stage Bardenpho process, and no
effective DO control (gross over aeration with DO at say 4–6 mg/L) could gain considerable energy
and supplemental carbon chemical benefits by transitioning to ABAC. This seems like an obvious
improvement. However, we must be careful not decarbonize wastewater treatment at the expense
of plant capacity. Carrying our example further, if this same plant implements ABAC moving to an
average DO setpoint of say 0.5 mg/L, then the implication is an inherent loss in nitrification capacity;
decreased effective capacity of the plant in terms of aeration tank volume and secondary clarifier area.
This is often an unacceptable compromise, even though some optimization of this plant is clearly
warranted.
The goal of decarbonizing in the context of N removal must always consider capacity implications.
This is critical from the standpoint of gaining utility consensus and buy-in. In fact, as our industry
contemplates advanced N removal technologies, the ideal direction involves decarbonizing at the same
time as increasing effective capacity. This is collectively known as intensification. It is quite relevant
here, because the push towards shortcut N removal can provide the dual benefit of decarbonization
with no compromise in capacity, and in fact often depending on the baseline scenario, a dramatic
increase in plant capacity, and/or the excess aerobic zone capacity could be dedicated to anaerobic or
pre-anoxic zones to better utilize wastewater carbon for N and P removal. The real key to intensification
is directing NH4 + to anoxic oxidation to N2 by anammox, allowing for decreased operating/design
aerobic SRT while at the same time ensuring that reliably low TN and NH4 + limits can be met.
Sidestream PNA processes can be an important and nearly obvious part of that intensification
incentive. These processes are now mature and available, but they are not ‘plug-and-play,’ still requiring
considerable operator attention and knowledge for successful operation. Even then, upsets do occur.
After more than 15 years of research, mainstream PNA has been mostly unsuccessful. The few
reports of full-scale testing an implementation do not give any indication that significant amounts of
influent NH4 + are being directed to anammox. Pilot and laboratory testing results have shown promise
in some cases, but this has not led to scalable technology. The industry has generally determined
that the low growth rate of anammox can be accommodated by selective retention using biofilms
and granules. However, consistent and reliable out-selection of NOB has proven very difficult, and
this is perhaps made even more complicated by the existence of comammox. With what we know
now, there is little promise of legitimate mainstream anammox technology that relies on PNA. That
said, additional cost savings could be provided by processes that can periodically take advantage of
unsustained and unreliable NO2− production from NOB repression by directing residual NH4 + and
NO2− to anammox-based polishing processes.
Mainstream PdNA can provide this opportunity to take advantage of NO2− produced either
from partial denitrification or from periods of NOB out-selection but, more importantly, it offers
performance reliability and the intensification benefit. It is argued here that PdNA does not make
sense to be considered for sidestream treatment, but it is immediately valuable for mainstream in the
situations where low TN limits are required. PdNA seems best applied in the form of a post polishing
process or integrated into a downstream anoxic zone in a step feed BNR or Bardenpho-style process.
In these polishing PdNA applications (anoxic zone or post-secondary process), perhaps 10–20% of
the influent TKN can be directed to anammox. This is a huge benefit from a capacity standpoint, and
it offers considerable operating cost savings benefits. The question remains whether PdNA can be
further developed to allow a larger fraction of the influent TKN to be directed to anammox. This of
course requires two things – use of wastewater COD for partial denitrification and likely some degree
of carbon diversion. Although the benefits may never be quite as good as PNA, the foundation has
been laid for mainstream PdNA.
REFERENCES
Al-Omari A., Wett B., Nopens I., De Clippeleir H., Han M., Regmi P., Bott C. and Murthy S. (2015). Model-based
evaluation of mechanisms and benefits of mainstream shortcut nitrogen removal processes. Water Science
and Technology, 71(6), 840–847. https://doi.org/10.2166/wst.2015.022
Anthonisen A. C., Loehr R. C., Prakasam T. B. S. and Srinath E. G. (1976). Inhibition of nitrification by ammonia
and nitrous acid. Journal (Water Pollution Control Federation), 48(5), 835–852.
Beckinghausen A., Odlare M., Thorin E. and Schwede S. (2020). From removal to recovery: an evaluation of
nitrogen recovery techniques from wastewater. Applied Energy, 263, 114616. https://doi.org/10.1016/j.
apenergy.2020.114616
Campolong C., Klaus S., Rosenthal A., Sabba F., Baidme M., Wells G., Wett B., Clippeleir H. D., Chandran K.
and Bott C. (2019). Comparison of external carbon sources for a polishing partial denitrification/anammox
MBBR. Proceedings of the Water Environment Foundation, 56–63.
Cao Y. S., Kwok B. H., Yong W. H., Chua S. C., Wah Y. L. and Yahya A. G. (2013). The main stream autotrophic
nitrogen removal in the largest full scale activated sludge process in Singapore: process analysis.
Proceedings of WEF/IWA Nutrient Removal and Recovery 2013: Trends in Resource Recovery and Use.
Water Environment Federation, Alexandria, Virginia, pp. 28–31.
Cao Y., van Loosdrecht M. C. M. and Daigger G. T. (2017). Mainstream partial nitritation–anammox in municipal
wastewater treatment: status, bottlenecks, and further studies. Applied Microbiology and Biotechnology,
101(4), 1365–1383. https://doi.org/10.1007/s00253-016-8058-7
Cao Y., Kwok B. H., van Loosdrecht M. C. M., Daigger G., Png H. Y., Long W. Y. and Eng O. K. (2018). The
influence of dissolved oxygen on partial nitritation/anammox performance and microbial community of the
200 000 m3/d activated sludge process at the Changi water reclamation plant (2011 to 2016). Water Science
and Technology, 78(3), 634–643. https://doi.org/10.2166/wst.2018.333
Chen H., Zeng L., Wang D., Zhou Y. and Yang X. (2020). Recent advances in nitrous oxide production and mitigation
in wastewater treatment. Water Research, 184, 116168. https://doi.org/10.1016/j.watres.2020.116168
Courtens E. N. P., De Clippeleir H., Vlaeminck S. E., Jordaens R., Park H., Chandran K. and Boon N. (2015).
Nitric oxide preferentially inhibits nitrite oxidizing communities with high affinity for nitrite. Journal of
Biotechnology, 193, 120–122. https://doi.org/10.1016/j.jbiotec.2014.11.021
Cui B., Yang Q., Liu X., Wu W., Liu Z. and Gu P. (2020). Achieving partial denitrification-anammox in biofilter
for advanced wastewater treatment. Environment International, 138, 105612. https://doi.org/10.1016/j.
envint.2020.105612
Daigger G. T. (2014). Oxygen and carbon requirements for biological nitrogen removal processes accomplishing
nitrification, nitritation, and anammox. Water Environment Research, 86(3), 204–209. https://doi.org/10.2
175/106143013X13807328849459
Daims H., Lücker S. and Wagner M. (2016). A new perspective on microbes formerly known as nitrite-oxidizing
bacteria. Trends in Microbiology, 24(9), 699–712. https://doi.org/10.1016/j.tim.2016.05.004
Delgado Vela J., Stadler L. B., Martin K. J., Raskin L., Bott C. B. and Love N. G. (2015). Prospects for biological
nitrogen removal from anaerobic effluents during mainstream wastewater treatment. Environmental Science
& Technology Letters, 2(9), 234–244. https://doi.org/10.1021/acs.estlett.5b00191
Di Capua F., Pirozzi F., Lens P. N. L. and Esposito G. (2019). Electron donors for autotrophic denitrification.
Chemical Engineering Journal, 362, 922–937. https://doi.org/10.1016/j.cej.2019.01.069
Du R., Peng Y., Cao S., Wang S. and Niu M. (2016). Characteristic of nitrous oxide production in partial
denitrification process with high nitrite accumulation. Bioresource Technology, 203, 341–347.
Du R., Cao S., Li B., Niu M., Wang S. and Peng Y. (2017). Performance and microbial community analysis of a
novel DEAMOX based on partial-denitrification and anammox treating ammonia and nitrate wastewaters.
Water Research, 108, 46–56.
Duan H., van den Akker B., Thwaites B. J., Peng L., Herman C., Pan Y., Ni B.-J., Watt S., Yuan Z. and Ye L. (2020).
Mitigating nitrous oxide emissions at a full-scale wastewater treatment plant. Water Research, 185, 116196.
https://doi.org/10.1016/j.watres.2020.116196
Fofana R., Parsons M., Long C., Chandran K., Jones K., Klaus S., Trovato B., Wilson C., De Clippeleir H. and
Bott C. (submitted). Full-scale transition from denitrification to Partial denitrification – anammox (PdNA)
in Deep-Bed filters: Operational strategies for and benefits of PdNA implementation. Water Environment
Research.
Gao D.-W., Huang X.-L., Tao Y., Cong Y. and Wang X. (2015). Sewage treatment by an UAFB–EGSB biosystem with
energy recovery and autotrophic nitrogen removal under different temperatures. Bioresource Technology,
181, 26–31. https://doi.org/10.1016/j.biortech.2015.01.037
Gilbert E. M., Agrawal S., Brunner F., Schwartz T., Horn H. and Lackner S. (2014a). Response of different
nitrospira species to anoxic periods depends on operational DO. Environmental Science and Technology,
48(5), 2934–2941. https://doi.org/10.1021/es404992g
Gilbert E. M., Agrawal S., Karst S. M., Horn H., Nielsen P. H. and Lackner S. (2014b). Low temperature partial
nitritation/anammox in a moving bed biofilm reactor treating low strength wastewater. Environmental
Science and Technology, 48(15), 8784–8792. https://doi.org/10.1021/es501649m
Gustavsson D. J. I., Suarez C., Wilén B.-M., Hermansson M. and Persson F. (2020). Long-term stability of partial
nitritation-anammox for treatment of municipal wastewater in a moving bed biofilm reactor pilot system.
Science of the Total Environment, 714, 136342. https://doi.org/10.1016/j.scitotenv.2019.136342
Han M., Vlaeminck S. E., Al-Omari A., Wett B., Bott C., Murthy S. and De Clippeleir H. (2016). Uncoupling the
solids retention times of flocs and granules in mainstream deammonification: a screen as effective out-
selection tool for nitrite oxidizing bacteria. Bioresource Technology, 221, 195–204. https://doi.org/10.1016/j.
biortech.2016.08.115
Hellinga C., Schellen A., Mulder J. W., van Loosdrecht M. C. M. and Heijnen J. J. (1998). The SHARON process: an
innovative method for nitrogen removal from ammonium-rich waste water. Water Science and Technology,
37(9), 135–142. https://doi.org/10.2166/wst.1998.0350
Jetten M. S. M., Horn S. J. and van Loosdrecht M. C. M. (1997). Towards a more sustainable municipal
wastewater treatment system. Water Science and Technology, 35(9), 171–180. https://doi.org/10.2166/
wst.1997.0341
Jimenez J., Wise G., Regmi P., Burger G., Conidi D., Du W. and Dold P. (2020). Nitrite-shunt and biological
phosphorus removal at low dissolved oxygen in a full-scale high-rate system at warm temperatures. Water
Environment Research, 92(8), 1111–1122. https://doi.org/10.1002/wer.1304
Kampschreur M. J., Temmink H., Kleerebezem R., Jetten M. S. and van Loosdrecht M. C. (2009). Nitrous oxide
emission during wastewater treatment. Water Research, 43(17), 4093–4103. https://doi.org/10.1016/j.
watres.2009.03.001
Kartal B., Kuenen J. G. and van Loosdrecht M. C. M. (2010). Sewage treatment with anammox. Science, 328(5979),
702–703. https://doi.org/10.1126/science.1185941
Kornaros M., Dokianakis S. N. and Lyberatos G. (2010). Partial nitrification/denitrification can be attributed to
the slow response of nitrite oxidizing bacteria to periodic anoxic disturbances. Environmental Science and
Technology, 44(19), 7245–7253. https://doi.org/10.1021/es100564j
Lackner S., Gilbert E. M., Vlaeminck S. E., Joss A., Horn H. and van Loosdrecht M. C. M. (2014). Full-scale
partial nitritation/anammox experiences – an application survey. Water Research, 55, 292–303. https://doi
.org/10.1016/j.watres.2014.02.032
Lackner S., Welker S., Gilbert E. M. and Horn H. (2015). Influence of seasonal temperature fluctuations on two
different partial nitritation-anammox reactors treating mainstream municipal wastewater. Water Science
and Technology, 72(8), 1358–1363. https://doi.org/10.2166/wst.2015.301
Laureni M., Falås P., Robin O., Wick A., Weissbrodt D. G., Nielsen J. L., Ternes T. A., Morgenroth E. and Joss A.
(2016). Mainstream partial nitritation and anammox: long-term process stability and effluent quality at low
temperatures. Water Research, 101, 628–639. https://doi.org/10.1016/j.watres.2016.05.005
Le T., Peng B., Su C., Massoudieh A., Torrents A., Al-Omari A., Murthy S., Wett B., Chandran K., DeBarbadillo
C., Bott C. and Clippeleir H. D. (2019). Impact of carbon source and COD/N on the concurrent operation of
partial denitrification and anammox. Water Environment Research, 91(3), 185–197. https://doi.org/10.1002/
wer.1016
Liu M., Krikorian E., Knapp T., Melitas N., Zhao H. and Johnson M. (2018). Mainstream ANITA Mox pilot
testing at the joint water pollution control plant. Proceedings of the Water Environment Federation, 2018(5),
169–184. https://doi.org/10.2175/193864718824940385
Liu Y.-J., Gu J. and Liu Y. (2018). Energy self-sufficient biological municipal wastewater reclamation: present
status, challenges and solutions forward. Bioresource Technology, 269, 513–519. https://doi.org/10.1016/
j.biortech.2018.08.104
Lotti T., Kleerebezem R., Hu Z., Kartal B., Jetten M. S. M. and van Loosdrecht M. C. M. (2014). Simultaneous
partial nitritation and anammox at low temperature with granular sludge. Water Research, 66, 111–121.
https://doi.org/10.1016/j.watres.2014.07.047
Ma B., Zhang S., Zhang L., Yi P., Wang J., Wang S. and Peng Y. (2011). The feasibility of using a two-stage
autotrophic nitrogen removal process to treat sewage. Bioresource Technology, 102(17), 8331–8334. https://
doi.org/10.1016/j.biortech.2011.06.017
Ma B., Wang S., Cao S., Miao Y., Jia F., Du R. and Peng Y. (2016). Biological nitrogen removal from sewage via anammox:
recent advances. Bioresource Technology, 200, 981–990. https://doi.org/10.1016/j.biortech.2015.10.074
McCullough K., Klaus S., Parsons M., Wilson C. and Bott C. B. (2022). Advancing the Understanding of Mainstream
Shortcut Nitrogen Removal: Resource Efficiency, Carbon Redirection, and Plant Capacity. Water Research
(submitted).
Miller M. W., Bunce R., Regmi P., Hingley D. M., Kinnear D., Murthy S., Wett B. and Bott C. B. (2012). A/B
process pilot optimized for nitrite shunt: high rate carbon removal followed by BNR with ammonia-based
cyclic aeration control. Proceedings of the Water Environment Federation, 2012(10), 5808–5825. https://doi.
org/10.2175/193864712811709607
Morales N., Val del Río Á., Vázquez-Padín J. R., Méndez R., Campos J. L. and Mosquera-Corral A. (2016). The
granular biomass properties and the acclimation period affect the partial nitritation/anammox process
stability at a low temperature and ammonium concentration. Process Biochemistry, 51(12), 2134–2142.
https://doi.org/10.1016/j.procbio.2016.08.029
Mulder A. (2003). The quest for sustainable nitrogen removal technologies. Water Science and Technology, 48(1),
67–75. https://doi.org/10.2166/wst.2003.0018
Ni B.-J. and Yuan Z. (2015). Recent advances in mathematical modeling of nitrous oxides emissions from
wastewater treatment processes. Water Research, 87, 336–346. https://doi.org/10.1016/j.watres.2015.09.049
Nifong A., Nelson A., Johnson C. and Bott C. B. (2013). Performance of a full-scale sidestream DEMON®
deammonification installation. Proceedings of the Water Environment Federation, 2013(13), 3686–3709.
https://doi.org/10.2175/193864713813685700
Fofana R., Bachmann M., Akyon B., Jones K., Delgado J., Klaus S., Parsons M., Bott C., deBarbadillo C. and
De Clippeleir H. (2021). Carbon Source Selection for Deep-Bed Partial Denitrification: Anammox (PdNA)
Polishing Filters. WEFTEC 2021. Chicago, IL: WEF.
Pérez J., Lotti T., Kleerebezem R., Picioreanu C. and van Loosdrecht M. C. (2014). Outcompeting nitrite-oxidizing
bacteria in single-stage nitrogen removal in sewage treatment plants: a model-based study. Water Research,
66, 208–218. https://doi.org/10.1016/j.watres.2014.08.028
Piculell M., Christensson M., Jönsson K. and Welander T. (2016a). Partial nitrification in MBBRs for mainstream
deammonification with thin biofilms and alternating feed supply. Water Science and Technology, 73(6),
1253–1260. https://doi.org/10.2166/wst.2015.599
Piculell M., Suarez C., Li C., Christensson M., Persson F., Wagner M., Hermansson M., Jönsson K. and Welander T.
(2016b). The inhibitory effects of reject water on nitrifying populations grown at different biofilm thickness.
Water Research, 104, 292–302. https://doi.org/10.1016/j.watres.2016.08.027
Pijuan M., Torà J., Rodríguez-Caballero A., César E., Carrera J. and Pérez J. (2014). Effect of process parameters
and operational mode on nitrous oxide emissions from a nitritation reactor treating reject wastewater. Water
Research, 49, 23–33. https://doi.org/10.1016/j.watres.2013.11.009
Poot V., Hoekstra M., Geleijnse M. A., van Loosdrecht M. C. and Pérez J. (2016). Effects of the residual ammonium
concentration on NOB repression during partial nitritation with granular sludge. Water Research, 106,
518–530. https://doi.org/10.1016/j.watres.2016.10.028
Regmi P., Miller M. W., Holgate B., Bunce R., Park H., Chandran K., Wett B., Murthy S. and Bott C. B. (2014).
Control of aeration, aerobic SRT and COD input for mainstream nitritation/denitritation. Water Research,
57, 162–171. https://doi.org/10.1016/j.watres.2014.03.035
Regmi P., Holgate B., Miller M. W., Park H., Chandran K., Wett B., Murthy S. and Bott C. B. (2016). Nitrogen
polishing in a fully anoxic anammox MBBR treating mainstream nitritation–denitritation effluent.
Biotechnology and Bioengineering, 113(3), 635–642. https://doi.org/10.1002/bit.25826
Sharp R., Niemiec A., Khunjar W., Galst S. and Deur A. (2017). Development of a novel deammonification process
for cost effective separate centrate and main plant nitrogen removal. International Journal of Sustainable
Development and Planning, 12(1), 11–21.
Sliekers A. O., Haaijer S. C. M., Stafsnes M. H., Kuenen J. G. and Jetten M. S. M. (2005). Competition and
coexistence of aerobic ammonium- and nitrite-oxidizing bacteria at low oxygen concentrations. Applied
Microbiology and Biotechnology, 68(6), 808–817. https://doi.org/10.1007/s00253-005-1974-6
Soler-Jofra A., Pérez J. and van Loosdrecht M. C. M. (2021). Hydroxylamine and the nitrogen cycle: a review.
Water Research, 190, 116723. https://doi.org/10.1016/j.watres.2020.116723
Song K., Riya S., Hosomi M. and Terada A. (2015). Effect of carbon sources on nitrous oxide emission in a
modified Ludzak Ettinger process. Water Science and Technology, 72(4), 572–578.
Strous M., Heijnen J. J., Kuenen J. G. and Jetten M. S. M. (1998). The sequencing batch reactor as a powerful tool
for the study of slowly growing anaerobic ammonium-oxidizing microorganisms. Applied Microbiology and
Biotechnology, 50, 589–596. https://doi.org/10.1007/s002530051340
Tchobanoglous G., Burton F.L. and Stensel H.D. (2003). Wastewater Engineering: Treatment and Reuse, McGraw-
Hill, Boston.
van Kessel M. A. H. J., Speth D. R., Albertsen M., Nielsen P. H., Op den Camp H. J. M., Kartal B., Jetten M. S. M.
and Lücker S. (2015). Complete nitrification by a single microorganism. Nature, 528(7583), 555–559. https://
doi.org/10.1038/nature16459
Von Sperling M. and de Lemos Chernicharo C. (2005). Biological Wastewater Treatment in Warm Climate Regions
Volume I, IWA Publishing. See https://doi.org/10.2166/9781780402734 (Accessed 25 February 2022).
Vlaeminck S. E., Clippeleir H. D. and Verstraete W. (2012). Microbial resource management of one-
stage partial nitritation/anammox. Microbial Biotechnology, 5(3), 433–448. https://doi.org/10.1111/
j.1751-7915.2012.00341.x
Wang Q., Ye L., Jiang G., Hu S. and Yuan Z. (2014). Side-stream sludge treatment using free nitrous acid selectively
eliminates nitrite oxidizing bacteria and achieves the nitrite pathway. Water Research, 55, 245–255. https://
doi.org/10.1016/j.watres.2014.02.029
Welker S., Horn H. and Lackner S. (2016). Substrate contentment: Influence of residual ammonium and dissolved
oxygen concentrations on autotrophic nitrogen removal. Proceedings of WEF/IWA Nutrient Removal and
Recovery, Denver, USA. Water Environment Federation, Alexandria, Virginia.
Wett B., Podmirseg S. M., Gómez-Brandón M., Hell M., Nyhuis G., Bott C. and Murthy S. (2015). Expanding
DEMON sidestream deammonification technology towards mainstream application. Water Environment
Research, 87(12), 2084–2089. https://doi.org/10.2175/106143015X14362865227319
Winkler M. K. and Straka L. (2019). New directions in biological nitrogen removal and recovery from wastewater.
Current Opinion in Biotechnology, 57, 50–55. https://doi.org/10.1016/j.copbio.2018.12.007
Winkler M.-K. H., Kleerebezem R. and van Loosdrecht M. C. M. (2012). Integration of anammox into the aerobic
granular sludge process for main stream wastewater treatment at ambient temperatures. Water Research,
46(1), 136–144. https://doi.org/10.1016/j.watres.2011.10.034
Wunderlin P., Mohn J., Joss A., Emmenegger L. and Siegrist H. (2012). Mechanisms of N2O production in
biological wastewater treatment under nitrifying and denitrifying conditions. Water Research, 46(4), 1027–
1037. https://doi.org/10.1016/j.watres.2011.11.080
Ye Y., Ngo H. H., Guo W., Liu Y., Chang S. W., Nguyen D. D., Liang H. and Wang J. (2018). A critical review on
ammonium recovery from wastewater for sustainable wastewater management. Bioresource Technology,
268, 749–758. https://doi.org/10.1016/j.biortech.2018.07.111
Chapter 8
Decarbonization potentials in
phosphorus management in the
water sector
Annalisa Onnis-Hayden1, Dongqi Wang2, Ali Akbari3, Mi Nguyen3 and April Z. Gu3*
1Northeastern University, Boston, MA, USA
2Xi’an University of Technology, Xi’an, Shaanxi, China
3Cornell University, Ithaca, NY, USA
*Correspondence: [email protected]
© 2022 The Editors. This is an Open Access eBook distributed under the terms of the Creative Commons Attribution Licence
(CC BY-NC-ND 4.0), which permits copying and redistribution for non-commercial purposes with no derivatives, provided the
original work is properly cited (https://creativecommons.org/licenses/by-nc-nd/4.0/). This does not affect the rights licensed or
assigned from any third party in this book. The chapter is from the book Pathways to Water Sector Decarbonization, Carbon Capture
and Utilization, Zhiyong Jason Ren and Krishna Pagilla (Eds.)
Figure 8.1 Mass flow analysis of major global phosphorus sinks and sources depicting the significant P lost through
the cycle. The values in parentheses are annual load in million metric tons and retrieved from Liu et al. (2008),
Cordell et al. (2009) and Rittmann et al. (2011). Different places where decarbonization is relevant are included with
the associated carbon footprint. The carbon footprint of riparian buffer, and P removal from wastewater are from
Coats et al. (2011) and Styles et al. (2016), respectively. For transport the carbon footprint was calculated assuming
40 km transport of manure by a diesel truck.
The overall global P flow (Figure 8.1) starts with mining, and most of the mined P (80% of mined
P, 25.4 Mmmt. yr−1 including recycled P) is used for agricultural purposes (Rittmann et al., 2011).
The harvested crops are either utilized for livestock or directly used in the food industry. Industrial
applications and the production of detergents and P supplements for animal feeding constitute a
smaller fraction of P demand (Liu et al., 2008). Due to the extensive losses through various processing
stages, only13–17% of mined P (3 Mmt.yr−1) is eventually consumed by humans. Losses through non-
point sources such as agricultural run-off and erosion constitute at least half of mined P (>8 Mmt.yr−1
(Rittmann et al., 2011). The global P load into water resources (22 Mmt.yr−1) are mainly originated
from non-point sources (about 90%) (Cordell et al., 2009). As high as 30% of P used in croplands are
applied to soils with previously high residual P levels which makes soils the largest pool of P in the
environments (Alewell et al., 2020; Bouwman et al., 2013). The prolonged but finite accumulation
phase of P in soils requires adaptive management strategies to protect water resources. Despite the
significant volume of P released from non-point sources, the management of these sources has gained
relatively less attention. In comparison, there have been more advancements in removal and recovery
technologies from point sources, including municipal wastewater treatment plants or industrial
settings such as confined livestock facilities.
Integrated source-watershed management, containment, treatment, and recovery strategies are
needed to mitigate the environmental impacts associated with P flow through various processes and
environmental compartments. Holistic thinking requires that the associated carbon footprints with
various integrated interventions be taken into account.
For example, the phosphorus trading among point and non-point sources is getting more attention
and is likely to be implemented increasingly in the near future. In such a scenario, more stringent
effluent requirements from point-source treatment units are traded with implementing non-point
source management facilities such as constructed wetlands at the sink (area of release to surface
waterways). An example, resembling the carbon cap and trade, is a case in Dixie drain facility in
Boise, Idaho, where P trading led to a significantly less costly integrated, regional P management
(Macintosh et al., 2018). Because of the high carbon footprints of advanced chemical-based P removal
technologies needed to achieve the more stringent effluent requirements, integrated trading likely offers
a more sustainable resort. Previous studies have quantified significant total direct and indirect carbon
footprints of P removal processes in WWTPs (e.g., 0.1 kg CO2-e.m−3 treated wastewater (Coats et al.,
2011; Rahman et al., 2016) as a major source of emission in the P management cycle. That underlines
the significance of developing technologies with lower carbon footprints. The local, regional, direct,
and indirect impacts of P removal strategies should be assessed, and the environmental hotspots
identified toward sustainability-oriented improvements.
Figure 8.2 Various technologies related to P management from point or nonpoint sources at multiple levels. The
cost and environmental consequences are likely to increase from higher-level management practices for utilization
efficiency improvement to lower levels of containment treatment and recovery.
mitigation process due to the lower chemical use. Generally, preventive strategies are less expensive
and more sustainable alternatives to removal and treatment practices.
At the lowest level is the most expensive and environmentally consequential resort of removal
and recovery. These strategies are the last line of defense to protect the water resources and include
riparian buffers and wetlands for non-point sources and biological/chemical phosphorus removal and
recovery units in wastewater treatment plants. The surface layer in a wetland with high P content
provides an opportunity for recycling the lost P to the agriculture field. An external carbon-independent
modification of biological enhanced phosphorus removal units has been introduced, offering a lower
carbon footprint than conventional technologies (Wang et al., 2019). Novel treatments and recovery
technologies are yet to be developed toward more sustainable P management. A systematic comparative
sustainability analysis, particularly for non-point source P management practices, is currently missing.
Such holistic analyses should introduce and assess multi-tier, integrated management strategies.
programs including the Environmental Quality Incentives Program (EQIP) and the Conservation
Technical Assistance Program (CTA) are provided to farmers by the US Department of Agriculture
to encourage the implementation of nutrient management practices. The EQIP provides financial
and technical assistance to agricultural producers to address natural resource concerns and deliver
environmental benefits. The CTA program is to provide technical assistance to land users in planning
and implementing conservation systems including nutrient management for improving air, soil and
water quality. For municipalities and wastewater treatment plants, the US Clean Water State Revolving
Fund (CWSRF) loan program has been offering loans to help fund green projects. Through 2014, the
CWSRF program has provided 34 902 project assistance agreements ($105.4 billion) to communities.
At the middle level, preventive and containment measures provide better control of P release
into the environment with adopting technologies to prevent P runoff, promoting strategies to enable
downstream P removal and recovery, creating key partnerships and/or bridging institutions, and
building capacity for stakeholders. Technical and financial assistance to increase the capacity of
stakeholders, together with environmental education and outreach activities to raise public awareness
can encourage actions to reduce P losses into the environment. For example, the US Department of
Agriculture provides technical and financial support to farmers implementing P best management
practices to minimize P loss at the source, to mitigate the transport of P and problems associated
with excess P in water (Sharpley et al., 2006). For large regional issues, creating partnerships and/
or building bridging institutions provide a better platform for coordination among stakeholders. For
example, the Chesapeake Bay program as a bridging institution brought together stakeholders from
federal, state, local, academic, and nongovernmental organizations to build and adopt policies that
support the Bay’s restoration (Jones & Tippie, 1983). As part of the same program, states within the
Chesapeake Bay drainage area agreed to incorporate Chesapeake Bay issues into school curriculums
as part of the Chesapeake 2000 agreement.
The lowest level of P management is to provide P removal and recovery technologies. These
management strategies include providing technical and financial assistance for research and adopting
improvements in P removal and recovery. Technical and financial assistance is important for building
the knowledge and skills required to begin removing and recovering P from the waste streams.
To control P release from wastewater treatment plants into the environment, many countries
employed standards to regulate the concentration limit of P in treated municipal wastewater. The
main legal act regulating the quality of treated municipal wastewater in EU countries is the Council
Directive of May 21, 1991 (or so-called Wastewater Directive) (Council of European 1991). In the US,
the National Pollution Discharge Elimination System permit program created in the Clean Water Act
is the main regulation for treated wastewater discharged. In general, two major factors determining
the limit concentration of phosphorus in treated wastewater effluent are the size of the wastewater
treatment plant and the sensitivity to eutrophication of the water bodies receiving treated wastewater.
The limit concentration of total P ranges from 0.5 to 2 mg/L for EU members, 0.1–1 mg/L for states
in the US, 0.5 mg/L for China (Taihu Lake catchment), and 1 mg/L for Canada (Preisner et al., 2020).
In Europe, efforts to recover P from waste streams based on the circular economy paradigm have
been promoted in the European Green Deal and Circular Economy Action Plan (Bianchini & Rossi,
2020). Projects funded by the European Institute of Innovation and Technology Raw Materials and
Climate KIC provided recommendations and practical actions for a more sustainable P management
in Baltic Sea countries (Bianchini & Rossi, 2020). For example, in case of manure P which is one
of the major sources of P release into the environment, removal and recovery technologies can be
categorized into: (1) solid–liquid separation technologies; (2) technologies for processing the solid
fraction; and (3) technologies for processing the liquid section. A list of selected technologies for P
removal and recovery from animal manure is presented in section 8.3. More than 50 P removal and
recovery technical approaches are known and have been developed for municipal wastewater (Egle
et al., 2016). Based on the access points in wastewater treatment plants, P recovery can be conducted
via: (1) direct usage of sludge as soil amendments; (2) recovery from the aqueous phase either before or
after sludge dewatering process; (3) recovery from sludge during or after incineration (Egle et al., 2016).
A list of selected technologies for P recovery from municipal wastewater is presented in section 8.3.
In general, existing P management approaches were developed and conducted to reduce the
negative impacts of P losses into the watershed, which can accelerate freshwater eutrophication
and cause water-quality impairment (Litke, 1999; Sharpley & Tunney, 2000). The singular focus on
developing technologies for attaining near to complete P removal from waste streams would make
it more costly and un-sustainable when overall environmental impacts such as the carbon footprint
is overlooked. In the next section, we present the existing and emerging P management practices for
water quality improvements in conjunction with possible opportunities for reducing carbon footprint
to allow decision makers to select the most sustainable practices.
Table 8.2 P-rich waste stream, current approaches and opportunities for decarbonization.
Table 8.3 Strategies and technologies to recover resources and reduce the emissions of GHG in each step of the
manure management chain.
However, soil P surplus and potential pollution of water resources are common consequences of land
application of manure in regions with high concentrations of concentrated animal feeding operations
(CAFO). Manure transport for distances above ten miles is not economical and usually not practiced.
Another common manure management system includes uncovered anaerobic lagoons, which have
been identified by the US Environmental Protection Agency as the largest source of methane from
farms (Owen & Silver, 2015). These examples clearly indicate the need for management alternatives
to resolve agronomic P imbalances for more effective recycling of manure P, and simultaneous carbon
footprint reduction. Table 8.3 presents strategies and technologies to recover P and reduce the emission
of GHG in the management of manure in addition to land application.
Figure 8.3 Strategies and technologies to recover P resources and reduce the emissions of GHG in each step of the
domestic wastewater management chain.
Figure 8.4 Comparison of conventional and source separation systems. BW, GW, and FW refer to blackwater,
graywater and food waste, respectively. The figure is from Kjerstadius et al. (2017).
process were compared with the integrated processes that treat urine and wastewater separately,
and recover P via struvite precipitation. Their findings indicated that, advanced BNR processes
significantly increases energy and chemical consumption, requiring around 6 W per person, however,
the integrated treatment/recovery process could produce more than 1 W per person.
Numerous LCA studies have been conducted to compare the carbon footprint/environmental
impact of different scenarios for removal/recovery/recycling of phosphorus in human excreta or
removing them in an advanced WWTP (Bradford-Hartke et al., 2015; Hilton et al., 2020; Kavvada
et al., 2017; Spångberg et al., 2014; Xue et al., 2016), and all of the studies have indicated that urine
diversion reduced most environmental impacts through a wide range of conditions (reduced flushing,
reduce chemical use at BNR plants, better energy recovery from AD, etc.). It can be a particularly
effective decarbonization strategy in areas with high levels of nutrient removal, electricity produced
primarily from fossil fuels, and relatively little wastewater per capita.
Kjerstadius et al. (2017) also conducted an LCA study, but evaluated the decarbonization potential
of source separation systems for the management of domestic wastewater and food waste. In the
conventional system, blackwater (BW) and graywater (GW) are collected together and treated at a
wastewater treatment plant whereas separated food waste is collected by garbage trucks and treated
at a dedicated anaerobic digestion plant (see Figure 8.4). In the source separation system considered
by the study, BW was collected and then treated together with FW in an anaerobic digestion unit,
and nutrient recovery was performed on the digestate effluent by struvite precipitation and ammonia
stripping (to produce ammonium sulfate). GW was treated separately in an activated sludge unit.
The results for carbon footprint and nutrient recovery (phosphorus and nitrogen) concluded that the
source separation system could increase nutrient recovery (0.30–0.38 kg P per capita per year), while
decreasing the carbon footprint ( − 24 to − 58 kg CO2-eq. capita−1 year−1), compared to the conventional
system. The carbon footprint decreased, mainly due to energy recovery from the increased biogas
production, increased replacement of mineral fertilizer in agriculture and less emissions of nitrous
oxide from wastewater treatment.
Numerous opportunities for direct and indirect C reduction associated with P removal and recovery
processes at wastewater treatment plants exists and are evaluated in depth in section 8.3.
8.2.4 Phosphorus in urban runoff and best management practices for decarbonization
Residential lawns and turf areas (e.g., sports fields, golf courses and parks) in urban environments are
considered as ‘hotspots’ of total and dissolved phosphorus input into stormwater (Müller et al., 2020).
Moreover, fallen leaves and other detritus are often considered the primary contributors of nutrients
to urban stormwater, especially in areas with high overhead tree canopies (George et al., 2012).
A wide variety of stormwater control measures, also known as best management practices (BMPs),
are available that can play a significant role in the treatment of urban runoff, such as infiltration
beds (grass swales and porous pavements), filtration systems (sand filters, vegetated filter strips,
etc.), retention/detention basins (dry ponds, wet ponds and inline storage) and constructed wetlands
(Sample et al., 2012).
These strategies are used in conjunction with other measures to reduce the quantity of urban
runoff and reduce the impact of urban run-off on water quality (e.g., green roof, stormwater drain,
etc). All these strategies are part of large program and initiatives such as the ‘Sponge Cities’ in China
(Zevenbergen et al., 2018), ‘Water Sensitive Cities’ in Australia (Wong & Brown, 2009), ‘Sustainable
Urban Drainage Systems’ in the UK (Ashley et al., 2015), ‘Low Impact Development’ in the US
(Coffman, 2000) and New Zealand (Shaver, 2000).
Decarbonization opportunities with the implementation of these BMPs and strategies include
reduction in municipal pumping demand and energy costs associated with the added flood mitigation;
reduction of heat island effect, which results in heating and cooling energy savings; and carbon storage
in vegetation or trees in some of the BMPs.
Table 8.4 Summary of advantages and challenges of different decarbonization methods in P removal and
recovery processes.
(Continued)
Table 8.4 Summary of advantages and challenges of different decarbonization methods in P removal and
recovery processes (Continued).
A variety of carbon sources have been applied in EBPR systems for achieving high P removal
efficiency and stability (Table 8.5) which increases the overall treatment cost. The main drawbacks of
adding external carbon source in EBPR systems include:
1. The increase in the overall carbon dioxide emissions for the production and transport of added
commercial carbon sources leading to uneconomical and unsustainable operational practices.
2. Safety issues associated with the transport, handling, and storage of the external carbon
sources.
Table 8.5 Summary of advantages and limitations of different carbon sources for EBPR.
3. Long adaptation periods required in the startup process to acclimate the bacterial/PAO
community for preferential utilization specific carbon source.
4. The increased sludge production rate and the operational cost of water treatment and sludge
processing as a result of addition of external organic carbon.
Alternatively, the practice of on-site primary sludge fermentation is widely established in many
WWTPs, which reduces the overall carbon input in EBPR process. However, the carbon supply from
primary sludge fermentation is usually not adequate to ensure efficient P removal, as the VFA production
is often affected by several environmental and operational factors, such as influent properties, process
configuration, SRT, HRT, pH, temperature, and so on. Other drawbacks of implementing on-site
primary sludge fermentation in EBPR systems include:
• infeasible particularly for facilities with no primary treatment unit;
• increased footprint from additional construction and operation costs;
• potential odors from fermenter;
• reduced energy recovery via anaerobic digestion;
• potential effects of recalcitrant organic compounds and nutrients derived from fermentation
step on EBPR.
8.3.4 Additional technologies for phosphorus removal and recovery from wastewater streams
with carbon footprint reduction potential
8.3.4.1 Phosphorus recovery technologies at WWTP
As shown in Figure 8.3, there are various opportunities for P recovery from different waste streams at
wastewater treatment plants, such as: secondary treated effluent, digester supernatant, sewage sludge
(SS) and sewage sludge ash (SSA) (Montag & Pinnekamp, 2008). These streams differ widely in terms
of volume, P concentration, the form of P (dissolved as orthophosphate or biologically/chemically
bound), the characteristic of the source (liquid, liquid/solid, solid), pollutant content, requiring
different recovery technologies and therefore different decarbonization resorts.
More than 30 processes to recover phosphorus from waste streams have been identified. The
available solutions mainly focus on: struvite (crystals of magnesium ammonium phosphate)
precipitation from liquid fraction from different steps of wastewater treatment and they include:
(1) sludge after digestion; (2) wet chemical P recovery through an acid attack of ash leaching
phosphates; (3) thermal solubilization of phosphates in ash with simultaneous reduction of
heavy metals; and (4) use of ash for fertilizer manufacturing (Smol et al., 2020). Several life cycle
assessments case studies have identified opportunities and burdens associated with the advanced P
recovery from the different streams in wastewaters treatment plants. Amann et al. (2018) compared
the environmental impacts and GHG emissions of 18 P recovery technologies, described in Egle
et al. (2015). Amann et al. (2018) concluded that the recovery from the liquid phase generates less
emissions and has lower energy demands, but offers low rates of recovery, while recovery from
sludge (solid phase) has relatively higher emissions and higher energy demands. The recovery of
phosphorus from sludge ash, on the other hand, is the most promising option. It presents a higher
recycling rate, the possibility of heavy metal decontamination, and reduction of gaseous emissions
and energy demand (see Figure 8.5 extracted from Amann et al. 2018). According to Bradford-
Hartke et al. (2015), phosphorus recovery in an advanced BNR centralized water reclamation
facility led to a net reduction of 5 kgCO2eq/kgP. The net carbon footprint reduction is due to
avoided N2O emissions, lower power consumption, and reduced chemical usage for pH control (due
to reduced nitrification), which offset power and chemicals demands of dewatering liquid required
for struvite production.
Pradel and Aissani (2019) compared the environmental impacts of sludge-based phosphate
fertilizer production to producing mineral fertilizers from phosphate rocks. Their results indicated
Figure 8.5 Changes in carbon footprints of various P removal technologies. The figure is from previously published
study (Amann et al., 2018).
that sludge-based phosphate fertilizers appeared to have higher environmental impacts than mineral
phosphate fertilizers production, mainly due to their consumption of large amounts of electricity
and reactants needed to recover phosphorus, and their low phosphorus content in comparison with
phosphate rocks. Qualitatively similar conclusions were reported by Golroudbary et al. (2019). On the
contrary, Tonini et al. (2019) suggested that the environmental impacts of recovering sewage derived
P may be up to 81% less than mining P containing rock. They attribute the inconsistency of their
results with previous studies to the different assumptions, and that they have included the external
costs associated with all relevant emissions (including dissipated phosphate).
Other studies have indicated benefits from P recovery which are generally overlooked, for example,
the reduction in eutrophication potential (reduced phosphate rock mining and therefore lower P
water release from mining (Remy & Jossa, 2015), mitigation of Cd and U input into agricultural soils
(Bigalke et al., 2017), reduction of heavy metal input compared to conventional agricultural sewage
sludge application (Lederer & Rechberger, 2010), and decreased nitrogen emissions for technologies
which also recover nitrogen (Johansson et al., 2008). Overall, in spite of all inconsistences, these
results suggest that not all P recovery technologies offer decarbonization potential. Thorough, holistic
assessments of phosphorus recovery technologies are required.
8.3.4.3 Microalgae cultivation for joint nutrient removal and energy production
The incorporation of microalgae cultivation is a cost-effective and sustainable measure in WWTPs,
as microalgae can fix exogenous CO2 during autotrophic growth while assimilating N, P and metal in
wastewater. Harvested lipid-rich microalgae could be used for generation of biofuels (e.g., biodiesel)
with the potential to reduce greenhouse gas emissions through replacement of fossil fuels. The other
uses of algae biomass include carbon- and nutrient-rich soil amendment, animal feed, and bioplastic
production. Microalgae have been widely studied for CO2 capture and utilization, and extensive
research has been carried out on their use in large-scale (>5000 acres) cultivation systems.
genetically modified to overexpress phosphate-binding proteins (PBPs, also known as PstS or PhoS),
resulting in a highly improved P removal and recovery performance. Implementation of recombinant-
plasmid bacteria systems for selective P adsorption in actual wastewater treatment applications is
a challenging but attractive approach with relatively less chemical/energy input for configuration,
modification and maintenance compared to other processes/technologies, therefore deserving further
exploration.
8.4.1 LCA studies for the quantification of decarbonization potential for non-point sources
A broad understanding of the environmental impacts of the life cycle of non-point source P management
strategies is missing. There have been P-oriented LCA studies addressing eutrophication potential
from agricultural sources (Ortiz-Reyes & Anex, 2018). Those studies have focused on estimating the P
transport and discharge from agricultural sources and overlooked the indirect impacts of control and
mitigation strategies and their associated carbon footprint.
Few studies have provided critical reviews on non-point source management practices in relation
to practicality, cost-effectiveness, and regulatory requirements (Dinnes, 2004; Macintosh et al., 2018).
It has been suggested that source-oriented practices offer a cheaper and more effective solution than
endpoint alternatives such as wetlands. No study has compared the sustainability of containment
practices at the source with in-sink treatment practices or with combinational approaches. Wetlands
are considered low-tech with low energy demands solution. More than 80% of the environmental
impacts of the life cycle of wetlands are from the construction phase (Resende et al., 2019). Riparian
buffers can retain as high as 97% of P from run-off and eroded soil (Fox & Penn, 2013) and at the same
time offer a net saving (11.9 Mg CO2eq ha−1 year−1) in global warming potential (Styles et al., 2016).
Due to the diverse range and diffused nature of non-point sources, in-field management practices
may not be universally effective, cost-effective, or the optimum solution in relation to environmental
consequences. Also, there are often continuous P releases into the environment from accumulated
legacy P content of the soil (Powers et al., 2016), even when effective source containment strategies
are in place, necessitating combinational strategies at the source and in the sink for protecting
watersheds. The selection of a proper integrated management strategy requires a holistic picture of
the environmental net-benefit of various alternatives. Therefore, LCA is a key tool for a sustainable
P management of non-point sources in the water sector. Environmental release and transports of P
varies from site to site, and in the same site varies with time scales. Thus, future LCA studies should,
beyond common practices, consider such site-specific temporal variations.
8.4.2 LCA studies for P removal and recovery processes in WWTPs and quantification of
decarbonization potential
To maximize the potential of wastewater resources, a robust and integrative approach is needed to
quantitatively compare the environmental attributes of diverse technology options for P removal and
recovery technologies. In recent years few publications have discussed the environmental impacts of
nutrient removal technologies (Coats et al., 2011; Foley et al., 2010; Rahman et al., 2016), or P recovery
technologies (Amann et al., 2018; Bradford-Hartke et al., 2015), but little research has been published
on such a comparative sustainability assessment of both recovery and removal of P from wastewater.
An assessment of 27 nutrient removal technologies was carried out by Rahman et al. (2016); life
cycle impact assessment (LCIA) of the representative treatment process configurations with different
levels of treatment for both nitrogen and phosphorus was performed. Results showed that advanced
technologies that achieve high-level nutrient removal significantly decreased local eutrophication
potential, while chemicals and electricity use for these advanced treatments, particularly multistage
enhanced tertiary processes and reverse osmosis, increased indirect eutrophication potential and
contributed to other impacts including human and ecotoxicity, global warming potential, ozone
depletion, and acidification.
Regardless of the required effluent limit, when biological phosphorus removal processes are
compared to chemical processes in terms of environmental impacts, it would appear that best practices
would center wastewater treatment first on the biological process (Coats et al., 2011). The EBPR
process also produces significantly fewer biosolids and no chemical sludge, which allows for further
reduction of carbon footprint due to avoided transport/handling of these byproducts.
The LCA studies published that focused on P recovery schemes and technologies from urine and
various wastewater streams have highlighted the potential decarbonization associated with these
technologies, as already discussed in previous sections, but also identified some inconsistencies. The
typical approach in the published studies is the comparison of environmental impacts (often not
consistent among studies) in different scenarios involving different technologies and case-specific
waste streams. Consequently, quantitative results about environmental impact are strictly related to
the specific application and should not be used as a basis for deriving general conclusions about the
implications of P recovery. Further in-depth, comprehensive analyses about P recovery systems are
necessary to know their broad environmental impacts.
context, P trading policies should be developed. Implementing such policies should be contingent
on holistic life cycle analyses of net environmental impacts associated with each intervention. This
consideration would allow decision makers to select the most appropriate practices that meet both
water and air quality goals.
REFERENCES
Alewell C., Ringeval B., Ballabio C., Robinson D. A., Panagos P. and Borrelli P. (2020). Global phosphorus
shortage will be aggravated by soil erosion. Nature Communications, 11(1), 4546, https://doi.org/10.1038/
s41467-020-18326-7
Altinbas M., Ozturk I. and Aydin A. F. (2002). Ammonia recovery from high strength agro industry effluents.
Water Science and Technology, 45(12), 189–195, https://doi.org/10.2166/wst.2002.0426
Amann A., Zoboli O., Krampe J., Rechberger H., Zessner M. and Egle L. (2018). Environmental impacts of
phosphorus recovery from municipal wastewater. Resources, Conservation and Recycling, 130, 127–139,
https://doi.org/10.1016/j.resconrec.2017.11.002
Anderson D. M., Glibert P. M. and Burkholder J. M. (2002). Harmful algal blooms and eutrophication: nutrient
sources, composition, and consequences. Estuaries, 25(4), 704–726, https://doi.org/10.1007/BF02804901
Ashley R., Walker L., D’Arcy B., Wilson S., Illman S., Shaffer P., Woods-Ballard B. and Chatfield P. (2015). UK
Sustainable drainage systems: past, present and future. Proceedings of the Institution of Civil Engineers –
Civil Engineering, 168(3), 125–130, https://doi.org/10.1680/cien.15.00011
Barnard J. L., Dunlap P. and Steichen M. (2017). Rethinking the mechanisms of biological phosphorus removal.
Water Environment Research, 89(11), 2043–2054, https://doi.org/10.2175/106143017X15051465919010
Bianchini A. and Rossi J. (2020). An integrated industry-based methodology to unlock full-scale implementation
of phosphorus recovery technology. Sustainability, 12(24), 10632, https://doi.org/10.3390/su122410632
Bigalke M., Ulrich A., Rehmus A. and Keller A. (2017). Accumulation of cadmium and uranium in arable soils in
Switzerland. Environmental Pollution, 221, 85–93, https://doi.org/10.1016/j.envpol.2016.11.035
Bouwman L., Goldewijk K. K., Hoek K. W. V. D., Beusen A. H. W., Vuuren D. P. V., Willems J., Rufino M. C. and
Stehfest E. (2013). Exploring global changes in nitrogen and phosphorus cycles in agriculture induced by
livestock production over the 1900–2050 period. Proceedings of the National Academy of Sciences, 110(52),
20882–20887, https://doi.org/10.1073/pnas.1012878108
Bradford-Hartke Z., Lane J., Lant P. and Leslie G. (2015). Environmental benefits and burdens of phosphorus
recovery from municipal wastewater. Environmental Science and Technology, 49(14), 8611–8622, https://
doi.org/10.1021/es505102v
Bunce J. T., Ndam E., Ofiteru I. D., Moore A. and Graham D. W. (2018). A review of phosphorus removal
technologies and their applicability to small-scale domestic wastewater treatment systems. Frontiers in
Environmental Science, 6, 8, https://doi.org/10.3389/fenvs.2018.00008
Campolong C., Ferguson L., Klaus S., Wilson C., Wett B., Murthy S. and Bott C. B. (2018). Achieving
mainstream deammonification with biological phosphorus removal via sidestream RAS fermentation in
an A/B process. Proceedings of the Water Environment Federation, 2018(13), 2634–2639, https://doi.
org/10.2175/193864718825137142
Cao Y., Tang J., Henze M., Yang X., Gan Y., Li J., Kroiss H., Van Loosdrecht M., Zhang Y. and Daigger G. (2019).
The leakage of sewer systems and the impact on the ‘black and odorous water bodies’ and WWTPs in China.
Water Science and Technology, 79(2), 334–341, https://doi.org/10.2166/wst.2019.051
Carvalheira M., Oehmen A., Carvalho G., Eusébio M. and Reis M. A. M. (2014). The impact of aeration on the
competition between polyphosphate accumulating organisms and glycogen accumulating organisms. Water
Research, 66, 296–307, https://doi.org/10.1016/j.watres.2014.08.033
Chong J., Dominish E., Tjandraarmadja G., Prentice E., Retamal M. and Mitrovic S. (2019). Ecosystem Impacts
of Phosphorus and Surfactants in Consumer Products. Institute for Sustainable Futures. Report prepared by
the Institute for Sustainable Futures, University of Technology Sydney for Stewart Investors.
Coats E. R., Watkins D. L. and Kranenburg D. (2011). A comparative environmental life-cycle analysis for
removing phosphorus from wastewater: biological versus physical/chemical processes. Water Environment
Research, 83(8), 750–760, https://doi.org/10.2175/106143011X12928814444619
Coffman L. S. (2000). Low-impact Development Design: A New Paradigm for Stormwater Management Mimicking
and Restoring the Natural Hydrologic Regime, an Alternative Stormwater Management Technology. Prince
George’s County Department of Environmental Resources and US EPA (Report EPA 841-B-00-003), Maryland
County, USA. Retrieved from: http://www.epa.gov/owow/NPS/lidnatl.pdf (accessed 23 January 2021)
Cohen A. and Keiser D. A. (2017). The effectiveness of incomplete and overlapping pollution regulation: evidence
from bans on phosphate in automatic dishwasher detergent. Journal of Public Economics, 150, 53–74,
https://doi.org/10.1016/j.jpubeco.2017.03.005
Cordell D., Drangert J.-O. and White S. (2009). The story of phosphorus: global food security and food for thought.
Global Environmental Change, 19(2), 292–305, https://doi.org/10.1016/j.gloenvcha.2008.10.009
Cordell D., Rosemarin A., Schröder J. J. and Smit A. (2011). Towards global phosphorus security: a systems
framework for phosphorus recovery and reuse options. Chemosphere, 84(6), 747–758, https://doi.
org/10.1016/j.chemosphere.2011.02.032
Dinnes D. L. (2004). Assessments of Practices to Reduce Nitrogen and Phosphorus Nonpoint Source Pollution of
Iowa’s Surface Waters. USDA-ARS, National Soil Tilth Laboratory, Ames, IA, p. 366.
Dodds W. K., Bouska W. W., Eitzmann J. L., Pilger T. J., Pitts K. L., Riley A. J., Schloesser J. T. and Thornbrugh D. J.
(2009). Eutrophication of U.S. Freshwaters: analysis of potential economic damages. Environmental Science
and Technology, 43(1), 12–19, https://doi.org/10.1021/es801217q
Dubrovsky N., Burow K., Clark G. M., Gronberg J. M., Hamilton P. A., Hitt K. J., Mueller D. K., Munn M., Nolan
B. T., Puckett L. J., Rupert M. G., Short T. M., Spahr N. E., Sprague L. A. and Wilber W. G. (2010). The quality
of our nation’s waters – nutrients in the nation’s streams and groundwater, 1992–2004. US Geological Survey
Circular, 1350(2), 174.
Egle L., Rechberger H. and Zessner M. (2015). Overview and description of technologies for recovering
phosphorus from municipal wastewater. Resources, Conservation and Recycling, 105, 325–346, https://doi.
org/10.1016/j.resconrec.2015.09.016
Egle L., Rechberger H., Krampe J. and Zessner M. (2016). Phosphorus recovery from municipal wastewater: an
integrated comparative technological, environmental and economic assessment of P recovery technologies.
Science of the Total Environment, 571, 522–542, https://doi.org/10.1016/j.scitotenv.2016.07.019
European Commission. (1962). The Common Agricultural Policy at A Glance. European Commission, Brussels.
Available at: https://ec.europa.eu/info/food-farming-fisheries/key-policies/common-agricultural-policy/
cap-glance_en (accessed 23 January 2021)
European Commission. (2011). EP Supports Ban of Phosphates in Consumer Detergents. European Commission,
Brussels. Available at: https://ec.europa.eu/commission/presscorner/detail/en/IP_11_1542
Foley J., Haas D. D., Hartley K. and Lant P. (2010). Comprehensive life-cycle inventories of alternative wastewater
treatment systems. Water Research, 44(5), 1654–1666.
Fox G. A. and Penn C. J. (2013). Empirical model for quantifying total phosphorus reduction by vegetative filter
strips. Transactions of the ASABE, 56(4), 1461–1469.
Gao H., Liu M., Griffin J. S., Xu L., Xiang D., Scherson Y. D., Liu W.-T. and Wells G. F. (2017). Complete nutrient
removal coupled to nitrous oxide production as a bioenergy source by denitrifying polyphosphate-accumulating
organisms. Environmental Science and Technology, 51(8), 4531–4540, https://doi.org/10.1021/acs.est.6b04896
George H., Terril N., Unruh J. B., Laurie T. and Jerry S. (2012). Potential unintended consequences associated
with urban fertilizer bans in Florida – a scientific review. HortTechnology Hortte, 22(5), 600–616, https://
doi.org/10.21273/HORTTECH.22.5.600
Golroudbary S. R., Wali M. E. and Kraslawski A. (2019). Environmental sustainability of phosphorus recycling
from wastewater, manure and solid wastes. Science of the Total Environment, 672, 515–524, https://doi.
org/10.1016/j.scitotenv.2019.03.439
Gu A., Tooker N., Onnis-Hayden A., Wang D., Srinivasan V., Li G., Takács I. and Vargas E. (2019). Optimization
and Design of A Side-Stream EBPR Process as A Sustainable Approach for Achieving Stable and Efficient
Phosphorus Removal. The Water Research Foundation, DC, USA.
Hilton S. P., Keoleian G. A., Daigger G. T., Zhou B. and Love N. G. (2020). Life-cycle assessment of urine diversion
and conversion to fertilizer products at the city scale. Environmental Science and Technology, 55(1), 593–
603, https://doi.org/10.1021/acs.est.0c04195
Jasinski S. M. (2020). Phosphate Rock Data Sheet – Mineral Commodity Summaries 2020. US Geological Survey,
Reston, Virginia.
Johansson K., Perzon M., Fröling M., Mossakowska A. and Svanström M. (2008). Sewage sludge handling with
phosphorus utilization–life-cycle assessment of four alternatives. Journal of Cleaner Production, 16(1), 135–
151, https://doi.org/10.1016/j.jclepro.2006.12.004
Jones G. A. and Tippie V. K. (1983). Chesapeake Bay Program: Findings and Recommendations. US Environmental
Protection Agency, Philadelphia.
Kavvada O., Tarpeh W. A., Horvath A. and Nelson K. L. (2017). Life-cycle cost and environmental assessment of
decentralized nitrogen recovery using ion exchange from source-separated urine through spatial modeling.
Environmental Science and Technology, 51(21), 12061–12071, https://doi.org/10.1021/acs.est.7b02244
Kishida N., Tsuneda S., Kim J. H. and Sudo R. (2009). Simultaneous nitrogen and phosphorus removal from high-
strength industrial wastewater using aerobic granular sludge. Journal of Environmental Engineering, 135(3),
153–158, https://doi.org/10.1061/(ASCE)0733-9372(2009)135:3(153)
Kjerstadius H., Bernstad Saraiva A., Spångberg J. and Davidsson Å. (2017). Carbon footprint of urban source
separation for nutrient recovery. Journal of Environmental Management, 197, 250–257, https://doi.
org/10.1016/j.jenvman.2017.03.094
Kuba T., Van Loosdrecht M. and Heijnen J. (1996). Phosphorus and nitrogen removal with minimal COD
requirement by integration of denitrifying dephosphatation and nitrification in a two-sludge system. Water
Research, 30(7), 1702–1710, https://doi.org/10.1016/0043-1354(96)00050-4
Kundu S., Bhattacharyya R., Prakash V., Ghosh B. N. and Gupta H. S. (2007). Carbon sequestration and
relationship between carbon addition and storage under rainfed soybean–wheat rotation in a sandy loam soil
of the Indian Himalayas. Soil and Tillage Research, 92(1), 87–95, https://doi.org/10.1016/j.still.2006.01.009
Lal R. (2004). Soil carbon sequestration impacts on global climate change and food security. Science, 304(5677),
1623–1627, https://doi.org/10.1126/science.1097396
Lal R. and Kimble J. M. (1997). Conservation tillage for carbon sequestration. Nutrient Cycling in Agroecosystems,
49(1), 243–253, https://doi.org/10.1023/A:1009794514742
Lederer J. and Rechberger H. (2010). Comparative goal-oriented assessment of conventional and alternative
sewage sludge treatment options. Waste Management, 30(6), 1043–1056, https://doi.org/10.1016/j.
wasman.2010.02.025
Litke D. (1999). Review of Phosphorus Control Measures in the United States and Their Effects on Water Quality.
US Geological Survey, Reston, Virginia.
Liu Y., Villalba G., Ayres R. U. and Schroder H. (2008). Global phosphorus flows and environmental
impacts from a consumption perspective. Journal of Industrial Ecology, 12(2), 229–247, https://doi.
org/10.1111/j.1530-9290.2008.00025.x
Liu K., Elliott J. A., Lobb D. A., Flaten D. N. and Yarotski J. (2014). Conversion of conservation tillage to
rotational tillage to reduce phosphorus losses during snowmelt runoff in the Canadian prairies. Journal of
Environmental Quality, 43(5), 1679–1689, https://doi.org/10.2134/jeq2013.09.0365
Lundie S., Peters G. M. and Beavis P. C. (2004). Life-cycle assessment for sustainable metropolitan water systems
planning. Environmental Science and Technology, 38(13), 3465–3473, https://doi.org/10.1021/es034206m
Macintosh K. A., Mayer B. K., McDowell R. W., Powers S. M., Baker L. A., Boyer T. H. and Rittmann B. E. (2018).
Managing diffuse phosphorus at the source versus at the sink. Environmental Science and Technology,
52(21), 11995–12009, https://doi.org/10.1021/acs.est.8b01143
Maktabifard M., Zaborowska E. and Makinia J. (2018). Achieving energy neutrality in wastewater treatment
plants through energy savings and enhancing renewable energy production. Reviews in Environmental
Science and Bio/Technology, 17(4), 655–689, https://doi.org/10.1007/s11157-018-9478-x
Maurer M., Pronk W. and Larsen T. A. (2006). Treatment processes for source-separated urine. Water Research,
40(17), 3151–3166, https://doi.org/10.1016/j.watres.2006.07.012
McDowell R. W. (2015). Treatment of pasture topsoil with alum to decrease phosphorus losses in subsurface
drainage. Agriculture, Ecosystems and Environment, 207, 178–182, https://doi.org/10.1016/j.agee.2015.04.017
Mihelcic J. R., Fry L. M. and Shaw R. (2011). Global potential of phosphorus recovery from human urine and feces.
Chemosphere, 84(6), 832–839, https://doi.org/10.1016/j.chemosphere.2011.02.046
Montag D. M. and Pinnekamp J. (2008). Phosphorus Recovery During Wastewater Treatment: Development of a
Process for Integration in Municipal Wastewater Treatment Plants. PhD thesis, RWTH Aachen University,
Aachen, Germany.
Mulkerrins D., O’Connor E., Lawlee B., Barton P. and Dobson A. (2004). Assessing the feasibility of achieving
biological nutrient removal from wastewater at an Irish food processing factory. Bioresource Technology,
91(2), 207–214, https://doi.org/10.1016/S0960-8524(03)00173-1
Müller A., Österlund H., Marsalek J. and Viklander M. (2020). The pollution conveyed by urban runoff: a review
of sources. Science of the Total Environment, 709, 136125, https://doi.org/10.1016/j.scitotenv.2019.136125
Nsenga Kumwimba M., Meng F., Iseyemi O., Moore M. T., Zhu B., Tao W., Liang T. J. and Ilunga L. (2018).
Removal of non-point source pollutants from domestic sewage and agricultural runoff by vegetated drainage
ditches (VDDs): design, mechanism, management strategies, and future directions. Science of the Total
Environment, 639, 742–759, https://doi.org/10.1016/j.scitotenv.2018.05.184
Naidu R., Lamb D. T., Bolan N. S. and Gawandar J.. (2012). Recovery and reuse of phosphorus from wastewater
sources. Advanced nutrient management: gains from the past—goals for the future. Occasional Report 25.
See http://flrc.massey.ac.nz/publications.html.
Oehmen A., Lemos P. C., Carvalho G., Yuan Z., Keller J., Blackall L. L. and Reis M. A. (2007). Advances in
enhanced biological phosphorus removal: from micro to macro scale. Water Research, 41(11), 2271–2300,
https://doi.org/10.1016/j.watres.2007.02.030
Onnis-Hayden A., Srinivasan V., Tooker N. B., Li G., Wang D., Barnard J. L., Bott C., Dombrowski P., Schauer P.
and Menniti A. (2020). Survey of full-scale sidestream enhanced biological phosphorus removal (S2EBPR)
systems and comparison with conventional EBPRs in North America: process stability, kinetics, and
microbial populations. Water Environment Research, 92(3), 403–417, https://doi.org/10.1002/wer.1198
Ortiz-Reyes E. and Anex R. P. (2018). A life-cycle impact assessment method for freshwater eutrophication due
to the transport of phosphorus from agricultural production. Journal of Cleaner Production, 177, 474–482,
https://doi.org/10.1016/j.jclepro.2017.12.255
Owen J. J. and Silver W. L. (2015). Greenhouse gas emissions from dairy manure management: a review of field-
based studies. Global Change Biology, 21(2), 550–565, https://doi.org/10.1111/gcb.12687
Powers S. M., Bruulsema T. W., Burt T. P., Chan N. I., Elser J. J., Haygarth P. M., Howden N. J. K., Jarvie H. P., Lyu
Y., Peterson H. M., Sharpley A. N., Shen J., Worrall F. and Zhang F. (2016). Long-term accumulation and
transport of anthropogenic phosphorus in three river basins. Nature Geoscience, 9(5), 353–356, https://doi.
org/10.1038/ngeo2693
Pradel M. and Aissani L. (2019). Environmental impacts of phosphorus recovery from a ‘product’ life-
cycle assessment perspective: allocating burdens of wastewater treatment in the production of sludge-
based phosphate fertilizers. Science of the Total Environment, 656, 55–69, https://doi.org/10.1016/j.
scitotenv.2018.11.356
Preisner M., Neverova-Dziopak E. and Kowalewski Z. (2020). An analytical review of different approaches to
wastewater discharge standards with particular emphasis on nutrients. Environmental Management, 66(4),
694–708, https://doi.org/10.1007/s00267-020-01344-y
Rahman S. M., Eckelman M. J., Onnis-Hayden A. and Gu A. Z. (2016). Life-cycle assessment of advanced nutrient
removal technologies for wastewater treatment. Environmental Science and Technology, 50(6), 3020–3030,
https://doi.org/10.1021/acs.est.5b05070
Remy C. and Jossa P. (2015). Life-cycle Assessment of Selected Processes for P Recovery From Sewage Sludge,
Sludge Liquor, or Ash. P-Rex Deliverable 9 Report. Kompetenzzentrum Wasser, Berlin.
Renzoni R. and Germain A. (2007). Life-cycle assessment of water: from the pumping station to the wastewater
treatment plant. The International Journal of Life-Cycle Assessment, 12(2), 118–126, https://doi.org/10.1065/
lca2005.12.243
Resende J. D., Nolasco M. A. and Pacca S. A. (2019). Life-cycle assessment and costing of wastewater treatment
systems coupled to constructed wetlands. Resources, Conservation and Recycling, 148, 170–177, https://doi.
org/10.1016/j.resconrec.2019.04.034
Rheinhardt R. D., Brinson M. M., Meyer G. F. and Miller K. H. (2012). Carbon storage of headwater
riparian zones in an agricultural landscape. Carbon Balance and Management, 7(1), 41–54, https://doi.
org/10.1186/1750-0680-7-4
Rittmann B. E., Mayer B., Westerhoff P. and Edwards M. (2011). Capturing the lost phosphorus. Chemosphere,
84(6), 846–853, https://doi.org/10.1016/j.chemosphere.2011.02.001
Roots P., Sabba F., Rosenthal A. F., Wang Y., Yuan Q., Rieger L., Yang F., Kozak J. A., Zhang H. and Wells G. F.
(2019). Integrated low-energy and low carbon shortcut nitrogen removal with biological phosphorus removal
for sustainable mainstream wastewater treatment. bioRxiv, 772004.
Rudrappa L., Purakayastha T. J., Singh D. and Bhadraray S. (2006). Long-term manuring and fertilization effects
on soil organic carbon pools in a typic haplustept of semi-arid sub-tropical India. Soil and Tillage Research,
88(1), 180–192, https://doi.org/10.1016/j.still.2005.05.008
Sample D. J., Grizzard T. J., Sansalone J., Davis A. P., Roseen R. M. and Walker J. (2012). Assessing performance
of manufactured treatment devices for the removal of phosphorus from urban stormwater. Journal of
Environmental Management, 113, 279–291, https://doi.org/10.1016/j.jenvman.2012.08.039
Sharpley A. and Tunney H. (2000). Phosphorus research strategies to meet agricultural and environmental
challenges of the 21st century. Journal of Environmental Quality, 29(1), 176–181, https://doi.org/10.2134/
jeq2000.00472425002900010022x
Sharpley A. N., Daniel T., Gibson G., Bundy L., Cabrera M., Sims T., Stevens R., Lemunyon J., Kleinman P.
and Parry R. (2006). Best Management Practices to Minimize Agricultural Phosphorus Impacts on Water
Quality. USDA Agricultural Research Service Publication, Washington, DC.
Shaver E. (2000). Low Impact Design Manual for the Auckland Region. Regional Council Technical Report
TR2008/045, Auckland, NZ.
Srinivasan V. N., Li G., Wang D., Tooker N. B., Dai Z., Onnis-Hayden A., Bott C., Dombrowski P., Schauer P.,
Pinto A. and Gu A. Z. (2021). Oligotyping and metagenomics reveal distinct Candidatus accumulibacter
communities in side-stream versus conventional full-scale enhanced biological phosphorus removal (EBPR)
systems. Water Research, 206, 117725, https://doi.org/10.1016/j.watres.2021.117725
Smith L. C. and McDowell R. W. (2016). The use of alum to decrease phosphorus loss from dairy farm laneways in
southern New Zealand. Soil Use and Management, 32(1), 69–71, https://doi.org/10.1111/sum.12252
Smol M., Preisner M., Bianchini A., Rossi J., Hermann L., Schaaf T., Kruopienė J., Pamakštys K., Klavins M. and
Ozola-Davidane R. (2020). Strategies for sustainable and circular management of phosphorus in the Baltic
Sea Region: the holistic approach of the InPhos Project. Sustainability, 12(6), 2567, https://doi.org/10.3390/
su12062567
Spångberg J., Tidåker P. and Jönsson H. (2014). Environmental impact of recycling nutrients in human excreta to
agriculture compared with enhanced wastewater treatment. Science of the Total Environment, 493, 209–
219, https://doi.org/10.1016/j.scitotenv.2014.05.123
Styles D., Börjesson P., D’Hertefeldt T., Birkhofer K., Dauber J., Adams P., Patil S., Pagella T., Pettersson L. B.,
Peck P., Vaneeckhaute C. and Rosenqvist H. (2016). Climate regulation, energy provisioning and water
purification: quantifying ecosystem service delivery of bioenergy willow grown on riparian buffer zones
using life-cycle assessment. AMBIO, 45(8), 872–884, https://doi.org/10.1007/s13280-016-0790-9
Tonini D., Saveyn H. G. and Huygens D. (2019). Environmental and health co-benefits for advanced phosphorus
recovery. Nature Sustainability, 2(11), 1051–1061, https://doi.org/10.1038/s41893-019-0416-x
Urban Waste Water Treatment Directive (UWWTD). (1991). Council of the European communities. Council
directive of 21 May 1991 concerning urban waste water treatment (91/271/EEC). Official Journal, L, 135–140.
Vaccari D. A., Powers S. M. and Liu X. (2019). Demand-driven model for global phosphate rock suggests paths
for phosphorus sustainability. Environmental Science and Technology, 53(17), 10417–10425, https://doi.
org/10.1021/acs.est.9b02464
Wang D., Tooker N. B., Srinivasan V., Li G., Fernandez L. A., Schauer P., Menniti A., Maher C., Bott C. B. and
Dombrowski P. (2019). Side-stream enhanced biological phosphorus removal (S2EBPR) process improves
system performance-A full-scale comparative study. Water Research, 167, 115109, https://doi.org/10.1016/j.
watres.2019.115109
Wilsenach J. and van Loosdrecht M. (2003). Impact of separate urine collection on wastewater treatment systems.
Water Science and Technology, 48(1), 103–110, https://doi.org/10.2166/wst.2003.0027
Wong T. H. and Brown R. R. (2009). The water sensitive city: principles for practice. Water Science and Technology,
60(3), 673–682, https://doi.org/10.2166/wst.2009.436
Xia Y., Zhang M., Tsang D. C. W., Geng N., Lu D., Zhu L., Igalavithana A. D., Dissanayake P. D., Rinklebe
J., Yang X. and Ok Y. S. (2020). Recent advances in control technologies for non-point source pollution
with nitrogen and phosphorous from agricultural runoff: current practices and future prospects. Applied
Biological Chemistry, 63(1), 8, https://doi.org/10.1186/s13765-020-0493-6
Xu T. and Flapper J. (2009). Energy use and implications for efficiency strategies in global fluid-milk processing
industry. Energy Policy, 37(12), 5334–5341, https://doi.org/10.1016/j.enpol.2009.07.056
Xue X., Hawkins T. R., Schoen M. E., Garland J. and Ashbolt N. J. (2016). Comparing the life-cycle energy
consumption, global warming and eutrophication potentials of several water and waste service options.
Water, 8(4), 154.
Yuan Z., Pratt S. and Batstone D. J. (2012). Phosphorus recovery from wastewater through microbial processes.
Current Opinion in Biotechnology, 23(6), 878–883, https://doi.org/10.1016/j.copbio.2012.08.001
Zevenbergen C., Fu D. and Pathirana A. (2018). Sponge Cities: Emerging Approaches, Challenges and
Opportunities. MDPI, Basel, Switzerland.
Chapter 9
Decarbonization potentials using
photobiological systems
Lara Méndez1,2, Cristian A. Sepúlveda-Muñoz1,2, María del Rosario Rodero1,2,
Ignacio de Godos1,2 and Raúl Muñoz1,2*
1Institute of Sustainable Processes, University of Valladolid, Dr. Mergelina s/n., Valladolid 47011, Spain
2Department of Chemical Engineering and Environmental Technology, University of Valladolid, Dr. Mergelina s/n.,
Valladolid 47011, Spain
*Correspondence: [email protected]
9.1 INTRODUCTION
With a current focus on the need to limit the effects of climate change, greenhouse gas emissions
(GHG) associated with the wastewater treatment (WWT) sector need to be reduced. In this context,
significant reductions can be attained by using phototrophic microorganisms as a platform for WWT
combined with anaerobic digestion and biogas upgrading for bioenergy production. Microorganisms
like microalgae or purple photosynthetic bacteria (PPB) are capable of supporting a cost-effective
WWT with a lower energy consumption (with the associated reduction in indirect CO2 emissions),
an enhanced nutrient recovery and an in-situ assimilation of the CO2 produced during pollutant
oxidation compared to conventional WWT technologies. Thus, PPB constitute a promising biological
platform for the treatment of high strength wastewaters based on their high growth rates, ability to
use infrared radiation and tolerance to high salinity and low temperatures. Similarly, algal-bacterial
symbiosis supports complex interactions that contribute to a sustainable assimilation of carbon and
nutrients from multiple types of wastewaters. The photosynthetic biomass generated during WWT
can be further valorized into added-value products, including biogas. In this regard, biogas upgrading
into biomethane can be integrated in photosynthetic wastewater treatment schemes based on the
ability of PPB and microalgae to capture and utilize CO2 and use H 2S as an electron donor, which can
further decarbonize WWT. Finally, the recent advances in photobioreactor design that have boosted
the biodegradation potential of photobiological-based systems, while lowering their energy demand,
will be critically discussed in this book chapter.
Greenhouse gas (GHG) emissions, mainly carbon dioxide (CO2) from the combustion of fossil
fuels, have increased since the beginning of the industrial revolution with a higher incidence since the
1950s. The accumulation of GHG in the atmosphere has caused harmful effects on the environment
and climate change. The European Commission has recently targeted a GHG emissions reduction of
at least 55% by 2030 in comparison to 1990 emissions. Therefore, there is an urgent need to limit all
GHG emissions and further decarbonize both the energy system and all industrial sectors by boosting
© 2022 The Editors. This is an Open Access eBook distributed under the terms of the Creative Commons Attribution Licence
(CC BY-NC-ND 4.0), which permits copying and redistribution for non-commercial purposes with no derivatives, provided the
original work is properly cited (https://creativecommons.org/licenses/by-nc-nd/4.0/). This does not affect the rights licensed or
assigned from any third party in this book. The chapter is from the book Pathways to Water Sector Decarbonization, Carbon Capture
and Utilization, Zhiyong Jason Ren and Krishna Pagilla (Eds.)
and implementing innovative and sustainable production and waste management technologies (Qiao
et al., 2020).
In the context of the wastewater treatment (WWT) sector, the implementation of new technologies
based on phototrophic microorganisms in combination with anaerobic digestion (AD) systems have
proven to be cost-efficient for WW bioremediation and energy production (Fouilland et al., 2014; Maity
et al., 2014; Qiao et al., 2020). These technologies do not require external aeration, which significantly
reduces the energy demand and CO2 footprint of WWT. Microalgae are capable of assimilating
multiple pollutants like nitrogen compounds and phosphates present in aqueous effluents or CO2 from
off-gases with the consequent generation of oxygen and biomass during the photosynthetic process.
The implementation of microalgal-based systems allows reducing direct and indirect CO2 emissions,
supports a cost-effective WWT with lower energy consumption, and enhances nutrient recovery
compared to conventional WWT technologies. Moreover, the recovered microalgae biomass can be
further valorized into other added value compounds such as biofertilizers or employed as biofuels
substrates, that is biodiesel, integrating this process into the biorefinery concept and circular economy.
Likewise, purple phototrophic bacteria (PPB) have emerged as a novel and promising platform for the
removal of pollutants, which can be operated as a standalone technology or in combination with other
technologies such as algal-bacterial systems. PPB-based systems can use organic matter and infrared
spectra from solar radiation as energy source. These microorganisms have been reported to tolerate
and grow in high strength waters with a high salinity or toxicity and support high rates of organic
matter and nutrient assimilation (Batstone et al., 2015; Hülsen et al., 2014).
These photosynthetic platforms can be integrated with AD. AD is a biological process driven by
symbiotic microbial communities in the absence of oxygen, in which biodegradable organic matter is
bioconverted into biogas and nutrients are mineralized in the digestate. Indeed, AD can cope with
multiple organic residues, which are used as feedstocks for biogas generation (Čater et al., 2015; González-
Fernández et al., 2008; Mendez et al., 2014a; Nkemka & Murto, 2013; Rani et al., 2012). Biogas is a
mixture of gases typically composed of CH4 and CO2, and other components such as H2S, O2 or N2 at
lower concentrations. This biogas can be employed for the generation of renewable heat and electricity
in CHP engines or being upgraded to biomethane by removing biogas impurities. Thus, biomethane is a
purified form of raw biogas that can be injected into the natural gas grid or use as a vehicle fuel substitute.
The removal of CO2, H2O, H2S and other impurities is required during biogas upgrading according to
most biomethane standards. In this context, biogas upgrading can be integrated to WWT in microalgal
or PPB ponds based on the ability of these photosynthetic organisms to fix gaseous CO2 using the
residual nutrients present in digestates (López et al., 2013). For instance, CO2 is fixed as organic carbon
by the photosynthetic apparatus of microalgae, with the concomitant production of oxygen. This O2 can
be used by sulfur oxidizing bacteria to remove H2S from biogas (Muñoz et al., 2015).
The so called digestate, which is the liquid effluent of AD, is rich in phosphorous and nitrogen.
This digestate, prior solid–liquid separation, can be used as culture media in the microalgal pond
for the generation of an algal biomass that can be further valorized as a feedstock for biostimulant
or biofertilizer production. This will ultimately reduce the energy demand (and CO2 footprint) of
digestate management (Guilayn et al., 2020).
Overall, the lower energy demand of photosynthetic systems during WWT, along with the potential
integration of biogas upgrading in photobioreactors fed with digestates and the inherent capture of the
CO2 generated during organic matter mineralization, bring new opportunities to decarbonize wastewater
treatment and enhance nutrient recovery in areas with high irradiations and moderate temperatures.
addition, aerobic activated sludge-based treatments, anaerobic digestion, and so on. (Englande et al.,
2015). However, conventional treatment configurations entail a high-energy consumption, high CO2
footprint, low nutrient recovery and environmental impacts, despite providing satisfactory reductions
of carbon, nitrogen and phosphorous concentrations in the treated water (Posadas et al., 2017a).
Microalgal technologies are eco-friendly processes that exhibit low environmental impacts, while
reducing the operating costs of conventional WWT technologies. Microalgal-based WWT is carried
out by photoautotrophic microorganisms that support a photosynthetic carbon dioxide fixation, where
sunlight energy is converted into chemical energy (biomass) by a light-driven redox reaction using
H 2O as electron donor. This process results in the assimilation of CO2 and nutrients in the form of
algal-biomass concomitantly with the release of oxygen as side-product, which can be used by bacteria
to oxidize organic matter to CO2 and ammonium to nitrate/nitrite (Masojídek et al., 2013; Rochaix
2016). In this context, algal-bacterial symbiotic bioprocesses entail lower operational costs than
conventional activated sludge systems (Figure 9.1) and a more sustainable WWT (Barreiro-Vescovo
et al., 2020; Posadas et al., 2017a). Since the photosynthetic oxygen released can replace mechanical
aeration, nutrient recovery is enhanced as a result of microalgae growth, and the overall (direct and
indirect) CO2 emissions are reduced. The so called ‘microalgae’ are photoautotrophic microorganisms
that comprise both eukaryotic microalgae and prokaryotic blue green algae or cyanobacteria (Singh
& Dhar, 2019), which possess heterotrophic and autotrophic metabolisms.
Figure 9.1 Schematic diagram showing the synergy between microbial aerobic treatment on a WWTP and a
microalgae system.
Biological CO2 fixation is typically carried out by terrestrial plants, which are able to remove only
3–6% of the CO2 supplied. In this context, the uncomplicated cellular structures and rapid growth of
microalgae endow them with higher photosynthetic and CO2 fixation efficiencies that enable 10–50
times faster CO2 fixation rates (Cuellar-Bermudez et al., 2015; Iasimone et al., 2017). A typical carbon
content of microalgal biomass averages 50% of their dry weight, which entails 1.8 kg of CO2 demand
per kg of microalgae produced (Curtis 2010; Molazadeh et al., 2019; Posadas et al., 2017a; Schediwy
et al., 2019). Optimal growth of algae requires several elements in the culture broth, mainly C/N/P, in
stoichiometric proportion to that found in the composition of the algal biomass, which is determined
by the Redfield ratio (106:16:1 C/N/P). Most WWs contain a lower C/N/P ratio than that needed
for microalgal growth, leading to carbon limitation and therefore hindering biomass growth and
nutrient recovery (Toledo-Cervantes et al., 2018). Hence, the supplementation of external C sources
is a common strategy to sustain an active microalgae growth. This additional CO2 can be obtained
from alternative emission sources such as power plants flue gas, industrial off-gases or biogas. A direct
injection of the CO2 laden gas stream into the microalgae culture via fine bubble diffusers improves
the mass transfer of CO2, increasing the concentration of inorganic carbon, and thus enhancing
microalgal biomass productivity (Rezvani et al., 2016; Toledo-Cervantes et al., 2018).
Microalgae growth is governed by environmental conditions. The most important factors
determining microalgae growth in WW are nutrients concentration, pH and alkalinity, light and
temperature. The pH of the cultivation broth is highly dependent on photosynthetic activity, alkalinity,
and microbial respiration (Posadas et al., 2017a). Photosynthetic activity increases pH as a result of
microalgal CO2 uptake from the cultivation broth. This pH can reach values of up to 11, which can
inhibit microalgal growth, although a pH range of 7–8 is considered optimum for microalgae growth.
pH can also modify the equilibrium of the nutrient species available in the cultivation broth and
impact on the gas–liquid CO2 mass transfer. Equation (9.1) shows carbon distribution in aqueous
medium as a function of the pH:
The preferred form of inorganic carbon for microalgae is species dependent. Many species are able
to use both CO2 and HCO3−, while some others are limited to only one of them (Markou et al., 2014). In
this context, since the pH of microalgae cultivation broths can range from 6.5 to 10, bicarbonate is the
dominant inorganic carbon species in most photobioreactors (Canon-Rubio et al., 2016; Chi et al., 2011).
In addition, alkalinity has a significant role in inorganic carbon speciation and governs the CO2
gas-liquid mass transfer rate to the cultivation broth. The CO2 present in gas streams such as biogas is
absorbed into the aqueous algal broth and reacts with OH− and water to form carbonate-bicarbonate
ions, while increasing total dissolved inorganic carbon (DIC) (Canon-Rubio et al., 2016; Markou et al.,
2014). Therefore, the external supplementation of CO2 does not only support pH control but also
increases DIC availability (Choi et al., 2019; Posadas et al., 2015a, 2017a).
Most biotechnologies devoted to WWT also involve the remediation of the nitrogen and
phosphorous present in wastewaters. The bioremediation potential of microalgae for inorganic
nitrogen and phosphorus from sewage has been consistently reported in literature, along with the
capacity of microalgae to remove trace organic micropollutants and heavy metals (González et al.,
2008; Ji et al., 2013; López-Serna et al., 2019; Mendez et al., 2016; Whitton et al., 2015; Yang et al.,
2015). The assimilation of nutrients by microalgae entails significantly lower energy consumptions and
CO2 emissions than conventional nitrification–denitrification processes (e.g. 1.5 kwh kg Nremoved−1) or
phosphate precipitation. Table 9.1 summarizes some of the research carried out in different wastewater
phytoremediation studies.
Nitrogen and phosphorous are also essential nutrients for the development of microalgal cultures
and both (phosphorous to a lesser extent) are limiting factors in the growth of algae (Curtis, 2010).
Nitrogen approximately represents around 5–10% of algal composition (Markou et al., 2014) and
phosphorous is around 1% (Solovchenko et al., 2016). Nitrogen concentration markedly affects the
composition of the microalgae. A limitation of the nitrogen source available in the culture media
implies the use of intracellular nitrogen to carry out metabolic functions (Pancha et al., 2014).
However, high concentration of nitrogen can lead to inhibitory effects (He et al., 2013). Likewise, pH
regulates the balance NH4 +/NH3 concentrations as shown in Equation (9.2):
pk =9.25
NH4+ + OH – NH3 + H 2O (9.2)
It has been reported that NH4 + inhibits photosynthetic activity in some microalgae species at
concentrations higher than 100 mg N-NH4 + · L −1 and pH >8 as a result of NH3 toxicity since NH3
is highly inhibitory to microalgae growth (Abeliovich & Azov, 1976; Posadas et al., 2014). Besides,
NH3 stripping occurs in open photobioreactors operated at high pH. Despite being soluble in water,
NH3 is a highly volatile compound that is dominant at high pH, which can lead to the loss of N-NH3
by volatilization. This fact is particularly relevant during WWT with mechanical aeration (Cai et al.,
2013; Jamieson et al., 2003; Mendez et al., 2016), which increases the overall CO2 footprint of the
WWT plant since NH3 is a precursor of N2O. Additionally, nitrification–denitrification processes can
Figure 9.2 Simplified metabolic diagram of purple phototrophic bacteria based on the metabolism of
Rhodopseudomonas palustris.
can even grow simultaneously utilizing light energy and organic substrates under mixotrophic growth.
Under anaerobic conditions in the presence of light energy, phototrophic growth is favored and results
in the production of energy by anoxygenic photosynthesis and biosynthesis of adenosine triphosphate
(ATP) by photophosphorylation. At this point, it should be stressed that ATP is the main molecule
used by PPB to preserve energy to be used in different metabolic routes. On the other hand, under
aerobic conditions and absence of light, the predominant metabolism will be chemotrophic since
the presence of oxygen inhibits the synthesis of bacteriochlorophyll and affects the photosynthetic
capacity of PPB (Izu et al., 2001). Therefore, oxidative phosphorylation for ATP synthesis is favored
under aerobic conditions (Lu et al., 2011). Under aerobic conditions, PPB use oxygen as an electron
acceptor and obtain energy from the proton motive force generated by consumption of NADH
(molecules synthesized from the degradation of organic compounds).
PPB can use CO2 under autotrophic mode or organic compounds under heterotrophic mode
as a carbon source. Indeed, PPB are capable of metabolizing different carbon sources due to their
great metabolic plasticity. For instance, PPB can use as a carbon source small molecules of some
carbohydrates, fatty acids and alcohols via photoheterotrophic metabolism (Lu et al., 2019b) using
multiple metabolic pathways such as tricarboxylic acid cycle, Embden–Meyerhof pathway, pentose
phosphate route or fatty acid metabolism (Larimer et al., 2004). Likewise, PPB can fix CO2 via the
Calvin–Benson–Bassham cycle, which can partially support the decarbonization of WWT (Lo et al.,
2018). For instance, an efficient metabolism in the reuse of carbon has been described in R. palustris,
which is able to use the CO2 produced by catabolic routes and fix it to synthesize biomass (Navid
et al., 2019). On the other hand, PPB can use CO2 as an electron acceptor under photoheterotrophic
mode. PPB can metabolize all forms of inorganic nitrogen like NH4 +, NO2−, NO3− and fix atmospheric
N2 (Sepúlveda-Muñoz et al., 2020b) and organic compounds containing nitrogen (i.e amino acids
or proteins). PPB can use NH3 as an electron donor under chemoautotrophic mode and NO3− as an
electron acceptor under chemoheterotrophic mode.
This inherent ability of PPB to assimilate most types of carbon and nitrogen at high biomass
yields has attracted great interest in PPB-based WWT (Capson-Tojo et al., 2020) as an alternative to
conventional biological treatments such as anaerobic digestion or activated sludge system where most
carbon and nitrogen present in wastewater is released to the atmosphere. In this context, the presence
of PNSB like R. palustris, R. sphaeroides and R. capsulatus has been reported in WWT systems due
to the low concentrations of hydrogen sulfide in this environment. The WWT ability of PPB has been
confirmed in both domestic wastewaters (Hülsen et al., 2014) and also in high strength WW such as
piggery WW (García et al., 2019; Sepúlveda-Muñoz et al., 2020a) or poultry WW (Hülsen et al., 2018).
High removal rates of carbon, nitrogen and other pollutants removal have been achieved, exceeding
removal efficiencies of 90% (Table 9.2). In addition, PPB have been proposed and validated as a
Table 9.2 Summary of wastewater treatment in batch photobioreactors with purple phototrophic bacteria.
platform for biogas upgrading coupled with WWT, reporting high efficiencies of CO2, H 2S and organic
matter (Marín et al., 2019b).
Finally, the PPB biomass generated from WWT is rich in compounds with added value such as
pigments (bacteriochlorophylls and carotenoids) and other molecules like coenzyme Q10, single-
cell protein, nutrients, polyhydroxyalkanoates, pantothenic acid and 5-aminolevulinic acid (5-ALA)
(Capson-Tojo et al., 2020; Lu et al., 2019b). These molecules can be extracted from the PPB biomass
in the context of a circular economy and WWT biorefinery.
as a modification of the widespread stabilization ponds where algae only developed on the surface
(Spellman & Drinan, 2014). Therefore, shorter hydraulic retention time (HRT) between 3 and 10
days can be applied in open raceways. Continuous mixing by paddle wheels allows for higher algae
biomass production and fully aerobic conditions in the cultivation broth. During the central hours of
the day, sunlight strikes perpendicularly to the cultivation surface, resulting in the higher treatment
capacities of the day. Indeed, oxygen oversaturation has been commonly reported under favorable
environmental conditions and moderate loading rates (Arbib et al., 2017; Hamouri, 2009). Biomass
harvesting in open ponds devoted to WWT has been carried out in settlers, settling ponds, lamella
settlers or DAF units (dissolved oxygen flotations) (Craggs et al., 2012; de Godos et al., 2016).
Open raceways are simple to build in horizontal surfaces delimited with walls or earth slopes
(Craggs et al., 2012). The original system designed by Oswald and co-workers was implemented in
the 1970s and 1980s in real scale facilities treating urban wastewater. Hence, 2- and 5-hectare open
raceways were operated for years in St. Helena and Hollister wastewater treatment plants in California
(Park et al., 2013). These demonstration units were constructed in combination with facultative ponds
and maturation ponds to optimize the process. Subsequent studies have used different integration of
the raceway reactors (also called high-rate algae ponds) into domestic WWT plants. The open algal
ponds are normally preceded by pretreatment units (mainly primary settlers) to remove suspended
materials (Craggs et al., 2012; Hamouri et al., 2003). Other demonstration units are based on the use
of anaerobically treated wastewater (Hamouri, 2009; Hamouri et al., 2003). Recently, a European
demonstration project (ALLGAS) based on algae production using urban wastewater as cultivation
media reached a positive energy production by combining anaerobic pretreatment of WW and biogas
production with algae biomass (de Godos et al., 2017). A total surface of 3.6 hectares was implemented
and bioenergy was produced as biomethane, which is used as a vehicle fuel resulting in a ratio of
energy return on investment of 2 ( www.all-gas.eu). A recent techno-economic evaluation performed
under the framework of this project showed a reduction in domestic wastewater treatment costs from
0.22 to 0.15 $ m−3 and a reduction in the energy demand by a factor of 4, which results in a significant
reduction in the indirect CO2 emissions (Acién et al., 2017)
The potential of open systems has been also evaluated for the treatment of livestock, industrial, agro-
industrial effluents (Mulbry et al., 2010, 2005; Olguín et al., 2007). In the case or piggery and cattle
wastewaters, dilution is applied in order to reduce turbidity of culture media and inhibition mediated
by elevated ammonia levels present in these livestock effluents (Godos et al., 2010). Dilution is typically
preceded by solid separation units such as sieves, settlers, and coagulation-flocculation units (Barlow
et al., 1975; González-Fernández et al., 2010). However, recent studies showed the feasibility of algae
growth in undiluted anaerobically digested swine manure using selected and acclimatized microalgae
species (Ayre et al., 2017). Algae grown in diluted cattle effluents has been recently studied as a protein
and HUFA (highly unsaturated fatty acid) feedstock for animal feed (Murry et al., 2019).
Tubular photobioreactors are made of transparent tubular pipes provided with a continuous liquid
recirculation (via centrifugal pumps or airlift units) to maintain cells suspension and light distribution
to the bulk algal broth. Different configurations have been developed for biomass production
applications: two plane tubular, near-horizontal tubular, the helical bubble reactor and α-tubular
photobioreactor (Molina-Grima et al., 2010). The arrangement of the solar collector tubes determines
the irradiated surface and light received by algae cells. The high construction and operational costs
compared to raceways have hampered the use of this complex technology for WWT purposes (Ibrahim
et al., 2020). In this context, de Godos et al. (2017) compared the performance of tubular reactors with
raceway systems and observed that although high nutrient removals were reached in the enclosed unit
(98% for N and P), collapse of algae cultures was reported after 30 days of operation due to biofouling.
Flat panels are vertical translucent plates containing the culture broth and receiving illumination
on both sides. Mixing is typically provided by continuous aeration. The first designs used glass sheets
connected with rubber, resulting in an expensive system difficult to scale-up (Samson & Leduy,
1985). Researchers have also proposed the use of plastic bags installed inside metal frames in order
to reduce the installation costs (Tredici & Materassi, 1992). Similarly to bubble columns and tubular
photobioreactors, no large-scale experiences using residual effluents as culture medium have been
reported with flat panels. The performance of a lab-scale flat panel photobioreactor for nutrient
removal from secondary effluents was studied by Ruiz et al. (2012), who achieved removal rates of 89
and 84% for nitrogen and phosphorous, respectively.
Figure 9.4 Schematic diagram of photoanaerobic membrane bioreactor for purple bacteria culture in wastewater.
are easily scaled-up and some experiences have been reported in outdoor conditions using larger
volumes of culture. For instance, 19 m3 pilot raceways were operated for more than one year using
domestic wastewater and molasses by University Nova de Lisboa and the company FCC Aqualia.
Visible light filters were placed in the complete surface to avoid microalgae growth and the process
was operated to promote the synthesis de polyhydroxyalkanoates.
9.3 ENHANCED BIOGAS PRODUCTION FROM MICROALGAE AND PPB FOR WWT
DECARBONIZATION
Circular bioeconomy has emerged in recent years as an essential component of sustainable and green
industrial activity. This approach focuses on utilizing all the potential of natural resources through
cascading biomass use and recycling, while ensuring that natural capital is preserved (Rajesh Banu
et al., 2020) In this context, photosynthetic biomass biorefining is presented as a promising approach
to convert algal or PBB biomass to value-added products, biofuels, and chemicals. Thus, microalgae
or PPB from WWT or residues of these biomass sources, for instance those from microalgal lipid
extraction for biodiesel production (Uggetti et al., 2017), can be used as a substrate of anaerobic
digestion processes, which in turn can provide digestate for microalgal growth based on the high
nutrient and inorganic carbon content of these AD effluents (Figure 9.5).
Anaerobic digestion involves a series of biological reactions where the breakdown of complex organic
matter is carried out in the absence of oxygen, nitrate, nitrite or sulfate (typically used as electron
acceptors). Anaerobic digestion occurs as a result of four sequential steps: hydrolysis, acidogenesis,
acetogenesis and methanogenesis. A complex network of anaerobic bacteria, archaea and fungi are
responsible for this process, which results in the production of biogas and the release of nutrients to
the anaerobic broth (Choi et al., 2019; Sanz et al., 2017; Uggetti et al., 2017). Biogas is a mixture of
gases, mainly CH4 and CO2 at concentrations in the range of 50–80 and 15–50%, respectively, with
other compounds such as water, H 2S, NH3, N2 and O2 present at lower concentrations. The methane
produced in AD can be used on-site as fuel gas in WWT plants to generate heat in boilers or heat and
power generation in gas engines or turbines, which allows for the decrease in the overall CO2 footprint
of the WWT. Alternative uses such as substitute of natural gas or biofuel for vehicles engines require
biogas upgrading for the removal of impurities, to increase the methane content up to 90% CH4
(Muñoz et al., 2015; Uggetti et al., 2017). Biogas upgrading contributes to an increase in the calorific
value of methane while reducing transportation costs and minimize corrosion in pipelines, engines
and biogas storage structures produced by some contaminants (Marín et al., 2019a).
Anaerobic digestion is a mature technology already established in activated sludge-based WWT
plants to obtain biogas from the treatment of the sludge generated during primary and secondary
treatment. However, the versatility of AD allows biogas production from a large range of substrates
including organic waste such as agricultural residues, animal manure, energy crops, micro and
microalgae and even PPB. Thus, biogas generation entails less technical complexity and environmental
impacts than the production of other biofuels since the extraction of specific components of the
biomass is not required, which increases the efficiency of the overall process.
Eukaryotic microalgae are composed of a semi-rigid structure or cell wall that confers cell protection
against physical, chemical and biological agents. The composition of microalgae cell wall is species-
specific and based on complex structures as cellulose, hemicellulose, pectin and glycoproteins, which
makes cell walls highly resistant structures (Gonzalez-Fernandez et al., 2017). The composition and
structure of microalgae cell wall influences AD performance and consequently the potential methane
yield. Indeed, the cell wall hinders the anaerobic biodegradability of most microalgae species due to
its high recalcitrance and resistance to microbial attack (Uggetti et al., 2017). In this context, a broad
variety of pretreatments can be found in literature for the disruption of microalgal cells in order to
increase the final yields of biogas production (Gonzalez-Fernandez et al., 2017; Mahdy et al., 2014a,
2014b; Mendez et al., 2014b; Passos et al., 2014, 2015a). However, the differences among composition
and structural characteristics of different microalgal species requires tailoring the selection of the
optimal method for each microalgal biomass. Pretreatments can be classified as chemical, thermal,
mechanical or biological. Thermal pretreatments entail the application of heat for organic matter
disruption and solubilization. Several studies have consistently proven the efficiency of thermal
pretreatments, however, the application of high temperature to microalgae biomass may lead to the
formation of recalcitrant compounds that could eventually inhibit anaerobic digestion (Atelge et al.,
2020; Carrère et al., 2016). Therefore, the application of thermal methods requires the optimization of
the operational conditions to each particular substrate. For instance, Mendez et al. (2015) evaluated
the efficiency of thermal pretreatments using C. vulgaris as a feedstock for methane production in batch
and semicontinuous mode in a CSTR reactor. Hence, the anaerobic digestion of raw and thermally
pretreated algal biomass at 120°C for 40 min was compared, and a significant increase from 138
and 85 mL CH4 g CODin.–1 with raw microalga up to 266 and 126 mL CH4 g CODin.–1 with thermally
pretreated algae was recorded under batch and semi-continuous mode, respectively. Similar results
were obtained by Schwede et al. (2013) in terms of increase in methane yield using Nannochloropsis
salina pretreated at 100 and 120°C for 2 and 8 h. The methane yields of the pretreated microalgae
compared with raw microalga increased by 2 to 2.85-fold in CSTR and batch mode, respectively.
Chemical pretreatments are commonly used to solubilize polymers. These pretreatments are also
associated with the potential formation of by-products that may induce inhibition of anaerobic
microbial communities. On the other hand, mechanical pretreatments support cell wall disruption
by structural fragmentation of the recalcitrant organic matter using mechanical forces. Mechanical
methods imply the application of shear forces, pressure or energy using bead milling, homogenization,
ultrasonication and microwaves. Mechanical pretreatments have been widely used in microalgae,
showing efficient results independently on the algal species employed (Barragán-Trinidad & Buitrón,
2020; Passos et al., 2015b). The energy requirements of these pretreatments are sometimes substantially
higher compared to the energy recovered as methane, which represents their main limitation. For
instance, Passos et al. (2013) reported a maximum biogas yield of 307 mL biogas·g VS−1, which
corresponded to a 78% increase compared with raw microalga when using microwaves at different
specific energies on microalgae cultivated with WW effluent from a hydrolytic up-flow sludge blanket
(HUSB) reactor fed with urban WW.
Biological pretreatments involve the use of enzymes responsible for the solubilization of recalcitrant
biomass. Enzymes are molecules that will bond to a specific target of the cell to perform the lysis and
solubilization of the organic compounds. Enzymatic methods represent a promising alternative to high
energy-consuming pretreatments and have recently generated great interest since hydrolytic enzymes
perform cell lysis under mild reaction conditions, without the generation of inhibitory side-products,
and with a low energy demand. However, despite the use of enzymes is an effective alternative to
thermal, chemical and physical methods, enzymes are costly molecules that need to be continuously
supplied to the feedstock (Barragán-Trinidad & Buitrón, 2020). The selection of adapted anaerobic
inocula to the substrate, in this particular case microalgal biomass, has been recently proposed as a low-
cost alternative to the use of pretreatments methods. For instance, Gonzalez-Fernandez et al. (2018)
reported differences in methane yields with inocula adapted to degrade sewage sludge and to digest
microalgae biomass. In this study, microalgal adapted sludge showed 79.2 ± 3.1 and 108.2 ± 1.9 mL CH4
g COD in−1 under mesophilic and thermophilic conditions, respectively, when digesting Scenedesmus
sp. biomass, while non-adapted anaerobic sludge only supported to 63.1 ± 3.1 mL CH4 g COD in−1 in
mesophilic conditions. However, there was no remarkable difference on the final yield (which ranged
from 105 to 114 mL CH4 g COD in−1 for the tested sludges) when digesting Chlorella sorokiniana. This
study confirmed the relevance of the previous adaptation of the sludge inocula to the biomass to be
digested as well as the key role of the microalgae biomass digested.
Despite PPB having been consistently proven to efficiently remove WW pollutants and operate as a
promising emerging tool for bioremediation technologies, the AD of PPB biomass has not been largely
assessed and literature related to digestibility of this biomass is scarce. Some authors have suggested
that since PPB are also anaerobic biomass, these organisms will be tolerant to the reducing conditions
of the digestor, which may limit the AD performance. Additionally, PPB are rich in proteins which
can eventually limit biomass degradability due to hydrolytic limitations. Hülsen et al. (2020) reported
anaerobic VS degradations of approximately 55% and methane yield of 330 ± 4.3 and 315 ± 2.1 mLCH4
gVS−1 under mesophilic and thermophilic conditions, respectively, in a continuous digester treating
PPB grown in domestic WW. However, the relatively low economic profits obtained from biogas
and the low digestibility exhibited by PPB grown in WW recommend anaerobic digestion only when
alternative PPB valorization strategies are not feasible (Capson-Tojo et al., 2020).
biogas to the algal-bacterial broth. For instance, the increase of the pH from 7 to 10 in an indoors
high-rate algal pond (HRAP) interconnected to an absorption column resulted in an increase of CO2
removal from less than 20% to almost 100% (Bahr et al., 2014). Although most microalgae show a
maximum activity at pH 7–8, some microalgae/cyanobacteria species such as Anabaena, Spirulina,
Chlorella, Chlorococcum and Scenedesmus are suitable for photosynthetic biogas upgrading due to
their ability to grow at high pH and CO2 concentrations (Bose et al., 2019). A high buffer capacity
mediated by a high alkalinity or inorganic carbon (IC) concentration in the cultivation broth is also
necessary to prevent an elevated pH decrease in the absorption column as a result of the overload
of these acidic gases or biological processes such as nitrification, which tend to acidify the algal-
bacterial broth of the photobioreactor. In this context, the use of the liquid fraction of digestates
instead of domestic wastewater as a low-cost nutrient source during photosynthetic biogas upgrading
is preferable due to its higher pH and alkalinity (Rodero et al., 2019). Nevertheless, IC concentrations
in the cultivation broth of >2400 mg C·L −1 involve a high salt content (carbonates), which exerts a
negative impact on photosynthetic activity along with an increase in CO2 stripping to the atmosphere
from the photobioreactor surface (Rodero et al., 2020b). In this regard, process operation with an
optimum alkalinity in the cultivation broth is a must to avoid acidification without compromising the
environmental benefits of this biotechnology.
Environmental factors such as temperature and dissolved oxygen (DO) in the cultivation broth
also have influence on both biomethane quality and the subsequent CO2 uptake. Although the optimal
temperature for microalgae growth often ranges between 28 and 35°C, low temperatures support
higher CO2 and H 2S removals due to a higher solubility of the gases in the cultivation broth. However,
this effect is minimum at medium-high alkaline cultivation conditions (Park et al., 2011; Rodero et al.,
2018). Otherwise, large amounts of DO in the cultivation broth result in a high O2 desorption from
the liquid to the biogas as well as the inhibition of photosynthetic activity (Pawlowski et al., 2015;
Posadas et al., 2015b).
On the other hand, an optimum design and operation of the process is also necessary to enhance
upgraded biogas quality. In this context, the mass transfer of CO2 and H 2S from the biogas to
the cultivation broth can take place in the photobioreactor or in an external absorption column
interconnected to the photobioreactor (i.e. HRAP or tubular), the latter configuration being preferred
since it prevents a high oxygen stripping from the cultivation broth to the biogas besides entailing a
more effective gas–liquid mass transfer due to the larger biogas bubble residence times (Figure 9.6)
(Meier et al., 2015). The liquid to biogas (L/G) ratio is a key operational parameter determining
Figure 9.6 Schematic diagram of the algal–bacterial process for the simultaneous biogas upgrading and wastewater/
digestate treatment.
Experimental Set-up CO2-RE Upgraded Biogas Operational Parameters Microalgae Population References
Design (%) Composition (%)
180 L HRAP 40–95 O2<1 H 2S:0 Indoors Spirulina platensis, Bahr et al.
interconnected to a pH: 7–10 Phormidium, Oocystis, (2014)
0.8 L bubble column L/G ratio: 0.4–1.6 Microspora sp.
Nutrient source: synthetic
medium and diluted centrate
180 L HRAP 80 O2:0.3–3 N2:6–10 H 2S:0 Indoors Chlorella sp., Pseudanabaena Serejo et al.
interconnected to a pH: 8 sp., Chloromonas sp., (2015)
2.5 L bubble column L/G ratio: 0.5–67 Geitlerinema sp., Microspora
Nutrient source: diluted sp., Stigeoclonium sp.,
anaerobically digested vinasse Planktolyngbya sp.
180 L HRAP 72–79 CH4:81 Indoors Geitlerinema sp., Limnothrix Posadas et al.
interconnected to a CO2:6.8–9.2 O2:0.7–1.2 pH: 8 planktonica, Pseudanabaena (2015b)
2.5 L bubble column N2:5.9–7.2 L/G ratio: 10.7 minima, Stigeoclonium
H 2S:0 Nutrient source: diluted tenue, Leptolyngbya
H 2S:0–0.01 medium
(Continued)
by guest
160
Table 9.3 Experimental studies on photosynthetic biogas upgrading under different configurations (Continued).
Experimental Set-up CO2-RE Upgraded Biogas Operational Parameters Microalgae Population References
Design (%) Composition (%)
45.6 L tubular 85–96 CH4:90.8–97.2 Indoors Aphanothece sp., Chlorella Ángeles et al.
photobioreactor CO2:1.1–6.2 pH: 8.9–9.3 sp., Chlorella vulgaris, (2020a)
interconnected to an O2:0.4–0.8 L/G ratio: 0.5 Mayamaea sp., Chlorella
84 L mixing chamber N2: 1.6–2.0 Nutrient source: synthetic homosphaera, Pseudanabaena
and a 2.6 L bubble H 2S:0 medium sp.
column
Closed photobioreactor 98 CH4:50–53 CO2:1.2–2.5 Indoors Chlorella vulgaris Mann et al.
of 1 L O2:18.3–23.4 pH: 5.5–7 (2009)
H 2S:0 Nutrient source: synthetic
the CO2 and H 2S gas-liquid mass transfer and O2/N2 stripping in the absorption column. In fact, a
control strategy to guarantee a biomethane quality over time under the typical daily and seasonal
variations in environmental conditions based on the optimization of the L/G ratio was successfully
validated under semi-industrial scale (Rodero et al., 2020a). High L/G ratios enhance CO2 and H 2S
removal efficiencies due to lower acidification of the scrubbing liquid along the absorption column at
the expense of increasing O2 and N2 stripping from the liquid to the upgraded biogas (Rodero et al.,
2019; Serejo et al., 2015). The gas–liquid flow configuration in the absorption column also determines
the upgraded biogas composition. Co-current flow operation is preferred due to lower O2 and N2
desorption and no S accumulation in the diffuser, although counter-current gas–liquid configuration
favors the gas–liquid mass transfer rates (Toledo-Cervantes et al., 2017). On the other hand, innovative
operational strategies to enhance biomethane quality have been recently evaluated. In this context,
the installation of a hollow fibre membrane prior to the biogas absorption column and the increase
in the operational pressure in the biogas absorption column to minimize N2 and O2 content in the
upgraded biogas have been validated at pilot scale as promising approaches (Ángeles et al., 2020b,
2020c). Moreover, the location of an HRAP inside a greenhouse and mechanical CO2 stripping during
winter conditions have been demonstrated to increase CO2 removal under unfavorable climatic
conditions (Marín et al., 2021).
Despite CO2 capture biotechnology still being under demo-scale validation, it constitutes
an attractive alternative to the conventional technologies for biogas upgrading based on its cost-
effectiveness and environmentally friendliness. In this context, CO2 and H 2S removal efficiencies as
high as 99 and 100%, respectively, with a final CH4 content in the upgraded biogas as high as 98%
being reported (Table 9.3). Furthermore, this biotechnology exhibits low CH4 losses (≤5%) as a result
of the poor aqueous solubility of this potent greenhouse gas (Posadas et al., 2017b).
ACKNOWLEDGEMENT
The Spanish Ministry of Science, Innovation and Universities (FJC 2018-038402-I), the Regional
Government of Castilla y León and the EU-FEDER programme (CLU 2017-09, VA281P18 and
UIC 71) and CONICYT (PFCHA/DOCTORADO BECAS CHILE/2017–72180211) are gratefully
acknowledged.
REFERENCES
Abeliovich A. and Azov Y. (1976). Toxicity of ammonia to algae in sewage oxidation ponds. Applied and
Environmental Microbiology, 31, 801–806, https://doi.org/10.1128/aem.31.6.801-806.1976
Acién F. G., Fernández-Sevilla J. M. and Molina Grima E. (2013). Photobioreactors for the production of
microalgae. Reviews in Environmental Science and Bio/Technology, 12, 131–151, https://doi.org/10.1007/
s11157-012-9307-6
Acién F. G., Molina E., Fernández-Sevilla J. M., Barbosa M., Gouveia L., Sepúlveda C., Bazaes J., and Arbib Z.
(2017). Economics of microalgae production. In: C. Gonzalez-Fernandez and R. Muñoz (eds). Microalgae-
Based Biofuels and Bioproducts: From Feedstock Cultivation to End-Products. Elsevier Inc. Amsterdam,
pp. 485–503.
Ángeles R., Arnaiz E., Gutiérrez J., Sepúlveda-Muñoz C. A., Fernández-Ramos O., Muñoz R. and Lebrero R. (2020a).
Optimization of photosynthetic biogas upgrading in closed photobioreactors combined with algal biomass
production. Journal of Water Process Engineering, 38, 101554, https://doi.org/10.1016/j.jwpe.2020.101554
Ángeles R., Marín D., Rodero M. del R., Pascual C., González-Sanchez A., de Godos I., Lebrero R. and Muñoz
R. (2020b). Biogas treatment for H 2S, CO2, and other contaminants removal. In: Biofiltration to Promising
Options in Gaseous Fluxes Biotreatment, G. Soreanu and É. Dumont (eds.), Elsevier, Amsterdam, pp.
153–176.
Ángeles R., Rodríguez Á, Domínguez C., García J., Prádanos P., Muñoz R. and Lebrero R. (2020c). Strategies for
N2 and O2 removal during biogas upgrading in a pilot algal-bacterial photobioreactor. Algal Research, 48,
101920, https://doi.org/10.1016/j.algal.2020.101920
Arbib Z., de Godos Crespo I., Corona E. L. and Rogalla F. (2017). Understanding the biological activity of high
rate algae ponds through the calculation of oxygen balances. Applied Microbiology and Biotechnology, 101,
5189–5198, https://doi.org/10.1007/s00253-017-8235-3
Arun J., Shreekanth S. J., Sahana R., Raghavi M. S., Gopinath K. P. and Gnanaprakash D. (2017). Studies on influence
of process parameters on hydrothermal catalytic liquefaction of microalgae (chlorella vulgaris) biomass
grown in wastewater. Bioresource Technology, 244, 963–968, https://doi.org/10.1016/j.biortech.2017.08.048
Atelge M. R., Atabani A. E., Rajesh Banu J., Krisa D., Kaya M., Eskicioglu C., Kumar G., Lee C., Yildiz Y. Ş., Unalan
S., Mohanasundaram R. and Duman F. (2020). A critical review of pretreatment technologies to enhance
anaerobic digestion and energy recovery. Fuel, 270, 117494, https://doi.org/10.1016/j.fuel.2020.117494
Awe O. W., Yaqian Z., Ange N., Doam Pham M. and Nathalie L. (2017). A review of biogas utilisation, purification
and upgrading technologies. Waste and Biomass Valorization, 8, 267–283, https://doi.org/10.1007/
s12649-016-9826-4
Ayre J. M., Moheimani N. R. and Borowitzka M. A. (2017). Growth of microalgae on undiluted anaerobic
digestate of piggery effluent with high ammonium concentrations. Algal Research, 24, 218–226, https://doi.
org/10.1016/j.algal.2017.03.023
Bahr M., Díaz I., Dominguez A., González Sánchez A. and Muñoz R. (2014). Microalgal-biotechnology as a
platform for an integral biogas upgrading and nutrient removal from anaerobic effluents. Environmental
Science and Technology, 48, 573–581, https://doi.org/10.1021/es403596m
Barlow E. W. R., Boersma L., Phinney H. K. and Miner J. R. (1975). Algal growth in diluted pig waste. Agriculture
and Environment, 2(4), 339–355, https://doi.org/10.1016/0304-1131(75)90040-5
Barragán-Trinidad M. and Buitrón G. (2020). Hydrogen and methane production from microalgal biomass
hydrolyzed in a discontinuous reactor inoculated with ruminal microorganisms. Biomass and Bioenergy,
143, 105825, https://doi.org/10.1016/j.biombioe.2020.105825
Barreiro-Vescovo S., González-Fernández C., Ballesteros M. and de Godos I. (2020). Activity determination of
an algal-bacterial consortium developed during wastewater treatment based on oxygen evolution. Journal of
Water Process Engineering, 36, 101278, https://doi.org/10.1016/j.jwpe.2020.101278
Batstone D. J., Hülsen T., Mehta C. M. and Keller J. (2015). Platforms for energy and nutrient recovery from
domestic wastewater: A review. Chemosphere, 140, 2–11, https://doi.org/10.1016/j.chemosphere.2014.10.021
Bose A., Lin R., Rajendran K., O’Shea R., Xia A. and Murphy J. D. (2019). How to optimise photosynthetic biogas
upgrading: a perspective on system design and microalgae selection. Biotechnology Advances, 37, 107444,
https://doi.org/10.1016/j.biotechadv.2019.107444
Budiman P. M., Wu T. Y., Ramanan R. N. and Hay J. X. W. (2014). Treatment and reuse of effluents from palm oil,
pulp, and paper mills as a combined substrate by using purple nonsulfur bacteria. Industrial & Engineering
Chemistry Research, 53(39), 14921–14931, https://doi.org/10.1021/ie501798f
Cai T., Park S. Y. and Li Y. (2013). Nutrient recovery from wastewater streams by microalgae: Status and prospects.
Renewable & Sustainable Energy Reviews, 19, 360–369, https://doi.org/10.1016/j.rser.2012.11.030
Canon-Rubio K. A., Sharp C. E., Bergerson J., Strous M. and De la Hoz Siegler H. (2016). Use of highly alkaline
conditions to improve cost-effectiveness of algal biotechnology. Applied Microbiology and Biotechnology,
100, 1611–1622, https://doi.org/10.1007/s00253-015-7208-7
Capson-Tojo G., Batstone D. J., Grassino M., Vlaeminck S. E., Puyol D., Verstraete W., Kleerebezem R., Oehmen
A., Ghimire A., Pikaar I., Lema J. M. and Hülsen T. (2020). Purple phototrophic bacteria for resource
recovery: challenges and opportunities. Biotechnology Advances, 43, 107567, https://doi.org/10.1016/j.
biotechadv.2020.107567
Carrère H., Antonopoulou G., Affes R., Passos F., Battimelli A., Lyberatos G. and Ferrer I. (2016). Review of
feedstock pretreatment strategies for improved anaerobic digestion: from lab-scale research to full-scale
application. Bioresource Technology, 199, 386–397, https://doi.org/10.1016/j.biortech.2015.09.007
Čater M., Fanedl L., Malovrh Š. and Logar R. M. (2015). Biogas production from brewery spent grain enhanced
by bioaugmentation with hydrolytic anaerobic bacteria. Bioresource Technology, 186, 261–269, https://doi.
org/10.1016/j.biortech.2015.03.029
Cembella A. D., Antia N. J. and Harrison P. J. (1982). The utilization of inorganic and organic phosphorous
compounds as nutrients by eukaryotic microalgae: a multidisciplinary perspective: part 1. Critical Reviews
in Microbiology, 10, 317–391, https://doi.org/10.3109/10408418209113567
Cheng P., Cheng J. J., Cobb K., Zhou C., Zhou N., Addy M., Chen P., Yan X. and Ruan R. (2020). Tribonema sp.
and Chlorella zofingiensis co-culture to treat swine wastewater diluted with fishery wastewater to facilitate
harvest. Bioresource Technology, 297, 122516, https://doi.org/10.1016/j.biortech.2019.122516
Chi Z., O’Fallon J. V. and Chen S. (2011). Bicarbonate produced from carbon capture for algae culture. Trends in
Biotechnology, 29(11), 537–541, https://doi.org/10.1016/j.tibtech.2011.06.006
Choi Y. Y., Patel A. K., Hong M. E., Chang W. S. and Sim S. J. (2019). Microalgae bioenergy with carbon capture
and storage (BECCS): An emerging sustainable bioprocess for reduced CO2 emission and biofuel production.
Bioresource Technology Reports, 7, 100270, https://doi.org/10.1016/j.biteb.2019.100270
Choorit W., Thanakoset P., Thongpradistha J., Sasaki K. and Noparatnaraporn N. (2002). Identification
and cultivation of photosynthetic bacteria in wastewater from a concentrated latex processing factory.
Biotechnology Letters, 24, 1055–1058, https://doi.org/10.1023/A:1016026412361
Chu W.-L. (2012). Biotechnological applications of microalgae. International Journal of Medical Science and
Education, 6, 24–37.
Converti A., Oliveira R. P. S., Torres B. R., Lodi A. and Zilli M. (2009). Biogas production and valorization by
means of a two-step biological process. Bioresource Technology, 100, 5771–5776, https://doi.org/10.1016/j.
biortech.2009.05.072
Correll D. L. (1998). The role of phosphorus in the eutrophication of receiving waters: A review. Journal of
Environmental Quality, 27(2), 261–266, https://doi.org/10.2134/jeq1998.00472425002700020004x
Craggs R., Sutherland D. and Campbell H. (2012). Hectare-scale demonstration of high rate algal ponds for
enhanced wastewater treatment and biofuel production. Journal of Applied Phycology, 24, 329–337, https://
doi.org/10.1007/s10811-012-9810-8
Cuellar-Bermudez S. P., Garcia-Perez J. S., Rittmann B. E. and Parra-Saldivar R. (2015). Photosynthetic bioenergy
utilizing CO2: an approach on flue gases utilization for third generation biofuels. Journal of Cleaner
Production, 98, 53–65, https://doi.org/10.1016/j.jclepro.2014.03.034
Curtis T. P. (2010). Low-energy wastewater treatment: strategies and technologies. In: Environmental Microbiology,
R. Mitchell and J.-D. Gu (eds.), John Wiley & Sons, Inc., Hoboken, NJ, USA, pp. 301–318.
Dalaei P., Ho D., Nakhla G. and Santoro D. (2019). Low temperature nutrient removal from municipal wastewater
by purple phototrophic bacteria (PPB). Bioresource Technology, 288, 121566, https://doi.org/10.1016/j.
biortech.2019.121566
Dalaei P., Bahreini G., Nakhla G., Santoro D., Batstone D. and Hülsen T. (2020). Municipal wastewater treatment
by purple phototropic bacteria at low infrared irradiances using a photo-anaerobic membrane bioreactor.
Water Research, 173(4), 115535, https://doi.org/10.1016/j.watres.2020.115535
De-Bashan L. E. and Bashan Y. (2004). Recent advances in removing phosphorus from wastewater and its future
use as fertilizer (1997–2003). Water Research, 38, 4222–4246, https://doi.org/10.1016/j.watres.2004.07.014
de Godos I., Blanco S., García-Encina P. A., Becares E. and Muñoz R. (2009). Long-term operation of high rate
algal ponds for the bioremediation of piggery wastewaters at high loading rates. Bioresource Technology,
100, 4332–4339, https://doi.org/10.1016/j.biortech.2009.04.016
de Godos I., Arbib Z., Lara E. and Rogalla F. (2016). Evaluation of high rate algae ponds for treatment of
anaerobically digested wastewater: effect of CO2 addition and modification of dilution rate. Bioresource
Technology, 220, 253–261, https://doi.org/10.1016/j.biortech.2016.08.056
de Godos I., Arbib Z., Lara E., Cano R., Muñoz R. and Rogalla F. (2017). Wastewater treatment in algal systems.
In: Innovative Wastewater Treatment & Resource Recovery Technologies: Impacts on Energy, Economy and
Environment, J. Lema and S. Suarez (eds.), IWA Publishing, London, pp. 76–95.
de las Heras I., Molina R., Segura Y., Hülsen T., Molina M. C., Gonzalez-Benítez N., Melero J. A., Mohedano A.
F., Martínez F. and Puyol D. (2020). Contamination of N-poor wastewater with emerging pollutants does
not affect the performance of purple phototrophic bacteria and the subsequent resource recovery potential.
Journal of Hazardous Materials, 385, 121617, https://doi.org/10.1016/j.jhazmat.2019.121617
de Mendonça H. V., Ometto J. P. H. B., Otenio M. H., Marques I. P. R. and dos Reis A. J. D. (2018). Microalgae-
mediated bioremediation and valorization of cattle wastewater previously digested in a hybrid anaerobic
reactor using a photobioreactor: comparison between batch and continuous operation. Science of the Total
Environment, 633, 1–11, https://doi.org/10.1016/j.scitotenv.2018.03.157
Eixler S., Karsten U. and Selig U. (2006). Phosphorus storage in Chlorella vulgaris (Trebouxiophyceae, Chlorophyta)
cells and its dependence on phosphate supply. Phycologia, 45, 53–60, https://doi.org/10.2216/04-79.1
Englande A. J., Krenkel P. and Shamas J. (2015). Wastewater treatment and water reclamation. Reference Module in
Earth Systems and Environmental Sciences, 639–670, https://doi.org/10.1016/B978-0-12-409548-9.09508-7
Fouilland E., Vasseur C., Leboulanger C., Le Floc’h E., Carré C., Marty B., Steyer J.-P. P. and Sialve B. (2014). Coupling
algal biomass production and anaerobic digestion: production assessment of some native temperate and
tropical microalgae. Biomass and Bioenergy, 70, 564–569, https://doi.org/10.1016/j.biombioe.2014.08.027
Franco-Morgado M., Alcántara C., Noyola A., Muñoz R. and González-Sánchez A. (2017). A study of photosynthetic
biogas upgrading based on a high rate algal pond under alkaline conditions: influence of the illumination
regime. Science of the Total Environment, 592, 419–425, https://doi.org/10.1016/j.scitotenv.2017.03.077
García D., Alcántara C., Blanco S., Pérez R., Bolado S. and Muñoz R. (2017). Enhanced carbon, nitrogen and
phosphorus removal from domestic wastewater in a novel anoxic-aerobic photobioreactor coupled with
biogas upgrading. Chemical Engineering Journal, 313, 424–434, https://doi.org/10.1016/j.cej.2016.12.054
García D., de Godos I., Domínguez C., Turiel S., Bolado S. and Muñoz R. (2019). A systematic comparison of the
potential of microalgae-bacteria and purple phototrophic bacteria consortia for the treatment of piggery
wastewater. Bioresource Technology, 276, 18–27, https://doi.org/10.1016/j.biortech.2018.12.095
Godos I. D., Vargas V. A., Blanco S., González M. C. G., Soto R., García-Encina P. A., Becares E. and Muñoz R. (2010).
A comparative evaluation of microalgae for the degradation of piggery wastewater under photosynthetic
oxygenation. Bioresource Technology, 101(14), 5150–5158, https://doi.org/10.1016/j.biortech.2010.02.010
González C., Marciniak J., Villaverde S., García-Encina P. A. and Muñoz R. (2008). Microalgae-based processes
for the biodegradation of pretreated piggery wastewaters. Applied Microbiology and Biotechnology, 80,
891–898, https://doi.org/10.1007/s00253-008-1571-6
González-Fernández C., León-Cofreces C. and García-Encina P. A. (2008). Different pretreatments for increasing
the anaerobic biodegradability in swine manure. Bioresource Technology, 99, 8710–8714, https://doi.
org/10.1016/j.biortech.2008.04.020
González-Fernández C., Molinuevo-Salces B. and García-González M. C. (2010). Open and enclosed photobioreactors
comparison in terms of organic matter utilization, biomass chemical profile and photosynthetic efficiency.
Ecological Engineering, 36, 1497–1501, https://doi.org/10.1016/j.ecoleng.2010.07.007
Gonzalez-Fernandez C., Mendez L., Ballesteros M. and Tomas-Pejó E. (2017). Hydrothermal processing of
microalgae. In: Hydrothermal Processing in Biorefineries: Production of Bioethanol and High Added-Value
Compounds of Second and Third Generation Biomass, A. Ruiz Leza, M. Thomsen and H. L. Trajano (eds.),
Springer International Publishing, New York, pp. 483–500.
Gonzalez-Fernandez C., Barreiro-Vescovo S., De Godos I., Fernandez M., Zouhayr A. and Ballesteros M. (2018).
Biochemical methane potential of microalgae biomass using different microbial inocula. Biotechnology for
Biofuels, 11, 1–11, https://doi.org/10.1186/s13068-018-1188-7
Guilayn F., Rouez M., Crest M., Patureau D. and Jimenez J. (2020). Valorization of digestates from urban or
centralized biogas plants: a critical review. Reviews in Environmental Science and Bio/Technology, 19,
419–462, https://doi.org/10.1007/s11157-020-09531-3
Hamouri B. E. (2009). Rethinking natural, extensive systems for tertiary treatment purposes: The high-rate algae
pond as an example. Desalination and Water Treatment, 4, 128–134, https://doi.org/10.5004/dwt.2009.367
Hamouri B. E., Rami A. and Vasel J. L. (2003). The reasons behind the performance superiority of a high rate
algal pond over three facultative ponds in series. Water Science and Technology, 48(2), 269–276, https://doi.
org/10.2166/wst.2003.0130
He P. J., Mao B., Shen C. M., Shao L. M., Lee D. J. and Chang J. S. (2013). Cultivation of Chlorella vulgaris on
wastewater containing high levels of ammonia for biodiesel production. Bioresource Technology, 129, 177–
181, https://doi.org/10.1016/j.biortech.2012.10.162
Hiraishi A., Shi J. L. and Kitamura H. (1989). Effects of organic nutrient strength on the purple nonsulfur bacteria
content and metabolic activity of photosynthetic sludge for wastewater treatment. Journal of Fermentation
and Bioengineering, 68(4), 269–276, https://doi.org/10.1016/0922-338X(89)90028-7
Hülsen T., Batstone D. J. and Keller J. (2014). Phototrophic bacteria for nutrient recovery from domestic wastewater.
Water Research, 50, 18–26, https://doi.org/10.1016/j.watres.2013.10.051
Hülsen T., Barry E. M., Lu Y., Puyol D. and Batstone D. J. (2016a). Low temperature treatment of domestic
wastewater by purple phototrophic bacteria: performance, activity, and community. Water Research, 100,
537–545, https://doi.org/10.1016/j.watres.2016.05.054
Hülsen T., Barry E. M., Lu Y., Puyol D., Keller J. and Batstone D. J. (2016b). Domestic wastewater treatment
with purple phototrophic bacteria using a novel continuous photo anaerobic membrane bioreactor. Water
Research, 100, 486–495, https://doi.org/10.1016/j.watres.2016.04.061
Hülsen T., Hsieh K., Tait S., Barry E. M., Puyol D. and Batstone D. J. (2018). White and infrared light continuous
photobioreactors for resource recovery from poultry processing wastewater – A comparison. Water Research,
144, 665–676, https://doi.org/10.1016/j.watres.2018.07.040
Hülsen T., Hsieh K. and Batstone D. J. (2019). Saline wastewater treatment with purple phototrophic bacteria.
Water Research, 160, 259–267, https://doi.org/10.1016/j.watres.2019.05.060
Hülsen T., Lu Y., Rodríguez I., Segura Y., Martínez F., Puyol D. and Batstone D. J. (2020). Anaerobic digestion
of purple phototrophic bacteria – The release step of the partition-release-recover concept. Bioresource
Technology, 306, 123125, https://doi.org/10.1016/j.biortech.2020.123125
Hunter C. N., Daldal F., Thurnauer M. C. and Beatty T. J. (2009). The Purple Phototrophic Bacteria, 1st edn.
Springer, The Netherlands.
Iasimone F., De Felice V., Panico A. and Pirozzi F. (2017). Experimental study for the reduction of CO2 emissions
in wastewater treatment plant using microalgal cultivation. Journal of CO2 Utilization, 22, 1–8, https://doi.
org/10.1016/j.jcou.2017.09.004
Ibrahim G., Muñoz R., Llamas B. and de Godos I. (2020). Carbon dioxide capture from carbon dioxide-rich gases
by microalgae. In: From Biofiltration to Promising Options in Gaseous Fluxes Biotreatment, G. Soreanu and
E. Dumont (eds.), Elsevier, Amsterdam, pp. 373–396.
Izu K., Nakajima F., Yamamoto K. and Kurisu F. (2001). Aeration conditions affecting growth of purple nonsulfur
bacteria in an organic wastewater treatment process. Systematic and Applied Microbiology, 24, 294–302,
https://doi.org/10.1078/0723-2020-00027
Jamieson T. S., Stratton G. W., Gordon R. and Madani A. (2003). The use of aeration to enhance ammonia
nitrogen removal in constructed wetlands. Canadian Biosystems Engineering, 45, 9–14.
Ji M.-K., Abou-Shanab R. A. I., Kim S.-H., Salama E.-S., Lee S.-H., Kabra A. N., Lee Y.-S., Hong S. and Jeon
B.-H. (2013). Cultivation of microalgae species in tertiary municipal wastewater supplemented with CO2 for
nutrient removal and biomass production. Ecological Engineering, 58, 142–148, https://doi.org/10.1016/j.
ecoleng.2013.06.020
Kalra R., Gaur S. and Goel M. (2020). Microalgae bioremediation: A perspective towards wastewater treatment
along with industrial carotenoids production. Journal of Water Process Engineering, 40, 101794, https://doi.
org/10.1016/j.jwpe.2020.101794
Karemore A., Ramalingam D., Yadav G., Subramanian G. and Sen R. (2016). Photobioreactors for improved algal
biomass production: analysis and design considerations. In: Algal Biorefinery: An Integrated Approach, D.
Das (ed.), Springer International Publishing, New York, pp. 103–124.
Kiran B., Pathak K., Kumar R. and Deshmukh D. (2014). Cultivation of chlorella sp. IM-01 in municipal wastewater
for simultaneous nutrient removal and energy feedstock production. Ecological Engineering, 73, 326–330,
https://doi.org/10.1016/j.ecoleng.2014.09.094
Larimer F. W., Chain P., Hauser L., Lamerdin J., Malfatti S., Do L., Land M. L., Pelletier D. A., Beatty J. T., Lang
A. S., Tabita F. R., Gibson J. L., Hanson T. E., Bobst C., Torres Y Torres J. L., Peres C., Harrison F. H., Gibson
J. and Harwood C. S. (2004). Complete genome sequence of the metabolically versatile photosynthetic
bacterium Rhodopseudomonas palustris. Nature Biotechnology, 22, 55–61, https://doi.org/10.1038/nbt923
Leong W. H., Lim J. W., Lam M. K., Uemura Y., Ho C. D. and Ho Y. C. (2018). Co-cultivation of activated sludge
and microalgae for the simultaneous enhancements of nitrogen-rich wastewater bioremediation and lipid
production. Journal of the Taiwan Institute of Chemical Engineers, 87, 216–224, https://doi.org/10.1016/j.
jtice.2018.03.038
Liu S., Zhang G., Zhang J., Li X. and Li J. (2016). Performance, carotenoids yield and microbial population dynamics
in a photobioreactor system treating acidic wastewater: effect of hydraulic retention time (HRT) and organic
loading rate (OLR). Bioresource Technology, 200, 245–252, https://doi.org/10.1016/j.biortech.2015.10.044
Liu S., Daigger G. T., Kang J. and Zhang G. (2019). Effects of light intensity and photoperiod on pigments
production and corresponding key gene expression of Rhodopseudomonas palustris in a photobioreactor
system. Bioresource Technology, 294, 122172, https://doi.org/10.1016/j.biortech.2019.122172
Lo K. J., Lin S. S., Lu C. W., Kuo C. H. and Liu C. T. (2018). Whole-genome sequencing and comparative analysis of
two plant-associated strains of Rhodopseudomonas palustris (PS3 and YSC3). Scientific Reports, 8, 12769,
https://doi.org/10.1038/s41598-018-31128-8
López J. C., Quijano G., Souza T. S. O., Estrada J. M., Lebrero R. and Muñoz R. (2013). Biotechnologies for
greenhouse gases (CH4, N2O, and CO2) abatement: state of the art and challenges. Applied Microbiology and
Biotechnology, 97, 2277–2303, https://doi.org/10.1007/s00253-013-4734-z
López-Serna R., Posadas E., García-Encina P. A. and Muñoz R. (2019). Removal of contaminants of emerging
concern from urban wastewater in novel algal-bacterial photobioreactors. Science of the Total Environment,
662, 32–40, https://doi.org/10.1016/j.scitotenv.2019.01.206
Lu H., Zhang G. and Dong S. (2011). Quantitative study of PNSB energy metabolism in degrading pollutants under
weak light-micro oxygen condition. Bioresource Technology, 102, 4968–4973, https://doi.org/10.1016/j.
biortech.2011.01.027
Lu H., Zhang G., Dai X., Schideman L., Zhang Y., Li B. and Wang H. (2013). A novel wastewater treatment
and biomass cultivation system combining photosynthetic bacteria and membrane bioreactor technology.
Desalination, 322, 176–181, https://doi.org/10.1016/j.desal.2013.05.007
Lu H., Peng M., Zhang G., Li B. and Li Y. (2019a). Brewery wastewater treatment and resource recovery through
long term continuous-mode operation in pilot photosynthetic bacteria-membrane bioreactor. Science of the
Total Environment, 646, 196–205, https://doi.org/10.1016/j.scitotenv.2018.07.268
Lu H., Zhang G., Zheng Z., Meng F., Du T. and He S. (2019b). Bio-conversion of photosynthetic bacteria from
non-toxic wastewater to realize wastewater treatment and bioresource recovery: A review. Bioresource
Technology, 278, 383–399, https://doi.org/10.1016/j.biortech.2019.01.070
Madukasi E. I. and Zhang G. (2010). Microaerobic biodegradation of high organic load wastewater by phototrophic
bacteria. African Journal of Biotechnology, 9(25), 3852–3860.
Mahdy A., Mendez L., Ballesteros M. and González-Fernández C. (2014a). Autohydrolysis and alkaline
pretreatment effect on chlorella vulgaris and scenedesmus sp. Methane production. Energy, 78, 48–52,
https://doi.org/10.1016/j.energy.2014.05.052
Mahdy A., Mendez L., Blanco S., Ballesteros M. and González-Fernández C. (2014b). Protease cell wall degradation
of chlorella vulgaris: effect on methane production. Bioresource Technology, 171, 421–427, https://doi.
org/10.1016/j.biortech.2014.08.091
Maity J. P., Bundschuh J., Chen C. Y. and Bhattacharya P. (2014). Microalgae for third generation biofuel production,
mitigation of greenhouse gas emissions and wastewater treatment: present and future perspectives – A mini
review. Energy, 78, 104–113, https://doi.org/10.1016/j.energy.2014.04.003
Mann G., Schlegel M., Schumann R. and Sakalauskas A. (2009). Biogas-conditioning with microalgae. Agronomy
Research, 7(1), 33–38.
Marella T. K., Datta A., Patil M. D., Dixit S. and Tiwari A. (2019). Biodiesel production through algal cultivation
in urban wastewater using algal floway. Bioresource Technology, 280, 222–228, https://doi.org/10.1016/j.
biortech.2019.02.031
Marín D., Posadas E., Cano P., Pérez V., Blanco S., Lebrero R. and Muñoz R. (2018). Seasonal variation of biogas
upgrading coupled with digestate treatment in an outdoors pilot scale algal-bacterial photobioreactor.
Bioresource Technology, 263, 58–66, https://doi.org/10.1016/j.biortech.2018.04.117
Marín D., Ortíz A., Díez-Montero R., Uggetti E., García J., Lebrero R. and Muñoz R. (2019a). Influence of
liquid-to-biogas ratio and alkalinity on the biogas upgrading performance in a demo scale algal-bacterial
photobioreactor. Bioresource Technology, 280, 112–117, https://doi.org/10.1016/j.biortech.2019.02.029
Marín D., Posadas E., García D., Puyol D., Lebrero R. and Muñoz R. (2019b). Assessing the potential of purple
phototrophic bacteria for the simultaneous treatment of piggery wastewater and upgrading of biogas.
Bioresource Technology, 281, 10–17, https://doi.org/10.1016/j.biortech.2019.02.073
Marín D., Carmona-Martínez A. A., Blanco S., Lebrero R. and Muñoz R. (2021). Innovative operational strategies
in photosynthetic biogas upgrading in an outdoors pilot scale algal-bacterial photobioreactor. Chemosphere,
264, 128470, https://doi.org/10.1016/j.chemosphere.2020.128470
Markou G., Vandamme D. and Muylaert K. (2014). Microalgal and cyanobacterial cultivation: the supply of
nutrients. Water Research, 65, 186–202, https://doi.org/10.1016/j.watres.2014.07.025
Masojídek J., Torzillo G. and Koblízek M. (2013). Photosynthesis in microalgae. In: Handbook of Microalgal
Culture: Applied Phycology and Biotechnology, A. Richmond and Q. Hu (eds.), John Wiley & Sons, Ltd.,
New Jersey, pp. 21–36.
Meier L., Pérez R., Azócar L., Rivas M. and Jeison D. (2015). Photosynthetic CO2 uptake by microalgae: An
attractive tool for biogas upgrading. Biomass and Bioenergy, 73, 102–109, https://doi.org/10.1016/j.
biombioe.2014.10.032
Meier L., Stará D., Bartacek J. and Jeison D. (2018). Removal of H 2S by a continuous microalgae-based
photosynthetic biogas upgrading process. Process Safety and Environmental Protection, 119, 65–68, https://
doi.org/10.1016/j.psep.2018.07.014
Mendez L., Mahdy A., Ballesteros M. and González-Fernández C. (2015). Biomethane production using fresh and
thermally pretreated chlorella vulgaris biomass: A comparison of batch and semi-continuous feeding mode.
Ecological Engineering, 84, 273–277, https://doi.org/10.1016/j.ecoleng.2015.09.056
Mendez L., Mahdy A., Ballesteros M. and González-Fernández C. (2014a). Methane production of thermally
pretreated chlorella vulgaris and scenedesmus sp. Biomass at increasing biomass loads. Applied Energy, 129,
238–242, https://doi.org/10.1016/j.apenergy.2014.04.110
Mendez L., Mahdy A., Demuez M., Ballesteros M. and González-Fernández C. (2014b). Effect of high pressure
thermal pretreatment on Chlorella vulgaris biomass: organic matter solubilisation and biochemical methane
potential. Fuel, 117, 674–679, https://doi.org/10.1016/j.fuel.2013.09.032
Mendez L., Sialve B., Tomás-Pejó E., Ballesteros M., Steyer J. P. and González-Fernández C. (2016). Comparison
of chlorella vulgaris and cyanobacterial biomass: cultivation in urban wastewater and methane production.
Bioprocess and Biosystems Engineering, 39, 703–712, https://doi.org/10.1007/s00449-016-1551-7
Mohd Udaiyappan A. F., Abu Hasan H., Takriff M. S. and Sheikh Abdullah S. R. (2017). A review of the potentials,
challenges and current status of microalgae biomass applications in industrial wastewater treatment. Journal
of Water Process Engineering, 20, 8–21, https://doi.org/10.1016/j.jwpe.2017.09.006
Molazadeh M., Ahmadzadeh H., Pourianfar H. R., Lyon S. and Rampelotto P. H. (2019). The use of microalgae
for coupling wastewater treatment with CO2 biofixation. Frontiers in Bioengineering and Biotechnology,
7(48), 1–12.
Molina-Grima E. M., Fernández Sevilla J. M. and Acién Fernández F. G. (2010). Microalgae, mass culture methods.
In: Encyclopedia of Industrial Biotechnology, M. C. Flickinger (ed.), Wiley, New Jersey, pp. 1–24.
Molinuevo-Salces B., González-Fernández C., Gómez X., García-González M. C. and Morán A. (2012). Vegetable
processing wastes addition to improve swine manure anaerobic digestion: evaluation in terms of methane
yield and SEM characterization. Applied Energy, 91, 36–42, https://doi.org/10.1016/j.apenergy.2011.09.010
Molinuevo-Salces B., Mahdy A., Ballesteros M. and González-Fernández C. (2016). From piggery wastewater
nutrients to biogas: microalgae biomass revalorization through anaerobic digestion. Renewable Energy, 96,
1103–1110, https://doi.org/10.1016/j.renene.2016.01.090
Morales-Amaral M. del M., Gómez-Serrano C., Acién F. G., Fernández-Sevilla J. M. and Molina-Grima E. (2015).
Production of microalgae using centrate from anaerobic digestion as the nutrient source. Algal Research, 9,
297–305, https://doi.org/10.1016/j.algal.2015.03.018
Mulbry W., Westhead E. K., Pizarro C. and Sikora L. (2005). Recycling of manure nutrients: use of algal biomass
from dairy manure treatment as a slow release fertilizer. Bioresource Technology, 96(4), 451–458, https://doi.
org/10.1016/j.biortech.2004.05.026
Mulbry W., Kangas P. and Kondrad S. (2010). Toward scrubbing the bay: nutrient removal using small algal turf
scrubbers on chesapeake Bay tributaries. Ecological Engineering, 36, 536–541, https://doi.org/10.1016/j.
ecoleng.2009.11.026
Muñoz R., Meier L., Díaz I., Jeison D., Diaz I. and Jeison D. (2015). A review on the state-of-the-art of physical/
chemical and biological technologies for biogas upgrading. Reviews in Environmental Science and
Biotechnology, 14, 727–759, https://doi.org/10.1007/s11157-015-9379-1
Murry M. A., Murinda S. E., Huang S. and Mark A. (2019). Bioconversion of agricultural wastes from the livestock
industry for biofuel and feed production. In: Advanced Bioprocessing for Alternative Fuels, Biobased
Chemicals, and Bioproducts, M. Hosseini (ed.), Elsevier Inc., Amsterdam, pp. 225–247.
Nagadomi H., Kitamura T., Watanabe M. and Sasaki K. (2000). Simultaneous removal of chemical
oxygen demand (COD), phosphate, nitrate and H 2S in the synthetic sewage wastewater using porous
ceramic immobilized photosynthetic bacteria. Biotechnology Letters, 22(17), 1369–1374, https://doi.
org/10.1023/A:1005688229783
Navid A., Jiao Y., Wong S. E. and Pett-Ridge J. (2019). System-level analysis of metabolic trade-offs during anaerobic
photoheterotrophic growth in Rhodopseudomonas palustris. BMC Bioinformatics, 20, 233, https://doi.
org/10.1186/s12859-019-2844-z
Nkemka V. N. and Murto M. (2013). Biogas production from wheat straw in batch and UASB reactors: the roles
of pretreatment and seaweed hydrolysate as a co-substrate. Bioresource Technology, 128, 164–172, https://
doi.org/10.1016/j.biortech.2012.10.117
Olguín E. J., Sánchez-Galván G. and Pérez-Pérez T. (2007). Assessment of the phytoremediation potential of
Salvinia minima Baker compared to Spirodela polyrrhiza in high-strength organic wastewater. Water Air
and Soil Pollution, 181(1), 135–147, https://doi.org/10.1007/s11270-006-9285-9
Oswald W. J. (1978). The engineering aspects of microalgae. In: CRC Handbook of Microbiology, 2nd edn. (II)
Fungi, Algae, Protozoa, and Viruses, A. I. Laskin and H. A. Lechevalier (eds.), CRC Press, West Palm Beach,
Florida, pp. 519–552.
Pancha I., Chokshi K., George B., Ghosh T., Paliwal C., Maurya R. and Mishra S. (2014). Nitrogen stress triggered
biochemical and morphological changes in the microalgae Scenedesmus sp. CCNM 1077. Bioresource
Technology, 156, 146–154, https://doi.org/10.1016/j.biortech.2014.01.025
Park J. B. K., Craggs R. J. and Shilton A. N. (2011). Wastewater treatment high rate algal ponds for biofuel
production. Bioresource Technology, 102, 35–42, https://doi.org/10.1016/j.biortech.2010.06.158
Park J. B. K., Craggs R. J. and Shilton A. N. (2013). Enhancing biomass energy yield from pilot-scale high rate algal
ponds with recycling. Water Research, 47, 4422–4432, https://doi.org/10.1016/j.watres.2013.04.001
Passos F., Solé M., García J. and Ferrer I. (2013). Biogas production from microalgae grown in wastewater: effect
of microwave pretreatment. Applied Energy, 108, 168–175, https://doi.org/10.1016/j.apenergy.2013.02.042
Passos F., Uggetti E., Carrère H. and Ferrer I. (2014). Pretreatment of microalgae to improve biogas production: a
review. Bioresource Technology, 172, 403–412, https://doi.org/10.1016/j.biortech.2014.08.114
Passos F., Carretero J. and Ferrer I. (2015a). Comparing pretreatment methods for improving microalgae anaerobic
digestion: thermal, hydrothermal, microwave and ultrasound. Chemical Engineering Journal, 279, 667–672,
https://doi.org/10.1016/j.cej.2015.05.065
Passos F., Uggetti E., Carrère H. and Ferrer I. (2015b). Algal biomass: physical pretreatments. In: Pretreatment of
Biomass, A. Pandey, S. Negi, P. Binod and C. Larroche (eds.), Elsevier, Amsterdam, pp. 195–226.
Pawlowski A., Mendoza J. L., Guzmán J. L., Berenguel M., Acién F. G. and Dormido S. (2015). Selective pH and
dissolved oxygen control strategy for a raceway reactor within an event-based approach. Control Engineering
Practice, 44, 209–218, https://doi.org/10.1016/j.conengprac.2015.08.004
Posadas E., García-Encina P. A., Soltau A., Domínguez A., Díaz I. and Muñoz R. (2013). Carbon and nutrient
removal from centrates and domestic wastewater using algal-bacterial biofilm bioreactors. Bioresource
Technology, 139, 50–58, https://doi.org/10.1016/j.biortech.2013.04.008
Posadas E., Bochon S., Coca M., García-González M. C., García-Encina P. A. and Muñoz R. (2014). Microalgae-
based agro-industrial wastewater treatment: a preliminary screening of biodegradability. Journal of Applied
Phycology, 26, 2335–2345, https://doi.org/10.1007/s10811-014-0263-0
Posadas E., Morales M. del M., Gomez C., Acién F. G. G. and Muñoz R. (2015a). Influence of pH and CO2
source on the performance of microalgae-based secondary domestic wastewater treatment in outdoors pilot
raceways. Chemical Engineering Journal, 265, 239–248, https://doi.org/10.1016/j.cej.2014.12.059
Posadas E., Serejo M. L., Blanco S., Pérez R., García-Encina P. A. and Muñoz R. (2015b). Minimization of
biomethane oxygen concentration during biogas upgrading in algal-bacterial photobioreactors. Algal
Research, 12, 221–229, https://doi.org/10.1016/j.algal.2015.09.002
Posadas E., Alcántara C., García-Encina P. A., Gouveia L., Guieysse B., Norvill Z., Acién F. G., Markou G.,
Congestri R., Koreiviene J. and Muñoz R. (2017a). Microalgae cultivation in wastewater. In: Microalgae-
Based Biofuels Bioproducts From Feedstock Cultivation to End-Products, R. Raul Muñoz and C. Gonzalez-
Fernandez (eds.), Elsevier, Amsterdam, pp. 67–91.
Posadas E., Marín D., Blanco S., Lebrero R. and Muñoz R. (2017b). Simultaneous biogas upgrading and centrate
treatment in an outdoors pilot scale high rate algal pond. Bioresource Technology, 232, 133–141, https://doi.
org/10.1016/j.biortech.2017.01.071
Puyol D., Monsalvo V. M., Marin E., Rogalla F., Melero J. A., Martínez F., Hülsen T. and Batstone D. J. (2020). Purple
phototrophic bacteria as a platform to create the next generation of wastewater treatment plants: energy and
resource recovery. In: Wastewater Treatment Residues as Resources for Biorefinery Products and Biofuels, J. A.
Olivares, D. Puyol, J. A. Melero and J. Dufour (eds.), Elsevier, Amsterdam, Netherlands, pp. 255–280.
Qiao S., Hou C., Wang X. and Zhou J. (2020). Minimizing greenhouse gas emission from wastewater treatment
process by integrating activated sludge and microalgae processes. Science of the Total Environment, 732,
139032, https://doi.org/10.1016/j.scitotenv.2020.139032
Rajesh Banu J., Preethi Kavitha S., Gunasekaran M. and Kumar G. (2020). Microalgae based biorefinery promoting
circular bioeconomy-techno economic and life-cycle analysis. Bioresource Technology, 302, 122822, https://
doi.org/10.1016/j.biortech.2020.122822
Rani R. U., Kumar S. A., Kaliappan S., Yeom I.-T., Banu J. R., Uma Rani R., Adish Kumar S., Kaliappan S., Yeom
I.-T. and Rajesh Banu J. (2012). Low temperature thermo-chemical pretreatment of dairy waste activated
sludge for anaerobic digestion process. Bioresource Technology, 103, 415–424, https://doi.org/10.1016/j.
biortech.2011.09.124
Rezvani S., Moheimani N. R. and Bahri P. A. (2016). Techno-economic assessment of CO2 bio-fixation using
microalgae in connection with three different state-of-the-art power plants. Computers & Chemical
Engineering, 84, 290–301, https://doi.org/10.1016/j.compchemeng.2015.09.001
Rochaix J. D. (2016). The dynamics of the photosynthetic apparatus in algae. In: Applied Photosynthesis – New
Progress, M. Najafpour (ed.), IntechOpen, London, pp. 23–52.
Rodero M. del R., Posadas E., Toledo-Cervantes A., Lebrero R. and Muñoz R. (2018). Influence of alkalinity and
temperature on photosynthetic biogas upgrading efficiency in high rate algal ponds. Algal Research, 33,
284–290, https://doi.org/10.1016/j.algal.2018.06.001
Rodero M. del R., Lebrero R., Serrano E., Lara E., Arbib Z., García-Encina P. A. A. and Muñoz R. (2019). Technology
validation of photosynthetic biogas upgrading in a semi-industrial scale algal-bacterial photobioreactor.
Bioresource Technology, 279, 43–49, https://doi.org/10.1016/j.biortech.2019.01.110
Rodero M. del R., Carvajal A., Arbib Z., Lara E., de Prada C., Lebrero R., Muñoz R., Rodero R., Carvajal A.,
Arbib Z., Lara E. and Prada C. D. (2020a). Performance evaluation of a control strategy for photosynthetic
biogas upgrading in a semi-industrial scale photobioreactor. Bioresource Technology, 307, 123207, https://
doi.org/10.1016/j.biortech.2020.123207
Rodero M. del R., Severi C. A., Rocher-Rivas R., Quijano G. and Muñoz R. (2020b). Long-term influence of
high alkalinity on the performance of photosynthetic biogas upgrading. Fuel, 281, 118804, https://doi.
org/10.1016/j.fuel.2020.118804
Ruiz J., Arbib Z., Barragán J. and Perales J. A. (2012). Performance of a flat panel reactor in the continuous culture
of microalgae in urban wastewater: prediction from a batch experiment. Bioresource Technology, 127, 456–
463, https://doi.org/10.1016/j.biortech.2012.09.103
Ryckebosch E., Drouillon M. and Vervaeren H. (2011). Techniques for transformation of biogas to biomethane.
Biomass and Bioenergy, 35, 1633–1645, https://doi.org/10.1016/j.biombioe.2011.02.033
Samson R. and Leduy A. (1985). Multistage continuous cultivation of blue-green alga spirulina maxima in the flat
tank photobioreactors with recycle. The Canadian Journal of Chemical Engineering, 63, 105–112, https://
doi.org/10.1002/cjce.5450630117
Sanz J. L., Rojas P., Morato A., Mendez L., Ballesteros M. and González-Fernández C. (2017). Microbial communities
of biomethanization digesters fed with raw and heat pre-treated microalgae biomasses. Chemosphere, 168,
1013–1021, https://doi.org/10.1016/j.chemosphere.2016.10.109
Sarkodie S. A., Adams S. and Leirvik T. (2020). Foreign direct investment and renewable energy in climate change
mitigation : does governance matter? Journal of Cleaner Production, 263, 121262, https://doi.org/10.1016/j.
jclepro.2020.121262
Schediwy K., Trautmann A., Steinweg C. and Posten C. (2019). Microalgal kinetics – a guideline for photobioreactor
design and process development. Engineering in Life Sciences, 19, 830–843, https://doi.org/10.1002/
elsc.201900107
Schwede S., Rehman Z.-U., Gerber M., Theiss C. and Span R. (2013). Effects of thermal pretreatment on
anaerobic digestion of nannochloropsis salina biomass. Bioresource Technology, 143, 505–511, https://doi.
org/10.1016/j.biortech.2013.06.043
Sepúlveda-Muñoz C. A., Ángeles R., de Godos I. and Muñoz R. (2020a). Comparative evaluation of continuous
piggery wastewater treatment in open and closed purple phototrophic bacteria-based photobioreactors.
Journal of Water Process Engineering, 38, 101608, https://doi.org/10.1016/j.jwpe.2020.101608
Sepúlveda-Muñoz C. A., de Godos I., Puyol D. and Muñoz R. (2020b). A systematic optimization of piggery
wastewater treatment with purple phototrophic bacteria. Chemosphere, 253, 126621, https://doi.
org/10.1016/j.chemosphere.2020.126621
Serejo M. L., Posadas E., Boncz M. A., Blanco S., Garcia-Encina P. and Muñoz R. (2015). Influence of biogas flow
rate on biomass composition during the optimization of biogas upgrading in microalgal-bacterial processes.
Environmental Science & Technology, 49, 3228–3236, https://doi.org/10.1021/es5056116
Singh J. and Dhar D. W. (2019). Overview of carbon capture technology: microalgal biorefinery concept and state-
of-the-art. Frontiers in Marine Science, 6(29), 1–10. https://doi.org/10.3389/fmars.2019.00029
Solovchenko A., Verschoor A. M., Jablonowski N. D. and Nedbal L. (2016). Phosphorus from wastewater to crops:
An alternative path involving microalgae. Biotechnology Advances, 34, 550–564, https://doi.org/10.1016/j.
biotechadv.2016.01.002
Spellman F. R. and Drinan J. E. (2014). Wastewater Stabilization Ponds. CRC Press, Boca Raton.
Syahirah N., Aron M., Shiong K., Wayne K., Veeramuthu A., Chang J. and Loke P. (2020). Microalgae cultivation
in wastewater and potential processing strategies using solvent and membrane separation technologies.
Journal of Water Process Engineering, 39, 101701.
Toledo-Cervantes A., Serejo M. L., Blanco S., Pérez R., Lebrero R. and Muñoz R. (2016). Photosynthetic biogas
upgrading to bio-methane: boosting nutrient recovery via biomass productivity control. Algal Research, 17,
46–52, https://doi.org/10.1016/j.algal.2016.04.017
Toledo-Cervantes A., Madrid-Chirinos C., Cantera S., Lebrero R. and Muñoz R. (2017). Influence of the gas-
liquid flow configuration in the absorption column on photosynthetic biogas upgrading in algal-bacterial
photobioreactors. Bioresource Technology, 225, 336–342, https://doi.org/10.1016/j.biortech.2016.11.087
Toledo-Cervantes A., Morales T., González Á, Muñoz R. and Lebrero R. (2018). Long-term photosynthetic
CO2 removal from biogas and flue-gas: exploring the potential of closed photobioreactors for high-value
biomass production. Science of the Total Environment, 640–641, 1272–1278, https://doi.org/10.1016/j.
scitotenv.2018.05.270
Tredici M. R. and Materassi R. (1992). From open ponds to vertical alveolar panels: the Italian experience in
the development of reactors for the mass cultivation of phototrophic microorganisms. Journal of Applied
Phycology, 4, 221–231, https://doi.org/10.1007/BF02161208
Uggetti E., Passos F., Solé M., Garfí M. and Ferrer I. (2017). Recent achievements in the production of biogas from
microalgae. Waste and Biomass Valorization, 8, 129–139, https://doi.org/10.1007/s12649-016-9604-3
UNE-EN 16723 (2017). Natural Gas and Biomethane for Use in Transport and Biomethane for Injection in the
Natural Gas Network. Asociación Española de Normalización. https://www.une.org/encuentra-tu-norma/
busca-tu-norma/norma/?c=N0058455 (Accessed 16 February 2022).
Wen S., Liu H., He H., Luo L., Li X., Zeng G., Zhou Z., Lou W. and Yang C. (2016). Treatment of anaerobically
digested swine wastewater by Rhodobacter blasticus and Rhodobacter capsulatus. Bioresource Technology,
222, 33–38, https://doi.org/10.1016/j.biortech.2016.09.102
Whitton R., Ometto F., Pidou M., Jarvis P., Villa R. and Jefferson B. (2015). Microalgae for municipal wastewater
nutrient remediation: mechanisms, reactors and outlook for tertiary treatment. Environmental Technology
Reviews, 4(1), 133–148, https://doi.org/10.1080/21622515.2015.1105308
Woertz I., Feffer A., Lundquist T. and Nelson Y. (2009). Algae grown on dairy and municipal wastewater for
simultaneous nutrient removal and lipid production for biofuel feedstock. Journal of Environmental
Engineering, 135, 1115–1122, https://doi.org/10.1061/(ASCE)EE.1943-7870.0000129
Yang J., Cao J., Xing G. and Yuan H. (2015). Lipid production combined with biosorption and bioaccumulation
of cadmium, copper, manganese and zinc by oleaginous microalgae Chlorella minutissima UTEX2341.
Bioresource Technology, 175, 537–544, https://doi.org/10.1016/j.biortech.2014.10.124
Yang A., Zhao W., Peng M., Zhang G., Zhi R. and Meng F. (2018). A special light-aerobic condition for photosynthetic
bacteria-membrane bioreactor technology. Bioresource Technology, 268, 820–823, https://doi.org/10.1016/j.
biortech.2018.08.008
Zhou Q., Zhang P. and Zhang G. (2015). Biomass and pigments production in photosynthetic bacteria wastewater
treatment: effects of light sources. Bioresource Technology, 179, 505–509, https://doi.org/10.1016/j.
biortech.2014.12.077
Zhou Q., Zhang G., Lu Y. and Wu P. (2016). Feasibility study and process optimization of citric acid wastewater
treatment and biomass production by photosynthetic bacteria. Desalination and Water Treatment, 57(14),
6261–6267, https://doi.org/10.1080/19443994.2015.1005687
Chapter 10
Sludge management and utilization
for decarbonization
Meltem Urgun-Demirtas1*, Rachel Dalke1 and Krishna R. Pagilla2
1Applied Materials Division, Argonne National Laboratory, 9700 S Cass Ave, Lemont, IL, 60439, USA
2Civil and Environmental Engineering, University of Nevada, Reno, 1664 N Virginia Street, Reno, NV, 89557, USA
*Correspondence: [email protected]
© 2022 The Editors. This is an Open Access eBook distributed under the terms of the Creative Commons Attribution Licence
(CC BY-NC-ND 4.0), which permits copying and redistribution for non-commercial purposes with no derivatives, provided the
original work is properly cited (https://creativecommons.org/licenses/by-nc-nd/4.0/). This does not affect the rights licensed or
assigned from any third party in this book. The chapter is from the book Pathways to Water Sector Decarbonization, Carbon Capture
and Utilization, Zhiyong Jason Ren and Krishna Pagilla (Eds.)
d Eurostat (2020).
a substantial amount of sludge each year (>500 000 tons/yr from 2009 to 2018) that was primarily
managed through landfilling (Eurostat, 2020; Rosiek, 2020). In 2016 a law was passed prohibiting
landfilling of sludge; thereby creating a need for the development of adequate alternative management
strategies. In 2016, up to 34 and 30% of sludge generated in Poland was managed via agricultural
use (soil formation, fertilizer) and thermal transformation (incineration, co-incineration, gasification,
pyrolysis, wet oxidation), respectively (Przydatek & Wota, 2020).
The objective of this chapter is to review and summarize current knowledge about biosolids
management and provide a basis for future practices. The chapter focuses on how current management
practices help decarbonization, the role of biosolids management strategies in achieving the
decarbonization targets of utilities, and how challenges (e.g., emerging contaminants, odors, public
scrutiny and upset) in management can be addressed in meeting such targets. Since Chapter XX
Figure 10.1 Percent of resident population connected to urban wastewater collecting system. (OECD.Stat 2021).
discusses AD application in detail, this chapter focuses on sludge treatment alternatives beyond AD. It
should be noted that residual digestate after AD of sludge also needs additional treatment and disposal.
New emerging concepts, namely Water-Energy Nexus, Circular Economy, and Nutrient Trading, are
important vehicles for decarbonization in shaping the future sludge management practices. These
concepts significantly help to reduce the financial burdens of sludge management on societies and
overcome ecological issues and resource scarcity. New technologies and approaches need to be
developed to extract energy and nutrients from sludge and improve process and energy efficiency.
Recovered energy and nutrients help utilities become a source of revenue generation, overcoming
their reputation as pollution mitigation entities. In turn, they will become entities contributing to
reduction in carbon emissions and achieve decarbonization of the water sector. Renewable energy
production and resource recovery are presented as areas of sludge management to close the linearity
of waste production and implement a circular economy of waste management. The chapter closes with
a section on implementation of decarbonization at the utilities and future strategies and pathways.
• Decreases the need for chemical fertilizers • Compressed mixtures have lower be mitigated
(Bruni et al., 2020) porosity and moisture content and cause
• Contributes to awareness and promotion unpleasant odors (Jędrczak, 2018)
of community-level waste management • Produces higher amount of greenhouse
practices (Bruni et al., 2020) gas emissions (Jędrczak, 2018)
Sludge management and utilization for decarbonization 175
Figure 10.2 Sludge management practice utilization in the United States, European Union, and China (Đurđević
et al., 2019; US EPA, 2019; Wei et al., 2020).
et al., 2021). However, it has been reported in the US that biogas from sludge processing has a low level
of utilization with less than half of WRRFs with influent flow rates above one million gallons per day
(MGD) or 3785 m3/day using AD, and few using the biogas for heat or electricity generation (Seiple
et al., 2020; Shen et al., 2015). Additionally, growing legal use or disposal requirements for the solid
products of common biological methods such as AD or composting complicate their application; thus,
efforts are being focused on alternative methods (Świerczek et al., 2021).
Sludge incineration in conjunction with phosphorus (P) recovery has been gaining a lot of attention.
In Germany, large wastewater treatment plants have been using sludge incineration to recover P in
ashes to meet the country’s goal of obtaining at least 20% of raw phosphate from sewage sludge.
Total authorized sludge incineration capacity in Germany is approximately 1.5 M dry metric tons/yr
and 80% of this capacity in use (Wiechmann et al., 2013). Co-incineration of sludge at coal/lignite
fired power plants and cement plants has many advantages including reduced fossil fuel and carbon
emissions and lowered cost; however, P is not recoverable during the process. Sludge incineration
has been also used for the production of building materials such as cement. This method is one of
the best options if sewage contains a high concentration of heavy metals (HM) since it transforms
bioavailable HMs to more stable forms, thereby alleviating their leaching toxicities into the receiving
environments (Cao et al., 2020). Although sludge incineration maximizes solids reduction (thereby
lowering land requirements) and yields energy recovery with a stable ash production, it has high
capital and operation costs. Toxicity of the ash also limits the overall feasibility of the management
strategy (Arias et al., 2021). In most cases, air emissions are an important issue and require additional
treatment to meet air quality specifications.
An alternative process for treating raw biosolids is pyrolysis, an oxygen-deficient thermal
degradation process. Pyrolysis is an inexpensive and robust procedure that has shown promise as
a treatment tool for raw biosolids by generating a liquid and solid stream that can be used for liquid
fuels and soil amendment, respectively (Callegari & Capodaglio, 2018). Thermal hydrolysis prior to
(an)aerobic digestion or composting has been reported to increase feedstock biodegradability, solids
loading capability (9 vs. 6%), and biosolid dewaterability, all while maintaining a better energy
balance (Flores-Alsina et al., 2021). The high temperatures and pressures of pyrolysis improve the
soil quality through the inactivation of pathogens, thereby expanding the applicability of the finished
product (landfill vs. land application) (Flores-Alsina et al., 2021). Biochar is one product generated
through pyrolysis that has garnered attention from researchers as a means to improve soil quality and
enhance nutrient levels. Applications of biochar are seen in adsorption of antibiotics, heavy metals,
dyes, and phenolic compounds in wastewater effluent. It also aids in N retainment in soils after
agricultural application of the product (Singh et al., 2020). Pyrolysis has also been utilized to convert
biosolids into granular activated carbon (GAC). Producing GAC via pyrolysis is reported to provide a
financial benefit by reducing facility operating costs by 20–40% (Mu’azu et al., 2019). However, GAC
has reportedly less effective adsorption of heavy metals (Singh et al., 2020).
Recently the efficacy of implementing thermal treatment methods has been studied; however, their
current utilization is limited due to economic limitations. It has been shown that thermal treatment
methods have benefits within the framework of a circular economy, allowing for energy and natural
resource recovery, as well as valuable byproduct generation (Tsybina & Wuensch, 2018). A model
study conducted in South Africa found using a centralized sludge management technique with
thermal pretreatment and AD resulted in 36.5% conversion of COD in the influent into methane and
41 and 65% N and P in the biosolids, respectively. This is a substantial improvement from the disposal
practice utilized in the study, which was landfilling (Flores-Alsina et al., 2021).
Hydrothermal liquefaction (HTL) is another process that has the potential to treat and valorize
biosolids via conversion into biocrude oil. A significant advantage to utilizing HTL over pyrolysis
is the capability to process wet feedstocks, thereby eliminating the need for a cost-intensive drying
step. A modeling study in the US determined that facilities with capacity greater than 17 400 m3
per day could supply approximately 10 million metric tons of feedstock and produce 3.7 million m3
of biocrude/year (Seiple et al., 2020). It has been shown that the type of feedstock processed and
operation conditions selected influence the elemental composition of biocrude. WRRF biosolids have
been shown to produce high biocrude yield (45%), as well as 55–80% carbon recovery in the oil product.
Additionally, it has been shown that the majority of inorganics (including P) become concentrated in
the solids product of post-HTL (weight percentages >70%), where they may be separated out and used
as fertilizer (Conti et al., 2020). HTL has shown promise as a technology suitable for the destruction
of micropollutants in sludge; however, the complex composition of the resultant process water limits
its scale-up applications (Silva Thomsen et al., 2020). While raw process water has been identified as
a health hazard due to a high concentration of toxic compounds, treated effluent has the potential
for valorization through energy and nutrient recovery (Watson et al., 2020). While HTL is a highly
efficient strategy for biomass conversion into fuels, process applications are hindered by expensive
upgrading techniques required to meet biofuel standards. Catalysts have been proposed as a means to
overcome this challenge by increasing biocrude quality, but many are non-recoverable and thus present
additional economic challenges. Heterogeneous catalysts have shown potential as a viable alternative
that reduce costs by not only increasing biocrude yield and quality but allowing for catalysis recovery
and reuse (Scarsella et al., 2020). Biocrude from HTL has a lower oxygen and moisture content
and higher heating value relative to pyrolysis, resulting in lower fixed and operational costs and a
subsequent competitive advantage as a bioconversion technology (Dimitriadis & Bezergianni, 2017).
Walker et al., 1994). Most land applications in the US are on agricultural land (pastures and cropland),
disturbed areas (e.g., brownfields), plant nurseries, forests, recreational areas (e.g., parks, golf courses),
lawns and gardens, cemeteries, highways, and airport runway medians (US EPA, 2003a, 2003b). The
use of sludge in agriculture within the EU is currently regulated only by the limits of heavy metals (Cd,
Cu, Hg, Ni, Pb and Zn) listed in Council Directive 86/278/EEC. Low doses of sludge applications have
shown beneficial effects on microbial biomass, organic carbon, and soil microbial activity. In some
cases, excessive application of sewage sludge with high heavy metals concentrations into soil has been
found to increase the bioavailability of heavy metals (Hudcová et al., 2019).
10.5 COMPOSTING
Composting is a biochemical process involving the degradation of organic matter by microorganisms
under natural or controlled conditions (Bruni et al., 2020; Onwosi et al., 2017; Sánchez et al.,
2017). Composting has been identified as a valuable waste stabilization technique due to its wide
environmental compatibility (Onwosi et al., 2017). The success of implementing composting is
dependent on the concentrations of heavy metals in the waste material. When properly managed,
composting reduces greenhouse gas emissions, increases soil fertility and biodiversity, and reduces the
need for chemical fertilizers (Bruni et al., 2020). However, challenges exist with leachate generation,
gas emissions, and lack of uniformity in compost modeling, which is used to determine how control
measures affect the overall composting process. Several methods have been applied to reduce the
negative effects of composting, including the addition of bulking agents such as sawdust, rice straw,
wood chips, and cotton gin waste (Onwosi et al., 2017). Biochar has also been used as a bulking agent
during composting, wherein the addition lowered concentrations of metals and arsenic in soil as well
as stronger adsorption and microbial community activity (Ye et al., 2019). A study in Poland showed
the maturation of composting of biosolids impacts the concentrations of organic carbon, nutrients,
and heavy metals (but not the percentage of their mobile or bioavailable forms) and is an effective
stabilization method (Boż ym & Siemiątkowski, 2018). Alongside Poland, Italy is making large strides
to implement improved biosolids management practices (Mininni et al., 2019). The country utilizes
numerous composting facilities within their waste management framework. Extensive efforts to
source-separate organic waste materials have helped to maximize recovery and meet the country’s
sustainability goals. Sewage sludge is the second main fraction in composting plants in Italy. However,
its percentage decreased from 17% in 2004 to 10.6% in 2017. Most of the compost products are used
in agriculture (approximately 70%) (Bruni et al., 2020).
An assessment of the environmental and economic sustainability of each sludge management option
including landfilling, composting, incineration, digestion, and land application is needed. Life Cycle
Assessment (LCA) and Life Cycle Cost Analysis (LCCA) have been used to provide a comparative
assessment of environmental and economic impacts of sludge management alternatives (Arias et al.,
2021; Lee et al., 2020; Yoshida et al., 2018).
Figure 10.3 Sludge production in different provinces of China in 2019 (Wei et al., 2020).
management strategies should also include the proper sludge treatment/disposal methods to minimize
greenhouse emissions in addition to pathogen inactivation and concerned contaminants removal.
China’s sludge management strategy dramatically changed recently. As a result of 18 standards and 12
regulations promulgated in China, land application of sludge reduced from 60.9% in 2009 to 21.9% in
2017 (Wei et al., 2020) (Figures 10.2 and 10.3).
Circular agreements and deals are emerging in the European Union as an alternative form of
governance to meet stakeholders’ needs across the value chain, hence increasing the circulation level of
materials. Through such agreements, individuals are provided an opportunity to redistribute risks and
responsibilities in ways more appropriate for achieving a circular economy (Johansson, 2021). Sweden
recently developed a certification program for wastewater treatment plants called REVAQ to alleviate
concerns about the land application of sludge for agricultural purposes. There is a continuous effort
to reduce concern contaminants in wastewater streams before intake to meet REVAQ certification
specifications. Approximately 45% of sludge produced in Sweden meets REVAQ quality standards
set for farm applications (Dagerskog & Olsson, 2020). Another example is the implementation of
the Green Deal in the Netherlands, which has promoted incinerator bottom ash as a construction
aggregate (Government of the Netherlands, 2016).
To meet increasingly strict quality requirements and alleviate P shortage, sludge incineration
for energy production with P recovery is becoming the most commonly applied process. In current
wastewater treatment applications, most P in wastewater has been captured in sludge as a result of the
chemical precipitation techniques used to meet discharge criterion for P removal. For decarbonization
of sludge management practices, it is crucial to close the nutrient recovery loop (Table 10.2). Nitrogen
recovery has significant potential in the decarbonization of wastewater treatment operations.
Recovered nitrogen can be used as a feedstock to produce fertilizers which require very energy-
intensive processes from fossil-based feedstocks. N2O emissions resulted from wastewater treatment
operations can be minimized or eliminated with this approach. However, nitrogen capture in sludge
streams is relatively small (∼25%) since a large portion of nitrogen ends up in treated water and
the atmosphere during the wastewater treatment operations. To maximize nitrogen recovery, there
is a need to identify nitrogen-rich streams. One nitrogen-rich stream is recycled water from sludge
dewatering processes. Separate collection and treatment of nitrogen-rich streams also provide
complete nitrogen recovery at decentralized plants. Nutrient recovery and materials production
(e.g. building/construction materials, adsorbents) with minimal environmental impacts compared to
their fossil-driven production routes are critical to lower GHG emissions, and hence are helpful in
achieving decarbonization targets.
Centralized wastewater treatment plants have been usually designed to treat gray water with
stormwater and industrial wastewater. This operation complicates the sludge management practices
because sludge quality and quantity depend on a wide range of parameters. Decentralized wastewater
treatment plants play a critical role in developing a closed nutrient recovery loop since they can be
easily designed to allow upstream source-separation of wastewater prior to intake. A new source-
separation approach has been developed in Sweden to separate and dry out the urine at the point
source, thereby diverting the bulk of nitrogen from sewer lines in a relatively easy way with minimum
retrofitting of the existing pipelines (Dagerskog & Olsson, 2020).
Various new alternative processes such as wet oxidation, hydrolysis, hydrothermal carbonization,
and supercritical water oxidation, currently at either embryonic stage or pilot-scale, will be scaled up
to large-scale applications. Implementation of such technologies should not only be looked at in terms
of economic and environmental feasibility at the specific site, but also in terms of carbon emissions
over the life cycle of sludge. In addition to LCA and LCCA, social impact analysis (e.g., employment,
income) will be widely used tools to guide the facilities in selecting the best sludge management
methods while focusing on decarbonization and sustainability goals. Useful products that recover
carbon from sludge in a sustainable manner to decarbonize the water sector is the ultimate goal.
REFERENCES
Abdel Wahaab R., Mahmoud M. and van Lier J. B. (2020). Toward achieving sustainable management of municipal
wastewater sludge in Egypt: the current status and future prospective. Renewable and Sustainable Energy
Reviews, 127, 109880, 1–11. https://doi.org/10.1016/j.rser.2020.109880
Appels L., Lauwers J., Degrève J., Helsen L., Lievens B., Willems K., Van Impe J. and Dewil R. (2011). Anaerobic
digestion in global bio-energy production: potential and research challenges. Renewable and Sustainable
Energy Reviews, 15(9), 4295–4301. https://doi.org/10.1016/j.rser.2011.07.121
Arias A., Feijoo G. and Moreira M. T. (2021). Benchmarking environmental and economic indicators of
sludge management alternatives aimed at enhanced energy efficiency and nutrient recovery. Journal of
Environmental Management, 279(1), 111594, 1–10. https://doi.org/10.1016/j.jenvman.2020.111594
Boż ym M. and Siemiątkowski G. (2018). Characterization of composted sewage sludge during the maturation
process: a pilot scale study. Environmental Science and Pollution Research, 25(34), 34332–34342. https://
doi.org/10.1007/s11356-018-3335-x
Brown S. and Henry C. (2001). Using Biosolids for Reclamation/Remediation of Disturbed Soils. University
of Washington, pp. 1–26. Available at https://www.epa.gov/sites/default/files/2015-05/documents/
biosolidswhitepaper-uwash.pdf (Accessed 23 February 2022).
Bruni C., Akyol Ç., Cipolletta G., Eusebi A. L., Caniani D., Masi S., Colón J. and Fatone F. (2020). Decentralized
community composting: past, present and future aspects of Italy. Sustainability, 12(8), 3319, 1–20. https://
doi.org/10.3390/su12083319
Callegari A. and Capodaglio A. (2018). Properties and beneficial uses of (bio)chars, with special attention to
products from sewage sludge pyrolysis. Resources, 7(1), 20, 1–22. https://doi.org/10.3390/resources7010020
Cao X., Ma R., Zhang Q., Wang W., Liao Q., Sun S., Zhang P. and Liu X. (2020). The factors influencing sludge
incineration residue (SIR)-based magnesium potassium phosphate cement and the solidification/stabilization
characteristics and mechanisms of heavy metals. Chemosphere, 261, 127789, 1–9. https://doi.org/10.1016/j.
chemosphere.2020.127789
Chen W.-T., Haque Md. A., Lu T., Aierzhati A. and Reimonn G. (2020). A perspective on hydrothermal processing
of sewage sludge. Current Opinion in Environmental Science & Health, 14, 63–73. https://doi.org/10.1016/j.
coesh.2020.02.008
Chow W. L., Chong S., Lim J. W., Chan Y. J., Chong M. F., Tiong T. J., Chin J. K. and Pan G.-T. (2020). Anaerobic
co-digestion of wastewater sludge: a review of potential co-substrates and operating factors for improved
methane yield. Processes, 8(1), 39, 1–21. https://doi.org/10.3390/pr8010039
Cie ś lik B. and Konieczka P. (2017). A review of phosphorus recovery methods at various steps of wastewater
treatment and sewage sludge management. The concept of ‘no solid waste generation’ and analytical
methods. Journal of Cleaner Production, 142, 1728–1740. https://doi.org/10.1016/j.jclepro.2016.11.116
Conti F., Toor S. S., Pedersen T. H., Seehar T. H., Nielsen A. H. and Rosendahl L. A. (2020). Valorization of
animal and human wastes through hydrothermal liquefaction for biocrude production and simultaneous
recovery of nutrients. Energy Conversion and Management, 216, 112925, 1–11. https://doi.org/10.1016/j.
enconman.2020.112925
Dagerskog L. and Olsson O. (2020). Swedish Sludge Management at the Crossroads, Report, Stockholm
Environment Institute, Stockholm, Sweden.
Dimitriadis A. and Bezergianni S. (2017). Hydrothermal liquefaction of various biomass and waste feedstocks for
biocrude production: a state of the art review. Renewable and Sustainable Energy Reviews, 68, 113–125.
https://doi.org/10.1016/j.rser.2016.09.120
Dubey M., Mohapatra S., Tyagi V. K., Suthar S. and Kazmi A. A. (2021). Occurrence, fate, and persistence of
emerging micropollutants in sewage sludge treatment. Environmental Pollution, 273, 116515, 1–18. https://
doi.org/10.1016/j.envpol.2021.116515
Đurđević D., Blecich P. and Jurić Ž. (2019). Energy recovery from sewage sludge: the case study of Croatia.
Energies, 12(10), 1927, 1–19. https://doi.org/10.3390/en12101927
Đurđević D., Trstenjak M. and Hulenić I. (2020). Sewage sludge thermal treatment technology selection by
utilizing the analytical hierarchy process. Water, 12(5), 1–16. https://doi.org/10.3390/w12051255
European Commission (2020). Waste Streams: Sewage Sludge. Available at https://ec.europa.eu/environment/
topics/waste-and-recycling/sewage-sludge_en (accessed 29 September 2021)
Eurostat (2020). Sewage Sludge Production and Disposal. Available at https://ec.europa.eu/eurostat/databrowser/
view/env_ww_spd/default/table?lang=en (accessed 29 September 2021)
Falk J., Skoglund N., Grimm A. and Öhman M. (2020). Fate of phosphorus in fixed bed combustion of biomass and
sewage sludge. Energy & Fuels, 34(4), 4587–4594. https://doi.org/10.1021/acs.energyfuels.9b03976
Flores-Alsina X., Ramin E., Ikumi D., Harding T., Batstone D., Brouckaert C., Sotemann S. and Gernaey K. V.
(2021). Assessment of sludge management strategies in wastewater treatment systems using a plant-wide
approach. Water Research, 190, 116714, 1–15. https://doi.org/10.1016/j.watres.2020.116714
Government of the Netherlands (2016). Green deals overview, Report, Government of the Netherlands,
Netherlands. Available at https://www.greendeals.nl/sites/default/files/uploads/2016/03/Progress_report_
Green_Deals_ENG.pdf (accessed 29 September 2021)
Gretzschel O., Schäfer M., Steinmetz H., Pick E., Kanitz K. and Krieger S. (2020). Advanced wastewater treatment
to eliminate organic micropollutants in wastewater treatment plants in combination with energy-efficient
electrolysis at WWTP Mainz. Energies, 13(14), 3599, 1–27. https://doi.org/10.3390/en13143599
Hudcová H., Vymazal J. and Rozkošný M. (2019). Present restrictions of sewage sludge application in agriculture
within the European Union. Soil and Water Research, 14(2), 104–120. https://doi.org/10.17221/36/2018-SWR
ITRC (2010). Chemical Stabilization: Phosphate and Biosolids Treatment. Mining Waste Treatment Technology
Selection. Available at https://frtr.gov/matrix/documents/Solidification-and-Stabilization/2010-Chemical-
Stabilization-Phosphate-and-Biosolids-Treatment.pdf (accessed 29 September 2021)
Jędrczak A. (2018). Composting and fermentation of biowaste – advantages and disadvantages of processes. Civil
and Environmental Engineering Reports, 28(4), 71–87. https://doi.org/10.2478/ceer-2018-0052
Johansson N. (2021). Circular agreements – exploring the role of agreements and deals as a political tool for a circular
economy. Circular Economy and Sustainability, 1, 499–505. https://doi.org/10.1007/s43615-021-00004-5
Kacprzak M., Neczaj E., Fijałkowski K., Grobelak A., Grosser A., Worwag M., Rorat A., Brattebo H., Almås Å. and
Singh B. R. (2017). Sewage sludge disposal strategies for sustainable development. Environmental Research,
156, 39–46. https://doi.org/10.1016/j.envres.2017.03.010
Krüger O. and Adam C. (2015). Recovery potential of German sewage sludge ash. Waste Management, 45, 400–
406. https://doi.org/10.1016/j.wasman.2015.01.025
Laurich, F. (2011). Hamburg Wasser’s efforts for a self-sufficient energy supply of Kohlbrandhoft/Dradenau
WWTP. International Water Week, Amsterdam.
Law K. P. and Pagilla K. R. (2018). Phosphorus recovery by methods beyond struvite precipitation. Water
Environment Research, 90(9), 840–850. https://doi.org/10.2175/106143017X15131012188006
Law K. P. and Pagilla K. R. (2019). Reclaimed phosphorus commodity reserve from water resource recovery
facilities – a strategic regional concept towards phosphorus recovery. Resources, Conservation and
Recycling, 150, 104429, 1–12. https://doi.org/10.1016/j.resconrec.2019.104429
Lee E., Oliveira D. S. B. L., Oliveira L. S. B. L., Jimenez E., Kim Y., Wang M., Ergas S. J. and Zhang Q. (2020).
Comparative environmental and economic life cycle assessment of high solids anaerobic co-digestion for
biosolids and organic waste management. Water Research, 171, 115443, 1–12. https://doi.org/10.1016/j.
watres.2019.115443
Mills N., Pearce P., Farrow J., Thorpe R. B. and Kirkby N. F. (2014). Environmental and economic life cycle
assessment of current and future sewage sludge to energy technologies. Waste Management, 34(1), 185–195.
https://doi.org/10.1016/j.wasman.2013.08.024
Mininni G., Mauro E., Piccioli B., Colarullo G., Brandolini F. and Giacomelli P. (2019). Production and
characteristics of sewage sludge in Italy. Water Science and Technology, 79(4), 619–626. https://doi.
org/10.2166/wst.2019.064
Mu’azu N. D., Essa M. H., Aga O. and Jarrah N. (2019). Life cycle assessment approach to sustainable sewage
sludge management for water pollution control. Journal of Physics: Conference Series, 1349, 012145, 1–8.
https://doi.org/10.1088/1742-6596/1349/1/012145
MWRD (2021). Biosolids Utilization. Available at https://legacy.mwrd.org/irj/go/km/docs/documents/MWRD/
internet/protecting_the_environment/Water_Reclamation_Plants/htm/Biosolids_Disposal.htm (accessed
29 September 2021)
Nakkasunchi S., Hewitt N. J., Zoppi C. and Brandoni C. (2021). A review of energy optimization modelling tools
for the decarbonisation of wastewater treatment plants. Journal of Cleaner Production, 279, 123811, 1–20.
https://doi.org/10.1016/j.jclepro.2020.123811
OECD.Stat (2021). Wastewater Treatment (% Population Connected). Available at https://stats.oecd.org/Index.
aspx?DataSetCode=WATER_TREAT (accessed 29 September 2021)
Onwosi C. O., Igbokwe V. C., Odimba J. N., Eke I. E., Nwankwoala M. O., Iroh I. N. and Ezeogu L. I. (2017).
Composting technology in waste stabilization: on the methods, challenges and future prospects. Journal of
Environmental Management, 190, 140–157. https://doi.org/10.1016/j.jenvman.2016.12.051
Przydatek G. and Wota A. K. (2020). Analysis of the comprehensive management of sewage sludge in Poland.
Journal of Material Cycles and Waste Management, 22(1), 80–88. https://doi.org/10.1007/s10163-019-00937-y
Righi S., Oliviero L., Pedrini M., Buscaroli A. and Della Casa C. (2013). Life cycle assessment of management
systems for sewage sludge and food waste: centralized and decentralized approaches. Journal of Cleaner
Production, 44, 8–17. https://doi.org/10.1016/j.jclepro.2012.12.004
Rosiek K. (2020). Directions and challenges in the management of municipal sewage sludge in Poland in the
context of the circular economy. Sustainability, 12(9), 3686, 1–28. https://doi.org/10.3390/su12093686
Roskosch A. and Heidecke P. (2018). Sewage Sludge Disposal in the Federal Republic of Germany, Report,
German Environment Agency, Dessau-Roßlau, Germany. Available at https://www.umweltbundesamt.de/
sites/default/files/medien/1410/publikationen/190116_uba_fb_klaerschlamm_engl_bf.pdf (accessed 29
September 2021)
Sánchez Ó. J., Ospina D. A. and Montoya S. (2017). Compost supplementation with nutrients and microorganisms
in composting process. Waste Management, 69, 136–153. https://doi.org/10.1016/j.wasman.2017.08.012
Scarsella M., de Caprariis B., Damizia M. and De Filippis P. (2020). Heterogeneous catalysts for hydrothermal
liquefaction of lignocellulosic biomass: a review. Biomass and Bioenergy, 140, 105662, 1–15. https://doi.
org/10.1016/j.biombioe.2020.105662
Schäfer M., Gretzschel O. and Steinmetz H. (2020). The Possible Roles of Wastewater Treatment Plants in Sector
Coupling. Energies, 13(8), 2088, 1–20. https://doi.org/10.3390/en13082088
Seiple T. E., Skaggs R. L., Fillmore L. and Coleman A. M. (2020). Municipal wastewater sludge as a renewable, cost-
effective feedstock for transportation biofuels using hydrothermal liquefaction. Journal of Environmental
Management, 270, 110852, 1–9. https://doi.org/10.1016/j.jenvman.2020.110852
Shaddel S., Bakhtiary-Davijany H., Kabbe C., Dadgar F. and Østerhus S. (2019). Sustainable sewage sludge
management: from current practices to emerging nutrient recovery technologies. Sustainability, 11(12),
3435, 1–13. https://doi.org/10.3390/su11123435
Shen Y., Linville J. L., Urgun-Demirtas M., Mintz M. M. and Snyder S. W. (2015). An overview of biogas production
and utilization at full-scale wastewater treatment plants (WWTPs) in the United States: challenges and
opportunities towards energy-neutral WWTPs. Renewable and Sustainable Energy Reviews, 50, 346–362.
https://doi.org/10.1016/j.rser.2015.04.129
Silva Thomsen L. B., Carvalho P. N., dos Passos J. S., Anastasakis K., Bester K. and Biller P. (2020). Hydrothermal
liquefaction of sewage sludge; energy considerations and fate of micropollutants during pilot scale processing.
Water Research, 183, 116101, 1–8. https://doi.org/10.1016/j.watres.2020.116101
Singh S., Kumar V., Dhanjal D. S., Datta S., Bhatia D., Dhiman J., Samuel J., Prasad R. and Singh J. (2020). A
sustainable paradigm of sewage sludge biochar: valorization, opportunities, challenges and future prospects.
Journal of Cleaner Production, 269, 122259, 1–16. https://doi.org/10.1016/j.jclepro.2020.122259
Smol M., Adam C. and Anton Kugler S. (2020). Inventory of Polish municipal sewage sludge ash (SSA) – mass
flows, chemical composition, and phosphorus recovery potential. Waste Management, 116, 31–39. https://
doi.org/10.1016/j.wasman.2020.07.042
Stillwell A., Hoppock D. and Webber M. (2010). Energy recovery from wastewater treatment plants in the United
States: a case study of the energy-water nexus. Sustainability, 2(4), 945–962. https://doi.org/10.3390/su2040945
Ś wierczek L., Cie ślik B. M. and Konieczka P. (2021). Challenges and opportunities related to the use of sewage
sludge ash in cement-based building materials – a review. Journal of Cleaner Production, 287, 125054, 1–12.
https://doi.org/10.1016/j.jclepro.2020.125054
Tsybina A. and Wuensch C. (2018). Analysis of sewage sludge thermal treatment in the contect of circular economy.
Detritus, 2(1), 1–15. https://doi.org/10.31025/2611-4135/2018.13668
Turunen V., Sorvari J. and Mikola A. (2018). A decision support tool for selecting the optimal sewage sludge
treatment. Chemosphere, 193, 521–529. https://doi.org/10.1016/j.chemosphere.2017.11.052
US EPA (1992). Technical Support Document for Reduction of Pathogens and Vector Sludge Attraction in
Sewage Sludge, Report EPA 822/R-93–004, Office of Water, US EPA, Washington, DC, USA. Available at
https://www.epa.gov/sites/production/files/2019-12/documents/tsd-reduction-pathogens-vector-1992.pdf
(accessed 29 September 2021)
US EPA (1995). Part 503 Implementation Guidance, Report EPA 833-R-95–001, Office of Water, US EPA, Washington,
DC, USA. Available at https://www3.epa.gov/npdes/pubs/owm0237.pdf (accessed 29 September 2021)
US EPA. (2003a). Biosolids Technology Fact Sheet, Fact Sheet No. 832-F-03–012, pp. 1–8). Office of Water, US
EPA, Washington, DC, USA. Available at https://www.epa.gov/sites/default/files/2018-11/documents/
biosolids-technology-factsheet.pdf (Accessed 23 February 2022).
US EPA. (2003b). Environmental Regulations and Technology: Control of Pathogens and Vector Attraction
in Sewage Sludge (625/R-92/013; pp. 1–186). US EPA Office of Research and Development, US EPA,
Washington, DC, USA. Available at https://www.epa.gov/sites/production/files/2015-07/documents/epa-
625-r-92-013.pdf (Accessed 23 February 2022).
US EPA (2019). Basic Information about Biosolids [Other Policies and Guidance]. Available at https://www.epa.
gov/biosolids/basic-information-about-biosolids (accessed 29 September 2020)
Walker J., Knight L. and Stein L. (1994). A Plain English Guide to the EPA Part 503 Biosolids Rule, Report
EPA/832/R-93/003, Office of Wastewater Management, US EPA, Washington, DC, USA. Available at
https://www.epa.gov/sites/production/files/2018-12/documents/plain-english-guide-part503-biosolids-rule.
pdf (accessed 29 September 2021)
Watson J., Wang T., Si B., Chen W.-T., Aierzhati A. and Zhang Y. (2020). Valorization of hydrothermal liquefaction
aqueous phase: pathways towards commercial viability. Progress in Energy and Combustion Science, 77,
100819, 1–45. https://doi.org/10.1016/j.pecs.2019.100819
Wei L., Zhu F., Li Q., Xue C., Xia X., Yu H., Zhao Q., Jiang J. and Bai S. (2020). Development, current state and
future trends of sludge management in China: based on exploratory data and CO2-equivalent emissions
analysis. Environment International, 144, 106093, 1–12. https://doi.org/10.1016/j.envint.2020.106093
Wiechmann B., Dienemann C., Kabbe C., Simone B., Ines V. and Andrea R. (2013). Sewage Sludge Management
in Germany, Report. Umweltbundesamt, Germany.
Xu Y., Lu Y., Zheng L., Wang Z. and Dai X. (2020). Perspective on enhancing the anaerobic digestion of waste activated
sludge. Journal of Hazardous Materials, 389, 121847, 1–16. https://doi.org/10.1016/j.jhazmat.2019.121847
Ye S., Zeng G., Wu H., Liang J., Zhang C., Dai J., Xiong W., Song B., Wu S. and Yu J. (2019). The effects of activated
biochar addition on remediation efficiency of co-composting with contaminated wetland soil. Resources,
Conservation and Recycling, 140, 278–285. https://doi.org/10.1016/j.resconrec.2018.10.004
Yoshida H., ten Hoeve M., Christensen T. H., Bruun S., Jensen L. S. and Scheutz C. (2018). Life cycle assessment
of sewage sludge management options including long-term impacts after land application. Journal of Cleaner
Production, 174, 538–547. https://doi.org/10.1016/j.jclepro.2017.10.175
Zhang Q., Zhang L., Guo B. and Liu Y. (2020). Mesophiles outperform thermophiles in the anaerobic digestion
of blackwater with kitchen residuals: insights into process limitations. Waste Management, 105, 279–288.
https://doi.org/10.1016/j.wasman.2020.02.018
Chapter 11
Decarbonization potentials in
intensified water and wastewater
systems using membrane-related
technologies
Boyan Xu, Shujuan Huang, Chuansheng Wang, Tze Chiang Albert Ng and How Yong Ng*
National University of Singapore Environmental Research Institute, 5A Engineering Drive 1, 117411, Singapore
*Correspondence: [email protected]
11.1 INTRODUCTION
The foremost role of the wastewater treatment plants (WWTPs) is to remove pollutants in wastewater
to produce a high-quality effluent in order to protect the environment, safeguard human health and/
or achieve water reuse. However, it may be insufficient for some conventional WWTPs to provide
qualified or reusable water for industrial or domestic recycling. For example, conventional activated
sludge (CAS) could poorly remove emerging micro-pollutants in wastewater such as pharmaceutically
active compounds (PhACs) owing to the rapid developing pharmaceutical industries (Radjenović
et al., 2009). To meet the more stringent effluent discharge standards and the increasing need of
wastewater reclamation (Melin et al., 2006), additional treatment processes are called for to polish the
produced wastewater by removing undesirable compounds. Membrane technology, which is at least
half a century old, can selectively remove undesirable components over a wide range of molecular
weights. In addition to the high-quality effluent, the small footprint was another advantage of
membrane technologies, especially considering the growing tension in urban land use. For example,
membrane bioreactors (MBRs) technology (Ng & Ng, 2010), in which membranes were immerged
into activated sludge, could achieve high biochemical efficacy and high quality of treated wastewater.
The increasing demand and consumption of freshwater due to development and further societal
growth leads to an insufficient availability of freshwater for many countries worldwide. Therefore,
desalination processes which extract portable water from non-conventional sources such as seawater
and brackish water become necessary, especially for some water-stressed countries such as Singapore.
Historically, it has been performed through evaporation using thermal energy, and in the late 1970s the
reverse osmosis (RO) membranes started to be implemented for seawater desalination (Figure 11.1a)
(Judd, 2017). The RO technology is not only light-weight, highly compact and productive, but also less
energy intensive than thermal desalination such as multi-stage flash (MSF), multiple-effect distillation
(MED) and thermal vapor compression (TVC), thus leading to the replacement of thermal desalination
technologies in many parts of the world (Ali et al., 2018). In general, it could thus be said that the
© 2022 The Editors. This is an Open Access eBook distributed under the terms of the Creative Commons Attribution Licence
(CC BY-NC-ND 4.0), which permits copying and redistribution for non-commercial purposes with no derivatives, provided the
original work is properly cited (https://creativecommons.org/licenses/by-nc-nd/4.0/). This does not affect the rights licensed or
assigned from any third party in this book. The chapter is from the book Pathways to Water Sector Decarbonization, Carbon Capture
and Utilization, Zhiyong Jason Ren and Krishna Pagilla (Eds.)
Figure 11.1 Configuration and principles of (a) RO, (b) FO, (c) PRO and (d) MD. Black arrows represent the water flow
direction.
membrane technologies have been widely implemented in municipal water sectors (i.e., wastewater
treatment and desalination plants).
According to the United Nations Food and Agriculture Organization (UN-FAO), more than 26 billion
m3 of domestic wastewater was treated in 2009. Moreover, wastewater treatment accounts for about
3% of the United States total electrical consumption annually (USEPA, 2006), and it is at average of
0.3% for China (Hao et al., 2015). The generation of electricity required to drive the WWTPs operation
in China could result in the production of more than 110 million tons of CO2 annually (Hao et al.,
2015). Therefore, because of increasing energy cost, tremendous fossil fuel consumption and climate
change, WWTPs including those using membrane technologies should be designed for improving
energy efficiency and consider resource recovery as a key performance indicator. Considering that the
membrane technologies used in wastewater treatment could be more energy intensive as compared to
conventional wastewater treatment processes (e.g., MBRs versus CAS) (Mannina et al., 2020), it could
increase the carbon footprint and energy consumption in wastewater applications. Moreover, global
desalination of seawater and brackish water via RO could contribute more than 50% of drinking water
production (Judd, 2017). Although RO is the most economical technology for seawater desalination at
the commercial scale as compared to the thermal technologies, and has achieved a decline in production
cost from $4.5 m−3 in 1997 to ∼$1.5 m−3 in 2000 for seawater reverse osmosis (SWRO) with the help
of the pressure exchanger (PX) as energy recovery devices (ERDs) (Judd, 2017), it still requires a
specific energy ranging between 3 and 4 kWh/m3 which is more than double of the theoretical energy
requirement (i.e., 1.06 kWh/m3 for seawater with the salt concentration of 35 000 ppm and at a 50%
recovery) (Ali et al., 2018). Thus, without doubt, the current high energy demands from wastewater
treatment and desalination need to be reduced to achieve both economically and environmentally
sustainable water/wastewater treatment processes.
Efforts by the water industry and scientific community in research led to the development of some
novel membrane processes, and an integration of membrane technologies to conventional systems
could have much potential for energy saving, resource recovery and decarbonization. Therefore, it is
of obvious interest, and it is necessary to summarize these latest developments in achieving energy
self-sufficiency, and even carbon neutrality, by using novel membrane technologies for wastewater
treatment/reclamation and desalination systems. Especially, novel membrane technologies that
could reduce energy consumption with high decarbonization potential are discussed for wastewater
treatment, including aerobic granular sludge membrane bioreactors (AGMBRs), algae membrane
bioreactors (A-MBRs), anaerobic membrane bioreactors (AnMBRs), membrane biofilm reactors
(MBfRs) and forward osmosis (FO) integrated processes. Moreover, this chapter also includes the
membrane technologies for desalination with decarbonization potential, consisting of pressure
retarded osmosis (PRO), forward osmosis-reverse osmosis (FO-RO) hybrid and forward osmosis-
membrane distillation hybrid (FO-MD).
which limits the wide application of AGMBR operated under the continuous mode (Chen et al., 2017;
Corsino et al., 2016; Li et al., 2014). Recently, a few researchers have attempted to overcome this
challenge by recreating or simulating similar hydraulic conditions as those in SBRs in a continuous
configuration (Chen et al., 2017; Corsino et al., 2016). An internal circulation AGMBR, which was
achieved by driving liquid downward in the anoxic zone and upward in the aerobic MBR zone, was
designed to mimic the hydraulic condition in SBR to cultivate AGS in a continuous flow bioreactor
by Chen et al. (2017). The granules were successfully cultivated after 35 days with an average particle
size of 0.228 mm; however, the membrane filtration behavior was not considered in this study.
Corsino et al. (2016) attempted to recreate the hydraulic conditions in SBR to cultivate granules in
a novel hydrodynamic configuration of AGMBR. The results showed that the membrane filtration
performance of the AGMBR was improved 90% as compared to a conventional MBR. However, the
granules disintegrated in less than 20 days under the continuous operation mode. The information
and clear mechanisms regarding the granules cultivation in continuous membrane reactors is still
limited due to the complex mechanism of granulation, and therefore the cultivation and stability of
granules is the key challenge for the scaling-up of the low-carbon-footprint AGMBRs.
affected by various factors such as the ratio of algae to biomass, influent wastewater characteristics
and operation conditions (Sun et al., 2018b).
Figure 11.2 Schematic process of potential energy recovery from wastewater by using AnMBR (adopted from Shin
& Bae, 2018) and a configuration of an AnMBR.
Yet, AnMBRs for treatment of domestic wastewater may not generate a high methane yield due to the
low organic loadings and high portion of dissolution of methane in the effluents (Crone et al., 2016).
For high-strength wastewater treatment, AnMBR presented a high possibility to achieve energy self-
sufficiency by converting waste to energy. For example, the methane content of biogas produced by an
AnMBR treating landfill leachate ranged from 70 to 90% with a methane yield of 0.34 L/g CODremoved
(Xie et al., 2014). Moreover, as for food waste digested by AnMBRs, net energy benefit could be
potentially obtained, which could be due to the methane-rich biogas production of 0.21 ± 0.1 L CH4/g
CODremoved (Galib et al., 2016; Jeong et al., 2017).
In general, AnMBRs have the potential to be an energy-positive technology for wastewater treatment/
reclamation and water reclamation. However, serious issues such as serious membrane fouling (Robles
et al., 2012), lack of nutrients removal and high dissolved methane in the effluent (Crone et al., 2016)
have to be addressed in energy-efficient and cost-effective ways in order for the worldwide full-scale
applications of AnMBRs. Instead of energy-intensive biogas sparging, more energy efficient AnMBRs
may adopt proper strategies for fouling control, including physical (e.g., rotating membranes), chemical
(e.g., NaClO) and biological methods (e.g., quorum quenching) (Shin & Bae, 2018; Xu et al., 2020b;
Yue et al., 2018). Dissolved methane could be removed by using a hollow fiber membrane contactor
which could achieve an average removal efficiency of more than 70% (Lim et al., 2019). Novel hybrid
technology like forward osmosis (FO)-AnMBR could help reduce dissolved methane in the effluent
(Chen et al., 2014). Noticeably, coupling anaerobic ammonium oxidation (anammox) with nitrite/
nitrate-dependent anaerobic methane oxidation (n-DAMO) simultaneously removed up to 85%
dissolved methane and more than 99% nitrogen from synthetic anaerobic effluent (Liu et al., 2020),
which would thus reduce greenhouse gas (i.e., CH4) emission as well as the external carbon addition
for the post-treatment of AnMBR’s effluent.
11.2.5 Forward osmosis (FO) integrated processes for wastewater treatment/reclamation and
resource recovery
11.2.5.1 Microfiltration forward osmosis membrane bioreactors (MF-FOMBRs)
Compared with hydraulic pressure-driven membrane technologies such as RO, the osmotic pressure-
driven forward osmosis (FO) is featured with low fouling propensity and thus low energy input (Parida
& Ng, 2013). In the FO process, the osmotic energy of the concentrated solution from concentrated
solution side could draw water molecules from the dilute solution side across the FO membrane,
while salts could be rejected (Figure 11.1b) (Parida & Ng, 2013). FO membranes could be used for
resource recovery. For example, Bao et al. (2020) stated that the FO membrane modified with a
moderate primary amine could achieve a high anti-fouling capability and recover ammonium with
rejection above 94% for concentrating domestic wastewater. Phosphorus is also a non-renewable
resource. However, due to the low concentration of phosphorus in domestic wastewater, there are
limited technologies for direct phosphorus recovery from domestic wastewater. Microfiltration
forward osmosis membrane bioreactor (MF-FOMBR), where a forward osmosis membrane bioreactor
(FOMBRs) and a microfiltration (MF) membrane are operated in parallel, may achieve wastewater
treatment/reclamation as well as more than 90% phosphorus recovery with a phosphorus content of
11.1–13.3% in recovered amorphous calcium phosphate precipitates (Figure 11.4a) (Qiu et al., 2015).
The FO membrane could reject the nutrients, while the MF membrane allows them to pass through, and
the phosphorus is then recovered from the nutrient-rich MF permeate without addition of Fe3+, Ca 2+
and Mg 2+. As no biological activity is required in MF-FOMBR, obviating the need for the enrichment
Figure 11.4 (a) System configuration of microfiltration forward osmosis membrane bioreactors (MF-FOMBRs) for
wastewater reclamation as well as nutrient recovery and (b) forward osmotic membrane bioreactors (FOMBRs)-MD
hybrid system (FOMBR-MD) for wastewater reclamation.
of phosphate accumulating organisms (PAOs) and the downstream disposal of phosphorus-rich sludge
could lead to energy saving. Moreover, the FO membrane in the FOMBRs offers the advantage of low
membrane fouling potential which would require lower hydraulic pressure, scouring intensity and
frequency of backwashing (Achilli et al., 2009; Qiu et al., 2015), although FO membrane fouling could
result from a reduction in the osmotic pressure driving force due to the increasing bioreactor salinity
(Holloway et al., 2015).
The energy consumption of sustainable MF-FOMBR should be competitive to the CAS systems,
followed by the advanced treatment processes for the high-quality water reuse (Holloway et al.,
2015). Currently, the intensive energy input for recovery of draw solution (DS) significantly limits the
scaling-up of FOMBR and other FO related technology including MF-FOMBR. Regeneration of DS
for FOMBRs may be achieved by the RO process (Holloway et al., 2015). According to the reverse
osmosis system analysis system (ROSA) design software (Dow Filmtec, Edina, MN), the optimized
approximate specific energy demand calculated to reconcentrate a draw solution of 40 g L −1-NaCl
is 1.6 kWh m−3 (Holloway et al., 2015). It can drop to approximately 1.1 kWh m−3 if a pressure
exchanger is to be incorporated into the RO system. However, this value would still exceed that of
current advanced wastewater treatment processes (Holloway et al., 2015). Thus, to overcome the
major barrier for the implementation of full-scale MF-FOMBR, developing creative and efficient draw
solution regeneration configurations is necessary to reduce the energy consumption associated with
FOMBRs. Another pressure-driven filtration, namely nanofiltration (NF), which could achieve a high
rejection of multivalent ions as well as a sufficiently low pressure for high water recovery rate, is a
promising option for DS regeneration. High quality effluent was generated from a FOMBR and NF
hybrid system for agricultural irrigation, and the DS replacement costs were reduced (Corzo et al.,
2018). Although seawater brine from SWRO process is considered a waste product, it can be used
as an existing and easily accessible high osmotic pressure draw solution that will help increase the
economic feasibility of existing FOMBR, as the DS reconcentration is not necessary (Qiu et al., 2015).
Moreover, it could create a more sustainable way for disposal of seawater brine, as the diluted brine
from FOMBR systems could have a reduced influence on marine ecosystems. Thus, from the energy
perspective, using seawater brine as DS for FOMBRs could be expected to make positive impacts on
the sustainability of wastewater treatment and carbon footprint, though the diluted brine cannot be
used as reusable water without further advanced treatment and the long-term operation of the system
with regard to the accumulation of pollutants and membrane fouling needs to be further investigated
(Qiu et al., 2015).
due to the identical water transfer rate (WTR) between the FO and MD process, resulting in much
lower energy consumption. Moreover, to achieve a concentration factor (CF) of 10 when treating
500 mL of textile wastewater using the symmetric FO membrane in the hybrid process, it only used
the lowest total cost of 0.17 USD among the three. However, possible improvements with regards to
the feasibility of various hard-to-remove compounds and other optimized draw solutes need to be
further investigated to examine the promising potential of symmetric FO membranes for various types
of wastewater treatment (Li et al., 2020).
Figure 11.5 (a) System configuration of river-to-sea PRO system for power generation, (b) SWRO-PRO system for
desalination and power generation and (c) FO and RO hybrid system for desalination and wastewater concentration.
Figure 11.6 Sankey diagram of a river-to-sea PRO facility adopted from a previous study (O’Toole et al., 2016). Total
energy calculated based on the Gibbs free energy of mixing at 18°C. The unit is kWh/m3.
Table 11.1 Hybrid membrane technologies for desalination with low energy consumption.
foulants, as it has been already pretreated by the RO pretreatment system such as ultrafiltration (UF).
It would thus eliminate the parasitic energy consumption, as compared to that of the river-to-sea PRO
system (Achilli et al., 2014). Moreover, the membrane power densities of the SWRO-PRO could be
higher than 5 W/m 2 which largely exceeds the reported 1.5 W/m 2 for the river-to-sea PRO pilot system
in Korea (Achilli et al., 2014; Kim et al., 2013). For the energy perspective, a SWRO-PRO pilot system
at 50% RO recovery could yield approximately 1.1 kWh/m3, which could thus reduce the specific
energy for SWRO to around 1 kWh/m3 (Prante et al., 2014).
However, wastewater retentate, which has a salinity close to that of river water, was preferred as the
feed water for the PRO system rather than river water due to the scarcity of freshwater in some regions
such as Singapore (Wan & Chung, 2015). Therefore, membrane fouling due to the foulants from
wastewater retentate could be still a major problem for the SWRO-PRO systems for power generation.
The reduction in the power densities could largely result from the fouling on the porous substrate of
the PRO membranes. Therefore, both UF and NF has been employed as pretreatment processes to
boost the power densities to 6.6 and 8.9 W/m 2, respectively, as they could mitigate fouling potential
of the wastewater retentate (Wan & Chung, 2015). Moreover, coagulation of feed wastewater could
also be helpful (Wan et al., 2019). As compared to the untreated wastewater which caused a 69.3%
flux reduction, AlCl3 and NaAlO2 have been demonstrated to increase the normalized water flux to 66
and 64%, respectively. The initial water fluxes have also been increased to 25.5 and 24.8 LMH at 20
bars, respectively (Wan et al., 2019). Rather than pretreatment, increasing the anti-fouling membrane
properties could also be demonstrated to improve the SWRO-PRO, and a wide range of nanomaterials
could be used to modify the membrane such as graphene-based materials, carbon nanotubes and
zeolites, which endowed the PRO membranes with favorable membrane structures and enhanced
desirable antifouling characteristics (Wan et al., 2020). In general, to further scale up more PRO into
RO systems for low-carbon-footprint desalination, PRO membrane fouling needs to be well-mitigated.
11.3.2 Forward osmosis-reverse osmosis (FO-RO) hybrid for desalination and wastewater
concentration
FO and RO hybrid system (FO-RO) is considered as another green technology because it can be
employed to desalt seawater and concentrate wastewater simultaneously (Figure 11.5c) (Linares et al.,
2016). Shaffer et al. (2015) announced that the ‘FO process is not intended to replace RO, but rather
is to be used to process feed waters that cannot be treated by RO’. The FO system can use seawater
on one side of the FO membrane and wastewater on the other side, and it could lead to the reduction
of osmotic pressure of seawater prior to RO desalination. Reducing the volume of wastewater could
reduce the energy consumption for transportation as well as treatment processes (Linares et al., 2016).
Moreover, it could be more efficient to harvest energy (e.g., biogas) and nutrients (e.g., phosphates)
from the resulting concentrated wastewater. The concentrated wastewater can potentially be treated
by anaerobic treatment with enhanced biogas (CH4) production (Amy et al., 2017). For example, the
methane yield of an AnMBR could progressively increase from 214 to 322 mL- CH4/g- COD when the
pre-concentration factor of domestic wastewater by FO increased from 1 to 10 (Vinardell et al., 2021).
The energy consumption of the overall FO-RO system can be offset by the enhanced biogas produced
(Amy et al., 2017).
Furthermore, potential electricity savings could be also achieved in the SWRO facilities by lowering
the operating hydraulic pressure (Linares et al., 2016). The brackish water RO membranes (BWRO)
could also be used instead of SWRO membranes. Moreover, higher flux could be employed for the
diluted seawater, which could thus increase the water recovery of the whole system. Furthermore,
discharging brines with lower salinity would have less adverse impacts on the aquatic ecosystem.
The specific energy consumption (SEC) associated to the FO-RO process for desalination ranged
between 1.3 and 1.5 kWh m−3, and it was calculated with the total capacity of 2400 m3 d−1 based
on a conservative estimate using a secondary wastewater effluent as the feed and seawater as the
draw solution (Yangali-Quintanilla et al., 2011). This range of SEC could be lower than that of the
conventional RO process for seawater desalination.
However, the critical aspect in the CAPEX of FO-RO hybrid systems results from the FO
membranes. The use of an FO-RO system could be more viable if commercial FO membrane modules
can be produced with a reasonable price (Table 11.1) (Linares et al., 2016). Overall, the FO-RO system
could have a 56% lower OPEX due to the energy saving for diluted seawater desalination as compared
to the conventional SWRO, although there is a 21% higher CAPEX due to the implementation of
the FO systems (Linares et al., 2016). To calculate the total cost per cubic meter of desalted water,
the FO-RO hybrid desalination system could achieve a cost reduction of 16% as compared to the
SWRO. Moreover, economic evaluation of the FO-RO hybrid process was evaluated comparing to a
conventional two-stage SWRO. Spiral wound FO and plate-and-frame FO-RO hybrid processes can
achieve cost reductions of $355.3 million and $310.2 million, respectively, over a period of 20 years
(Im et al., 2020).
direct contact MD (Francis et al., 2014), air gap MD (Alsaadi et al., 2015) and vacuum MD (Alsaadi
et al., 2014). Moreover, efficient internal heat recovery and satisfactory flux are also considered as
main obstacles for MD module scale-up (Amy et al., 2017). Overall, more research needs to be further
conducted to improve the performance as well as the sustainability of FO-MD systems for low-carbon-
footprint desalination.
11.4 CONCLUSIONS
In addition to meeting the tightened effluent discharge standards in many countries, membrane
technologies are commonly used to address other objectives including protection of public health
and ecological issues and producing reusable water. There is an increasing trend to consider energy
consumption associated with various membrane technologies for water and wastewater treatments.
To achieve truly sustainable treatment processes, the intensive energy consumption of membrane-
based wastewater/water treatment plants is a key challenge and must be dealt with accordingly. Novel
and modified membrane-based strategies are developed to reduce energy and carbon footprint and
recover resources from water within the circular economy.
The way towards energy self-sufficient operation of the above summarized membrane-based
processes for wastewater treatment/reclamation is aiming to directly capture energy and nutrients from
wastewater (e.g., P, N and biogas) and minimize energy consumption such as aeration. As membrane
fouling is still a main obstacle for membrane-based technologies, especially the AnMBRs and the direct
membrane filtration, further optimization should be conducted to address it in a more cost-effective
and holistic manner. Membrane-based desalination operations mainly rely on RO as the baseline
conventional technology. However, it is challenged by the significant specific energy consumption as
well as the adverse environmental impacts such as greenhouse gas emissions. Interesting membrane
processes including FO, PRO and MD were proposed, and the hybrid technologies encompassing
a mix of new and conventional technologies can generate clean water and sustainable electricity
simultaneously. However, parasitic drawbacks are with these novel membrane-based strategies for
desalination such as the serious fouling potential of PRO, the high membrane cost of FO and the
low membrane flux of MD. We summarized the pains and gains of each hybrid technologies in this
chapter and suggested that more research optimization and development is necessary for their next
steps towards practical and worldwide implementation for energy saving and low carbon footprint.
REFERENCES
Achilli A., Cath T. Y., Marchand E. A. and Childress A. E. (2009). The forward osmosis membrane bioreactor:
a low fouling alternative to MBR processes. DesalinationI, 239(1–3), 10–21, https://doi.org/10.1016/j.
desal.2008.02.022
Achilli A., Prante J. L., Hancock N. T., Maxwell E. B. and Childress A. E. (2014). Experimental results from
RO-PRO: a next generation system for low-energy desalination. Environmental Science and Technology,
48(11), 6437–6443, https://doi.org/10.1021/es405556s
Ali A., Tufa R. A., Macedonio F., Curcio E. and Drioli E. (2018). Membrane technology in renewable-energy-
driven desalination. Renewable and Sustainable Energy Reviews, 81, 1–21, https://doi.org/10.1016/j.
rser.2017.07.047
Alsaadi A. S., Francis L., Amy G. L. and Ghaffour N. (2014). Experimental and theoretical analyses of temperature
polarization effect in vacuum membrane distillation. Journal of Membrane Science, 471, 138–148, https://
doi.org/10.1016/j.memsci.2014.08.005
Alsaadi A. S., Francis L., Maab H., Amy G. L. and Ghaffour N. (2015). Evaluation of air gap membrane distillation
process running under sub-atmospheric conditions: experimental and simulation studies. Journal of
Membrane Science, 489, 73–80, https://doi.org/10.1016/j.memsci.2015.04.008
Amy G., Ghaffour N., Li Z., Francis L., Linares R. V., Missimer T. and Lattemann S. (2017). Membrane-based
seawater desalination: present and future prospects. Desalination, 401, 16–21, https://doi.org/10.1016/j.
desal.2016.10.002
Aybar M., Pizarro G., Boltz J., Downing L. and Nerenberg R. (2014). Energy-efficient wastewater treatment via the
air-based, hybrid membrane biofilm reactor (hybrid MfBR). Water Science and Technology, 69(8), 1735–1741,
https://doi.org/10.2166/wst.2014.086
Bao X., She Q., Long W. and Wu Q. (2020). Ammonium ultra-selective membranes for wastewater treatment and
nutrient enrichment: interplay of surface charge and hydrophilicity on fouling propensity and ammonium
rejection. Water Research, 190, 116678, https://doi.org/10.1016/j.watres.2020.116678
Bilad M., Arafat H. A. and Vankelecom I. F. (2014). Membrane technology in microalgae cultivation and
harvesting: a review. Biotechnology Advances, 32(7), 1283–1300, https://doi.org/10.1016/j.biotechadv.
2014.07.008
Chen L., Gu Y., Cao C., Zhang J., Ng J.-W. and Tang C. (2014). Performance of a submerged anaerobic membrane
bioreactor with forward osmosis membrane for low-strength wastewater treatment. Water Research, 50,
114–123, https://doi.org/10.1016/j.watres.2013.12.009
Chen C., Bin L., Tang B., Huang S., Fu F., Chen Q., Wu L. and Wu C. (2017). Cultivating granular sludge directly
in a continuous-flow membrane bioreactor with internal circulation. Chemical Engineering Journal, 309,
108–117, https://doi.org/10.1016/j.cej.2016.10.034
Cheng W., Ma J., Zhang X. and Elimelech M. (2019). Sub-1 µm free-standing symmetric membrane for osmotic
separations. Environmental Science & Technology Letters, 6(8), 492–498, https://doi.org/10.1021/acs.
estlett.9b00364
Chernicharo C., Van Lier J., Noyola A. and Ribeiro T. B. (2015). Anaerobic sewage treatment: state of the art,
constraints and challenges. Reviews in Environmental Science and Bio/Technology, 14(4), 649–679, https://
doi.org/10.1007/s11157-015-9377-3
Corsino S., Campo R., Di Bella G., Torregrossa M. and Viviani G. (2016). Study of aerobic granular sludge
stability in a continuous-flow membrane bioreactor. Bioresource Technology, 200, 1055–1059, https://doi.
org/10.1016/j.biortech.2015.10.065
Corzo B., de la Torre T., Sans C., Escorihuela R., Navea S. and Malfeito J. J. (2018). Long-term evaluation of
a forward osmosis-nanofiltration demonstration plant for wastewater reuse in agriculture. Chemical
Engineering Journal, 338, 383–391, https://doi.org/10.1016/j.cej.2018.01.042
Crone B. C., Garland J. L., Sorial G. A. and Vane L. M. (2016). Significance of dissolved methane in effluents of
anaerobically treated low strength wastewater and potential for recovery as an energy product: a review.
Water Research, 104, 520–531, https://doi.org/10.1016/j.watres.2016.08.019
Drexler I. L. C. and Yeh D. H. (2014). Membrane applications for microalgae cultivation and harvesting: a
review. Reviews in Environmental Science and Bio/Technology, 13(4), 487–504, https://doi.org/10.1007/
s11157-014-9350-6
Francis L., Ghaffour N., Alsaadi A. S., Nunes S. P. and Amy G. L. (2014). Performance evaluation of the DCMD
desalination process under bench scale and large scale module operating conditions. Journal of Membrane
Science, 455, 103–112, https://doi.org/10.1016/j.memsci.2013.12.033
Galib M., Elbeshbishy E., Reid R., Hussain A. and Lee H.-S. (2016). Energy-positive food wastewater treatment
using an anaerobic membrane bioreactor (AnMBR). Journal of Environmental Management, 182, 477–485,
https://doi.org/10.1016/j.jenvman.2016.07.098
Ge Q., Wang P., Wan C. and Chung T.-S. (2012). Polyelectrolyte-promoted forward osmosis–membrane distillation
(FO–MD) hybrid process for dye wastewater treatment. Environmental Science and Technology, 46(11),
6236–6243, https://doi.org/10.1021/es300784h
Ge Q., Han G. and Chung T.-S. (2016). Effective As (III) removal by a multi-charged hydroacid complex draw
solute facilitated forward osmosis-membrane distillation (FO-MD) processes. Environmental Science and
Technology, 50(5), 2363–2370, https://doi.org/10.1021/acs.est.5b05402
González D., Amigo J. and Suárez F. (2017). Membrane distillation: perspectives for sustainable and improved
desalination. Renewable and Sustainable Energy Reviews, 80, 238–259, https://doi.org/10.1016/j.
rser.2017.05.078
Hao X., Liu R. and Huang X. (2015). Evaluation of the potential for operating carbon neutral WWTPs in China.
Water Research, 87, 424–431, https://doi.org/10.1016/j.watres.2015.05.050
Hasar H. (2009). Simultaneous removal of organic matter and nitrogen compounds by combining a membrane
bioreactor and a membrane biofilm reactor. Bioresource Technology, 100(10), 2699–2705, https://doi.
org/10.1016/j.biortech.2008.12.024
Holloway R. W., Achilli A. and Cath T. Y. (2015). The osmotic membrane bioreactor: a critical review. Environmental
Science: Water Research and Technology, 1(5), 581–605, https://doi.org/10.1039/C5EW00103J
Im S. J., Jeong S., Jeong S. and Jang A. (2020). Techno-economic evaluation of an element-scale forward
osmosis-reverse osmosis hybrid process for seawater desalination. Desalination, 476, 114240, https://doi.
org/10.1016/j.desal.2019.114240
Iorhemen O. T., Hamza R. A., Zaghloul M. S. and Tay J. H. (2019). Aerobic granular sludge membrane bioreactor
(AGMBR): extracellular polymeric substances (EPS) analysis. Water Research, 156, 305–314, https://doi.
org/10.1016/j.watres.2019.03.020
Jeong Y., Hermanowicz S. W. and Park C. (2017). Treatment of food waste recycling wastewater using anaerobic
ceramic membrane bioreactor for biogas production in mainstream treatment process of domestic wastewater.
Water Research, 123, 86–95, https://doi.org/10.1016/j.watres.2017.06.049
Judd S. J. (2017). Membrane technology costs and me. Water Research, 122, 1–9, https://doi.org/10.1016/j.
watres.2017.05.027
Kim Y. C., Kim Y., Oh D. and Lee K. H. (2013). Experimental investigation of a spiral-wound pressure-retarded
osmosis membrane module for osmotic power generation. Environmental Science and Technology, 47(6),
2966–2973, https://doi.org/10.1021/es304060d
Lei Z., Yang S., Li Y.-Y., Wen W., Wang X. C. and Chen R. (2018). Application of anaerobic membrane bioreactors
to municipal wastewater treatment at ambient temperature: a review of achievements, challenges, and
perspectives. Bioresource Technology, 267, 756–768, https://doi.org/10.1016/j.biortech.2018.07.050
Li J., Cai A., Wang M., Ding L. and Ni Y. (2014). Aerobic granulation in a modified oxidation ditch with an adjustable
volume intraclarifier. Bioresource Technology, 157, 351–354, https://doi.org/10.1016/j.biortech.2014.01.130
Li J., Liu Y., Li X. and Cheng F. (2019). Reactor performance and membrane fouling of a novel submerged aerobic
granular sludge membrane bioreactor during long-term operation. Journal of Water Reuse and Desalination,
9(1), 1–9, https://doi.org/10.2166/wrd.2017.019
Li M., Li K., Wang L. and Zhang X. (2020). Feasibility of concentrating textile wastewater using a hybrid forward
osmosis-membrane distillation (FO-MD) process: performance and economic evaluation. Water Research,
172, 115488, https://doi.org/10.1016/j.watres.2020.115488
Liang Z., Liu Y., Ge F., Xu Y., Tao N., Peng F. and Wong M. (2013). Efficiency assessment and pH effect in
removing nitrogen and phosphorus by algae-bacteria combined system of Chlorella vulgaris and Bacillus
licheniformis. Chemosphere, 92(10), 1383–1389, https://doi.org/10.1016/j.chemosphere.2013.05.014
Liébana R., Modin O., Persson F. and Wilén B.-M. (2018). Integration of aerobic granular sludge and membrane
bioreactors for wastewater treatment. Critical Reviews in Biotechnology, 38(6), 801–816, https://doi.org/10
.1080/07388551.2017.1414140
Lim K., Evans P. J. and Parameswaran P. (2019). Long-term performance of a pilot-scale gas-sparged anaerobic
membrane bioreactor under ambient temperatures for holistic wastewater treatment. Environmental Science
and Technology, 53(13), 7347–7354, https://doi.org/10.1021/acs.est.8b06198
Linares R. V., Li Z., Yangali-Quintanilla V., Ghaffour N., Amy G., Leiknes T. and Vrouwenvelder J. S. (2016). Life
cycle cost of a hybrid forward osmosis–low pressure reverse osmosis system for seawater desalination and
wastewater recovery. Water Research, 88, 225–234, https://doi.org/10.1016/j.watres.2015.10.017
Liu T., Li J., Khai Lim Z., Chen H., Hu S., Yuan Z. and Guo J. (2020). Simultaneous removal of dissolved methane
and nitrogen from synthetic mainstream anaerobic effluent. Environmental Science and Technology, 54(12),
7629–7638, https://doi.org/10.1021/acs.est.0c00912
Low S. L., Ong S. L. and Ng H. Y. (2016). Characterization of membrane fouling in submerged ceramic membrane
photobioreactors fed with effluent from membrane bioreactors. Chemical Engineering Journal, 290, 91–102,
https://doi.org/10.1016/j.cej.2016.01.005
Luo W., Phan H. V., Li G., Hai F. I., Price W. E., Elimelech M. and Nghiem L. D. (2017a). An osmotic membrane
bioreactor–membrane distillation system for simultaneous wastewater reuse and seawater desalination:
performance and implications. Environmental Science and Technology, 51(24), 14311–14320, https://doi.
org/10.1021/acs.est.7b02567
Luo Y., Le-Clech P. and Henderson R. K. (2017b). Simultaneous microalgae cultivation and wastewater treatment
in submerged membrane photobioreactors: a review. Algal Research, 24, 425–437, https://doi.org/10.1016/j.
algal.2016.10.026
Mannina G., Cosenza A. and Rebouças T. F. (2020). A plant-wide modelling comparison between membrane
bioreactors and conventional activated sludge. Bioresource Technology, 297, 122401, https://doi.
org/10.1016/j.biortech.2019.122401
Martin K. J. and Nerenberg R. (2012). The membrane biofilm reactor (MBfR) for water and wastewater treatment:
principles, applications, and recent developments. Bioresource Technology, 122, 83–94, https://doi.
org/10.1016/j.biortech.2012.02.110
Melin T., Jefferson B., Bixio D., Thoeye C., De Wilde W., De Koning J., Van der Graaf J. and Wintgens T. (2006).
Membrane bioreactor technology for wastewater treatment and reuse. Desalination, 187(1–3), 271–282,
https://doi.org/10.1016/j.desal.2005.04.086
Morrow C. P., Furtaw N. M., Murphy J. R., Achilli A., Marchand E. A., Hiibel S. R. and Childress A. E. (2018).
Integrating an aerobic/anoxic osmotic membrane bioreactor with membrane distillation for potable reuse.
Desalination, 432, 46–54, https://doi.org/10.1016/j.desal.2017.12.047
Ng T. C. A. and Ng H. Y. (2010). Characterisation of initial fouling in aerobic submerged membrane bioreactors
in relation to physico-chemical characteristics under different flux conditions. Water Research, 44(7), 2336–
2348, https://doi.org/10.1016/j.watres.2009.12.038
Nguyen N. C., Nguyen H. T., Ho S.-T., Chen S.-S., Ngo H. H., Guo W., Ray S. S. and Hsu H.-T. (2016). Exploring
high charge of phosphate as new draw solute in a forward osmosis–membrane distillation hybrid system for
concentrating high-nutrient sludge. Science of the Total Environment, 557, 44–50, https://doi.org/10.1016/j.
scitotenv.2016.03.025
Nhat P. V. H., Ngo H., Guo W., Chang S., Nguyen D. D., Nguyen P., Bui X.-T., Zhang X. and Guo J. (2018). Can
algae-based technologies be an affordable Green process for biofuel production and wastewater remediation?
Bioresource Technology, 256, 491–501, https://doi.org/10.1016/j.biortech.2018.02.031
O’Toole G., Jones L., Coutinho C., Hayes C., Napoles M. and Achilli A. (2016). River-to-sea pressure retarded
osmosis: resource utilization in a full-scale facility. Desalination, 389, 39–51, https://doi.org/10.1016/j.
desal.2016.01.012
Parida V. and Ng H. Y. (2013). Forward osmosis organic fouling: effects of organic loading, calcium and membrane
orientation. Desalination, 312, 88–98, https://doi.org/10.1016/j.desal.2012.04.029
Prante J. L., Ruskowitz J. A., Childress A. E. and Achilli A. (2014). RO-PRO desalination: an integrated low-energy
approach to seawater desalination. Applied Energy, 120, 104–114, https://doi.org/10.1016/j.apenergy.2014.01.013
Qiu G., Law Y.-M., Das S. and Ting Y.-P. (2015). Direct and complete phosphorus recovery from municipal wastewater
using a hybrid microfiltration-forward osmosis membrane bioreactor process with seawater brine as draw
solution. Environmental Science and Technology, 49(10), 6156–6163, https://doi.org/10.1021/es504554f
Radjenović J., Petrović M. and Barceló D. (2009). Fate and distribution of pharmaceuticals in wastewater and
sewage sludge of the conventional activated sludge (CAS) and advanced membrane bioreactor (MBR)
treatment. Water Research, 43(3), 831–841, https://doi.org/10.1016/j.watres.2008.11.043
Robles A., Durán F., Ruano M., Ribes J. and Ferrer J. (2012). Influence of total solids concentration on membrane
permeability in a submerged hollow-fibre anaerobic membrane bioreactor. Water Science and Technology,
66(2), 377–384, https://doi.org/10.2166/wst.2012.196
Robles Á., Durán F., Giménez J. B., Jiménez E., Ribes J., Serralta J., Seco A., Ferrer J. and Rogalla F. (2020).
Anaerobic membrane bioreactors (AnMBR) treating urban wastewater in mild climates. Bioresource
Technology, 314, 123763, https://doi.org/10.1016/j.biortech.2020.123763
Shaffer D. L., Werber J. R., Jaramillo H., Lin S. and Elimelech M. (2015). Forward osmosis: where are we now?
Desalination, 356, 271–284, https://doi.org/10.1016/j.desal.2014.10.031
Shin C. and Bae J. (2018). Current status of the pilot-scale anaerobic membrane bioreactor treatments of
domestic wastewaters: a critical review. Bioresource Technology, 247, 1038–1046, https://doi.org/10.1016/j.
biortech.2017.09.002
Smith A. L., Stadler L. B., Love N. G., Skerlos S. J. and Raskin L. (2012). Perspectives on anaerobic membrane
bioreactor treatment of domestic wastewater: a critical review. Bioresource Technology, 122, 149–159,
https://doi.org/10.1016/j.biortech.2012.04.055
Smith A. L., Stadler L. B., Cao L., Love N. G., Raskin L. and Skerlos S. J. (2014). Navigating wastewater energy
recovery strategies: a life cycle comparison of anaerobic membrane bioreactor and conventional treatment
systems with anaerobic digestion. Environmental Science and Technology, 48(10), 5972–5981, https://doi.
org/10.1021/es5006169
Sun L., Tian Y., Zhang J., Cui H., Zuo W. and Li J. (2018a). A novel symbiotic system combining algae and sludge
membrane bioreactor technology for wastewater treatment and membrane fouling mitigation: performance
and mechanism. Chemical Engineering Journal, 344, 246–253, https://doi.org/10.1016/j.cej.2018.03.090
Sun L., Tian Y., Zhang J., Li H., Tang C. and Li J. (2018b). Wastewater treatment and membrane fouling with
algal-activated sludge culture in a novel membrane bioreactor: influence of inoculation ratios. Chemical
Engineering Journal, 343, 455–459, https://doi.org/10.1016/j.cej.2018.03.022
Sun S.-P. and Chung T.-S. (2013). Outer-selective pressure-retarded osmosis hollow fiber membranes from vacuum-
assisted interfacial polymerization for osmotic power generation. Environmental Science and Technology,
47(22), 13167–13174, https://doi.org/10.1021/es403270n
Tang T. and Hu Z. (2016). A comparison of algal productivity and nutrient removal capacity between algal CSTR
and algal MBR at the same light level under practical and optimal conditions. Ecological Engineering, 93,
66–72, https://doi.org/10.1016/j.ecoleng.2016.04.008
Ta şkan B., Hasar H. and Lee C. H. (2020). Effective biofilm control in a membrane biofilm reactor using a
quenching bacterium (Rhodococcus sp. BH4). Biotechnology and Bioengineering, 117(4), 1012–1023, https://
doi.org/10.1002/bit.27259
Thanh B. X., Visvanathan C. and Aim R. B. (2013). Fouling characterization and nitrogen removal in a batch
granulation membrane bioreactor. International Biodeterioration and Biodegradation, 85, 491–498, https://
doi.org/10.1016/j.ibiod.2013.02.005
Tirosh U. and Shechter R. (2020). Membrane aerated biofilm reactor (MABR)—distributed treatment of
wastewater at low energy consumption. Frontiers in Water-Energy-Nexus – Nature-Based Solutions,
Advanced Technologies and Best Practices for Environmental Sustainability, 513–515, https://doi.
org/10.1007/978-3-030-13068-8_128
Truong H. T. B., Nguyen T., Thi P. and Bui H. M. (2018). Integration of aerobic granular sludge and membrane
filtration for tapioca processing wastewater treatment: fouling mechanism and granular stability. Journal
of Water Supply: Research and Technology – Aqua, 67(8), 846–857, https://doi.org/10.2166/aqua.2018.104
US EPA. (2006). Wastewater Management Fact Sheet: Energy Conservation. EPA 832-F-06-024.
Vijayalayan P., Thanh B. X. and Visvanathan C. (2014). Simultaneous nitrification denitrification in a batch
granulation membrane airlift bioreactor. International Biodeterioration and Biodegradation, 95, 139–143,
https://doi.org/10.1016/j.ibiod.2014.05.020
Vinardell S., Astals S., Jaramillo M., Mata-Alvarez J. and Dosta J. (2021). Anaerobic membrane bioreactor
performance at different wastewater pre-concentration factors: An experimental and economic study.
Science of the Total Environment, 750, 141625, https://doi.org/10.1016/j.scitotenv.2020.141625
Wan C. F. and Chung T.-S. (2015). Osmotic power generation by pressure retarded osmosis using seawater brine as
the draw solution and wastewater retentate as the feed. Journal of Membrane Science, 479, 148–158, https://
doi.org/10.1016/j.memsci.2014.12.036
Wan C. F., Cui Y., Gai W. X., Cheng Z. L. and Chung T.-S. (2020). Nanostructured membranes for enhanced forward
osmosis and pressure-retarded osmosis. In Sustainable Nanoscale Engineering. Elsevier, pp. 373–394.
Wan C. F., Jin S. and Chung T.-S. (2019). Mitigation of inorganic fouling on pressure retarded osmosis (PRO)
membranes by coagulation pretreatment of the wastewater concentrate feed. Journal of Membrane Science,
572, 658–667, https://doi.org/10.1016/j.memsci.2018.11.051
Wang P., Cui Y., Ge Q., Tew T. F. and Chung T.-S. (2015). Evaluation of hydroacid complex in the forward osmosis–
membrane distillation (FO–MD) system for desalination. Journal of Membrane Science, 494, 1–7, https://doi.
org/10.1016/j.memsci.2015.07.022
Wang S., Chew J. W. and Liu Y. (2020). Development of an integrated aerobic granular sludge MBR and reverse
osmosis process for municipal wastewater reclamation. Science of the Total Environment, 748, 141309,
https://doi.org/10.1016/j.scitotenv.2020.141309
Winkler M. K. and Straka L. (2019). New directions in biological nitrogen removal and recovery from wastewater.
Current Opinion in Biotechnology, 57, 50–55, https://doi.org/10.1016/j.copbio.2018.12.007
Xie Z., Wang Z., Wang Q., Zhu C. and Wu Z. (2014). An anaerobic dynamic membrane bioreactor (AnDMBR) for
landfill leachate treatment: performance and microbial community identification. Bioresource Technology,
161, 29–39, https://doi.org/10.1016/j.biortech.2014.03.014
Xu B., Albert Ng T. C., Huang S. and Ng H. Y. (2020a). Effect of quorum quenching on EPS and size-fractioned
particles and organics in anaerobic membrane bioreactor for domestic wastewater treatment. Water
Research, 179, 115850, https://doi.org/10.1016/j.watres.2020.115850
Xu B., Albert Ng T. C., Huang S., Shi X. and Ng H. Y. (2020b). Feasibility of isolated novel facultative quorum
quenching consortiums for fouling control in an AnMBR. Water Research, 169, 115251, https://doi.
org/10.1016/j.watres.2019.115251
Yang Q., Li H., Wang D., Zhang X., Guo X., Pu S., Guo R. and Chen J. (2020). Utilization of chemical wastewater
for CO2 emission reduction: purified terephthalic acid (PTA) wastewater-mediated culture of microalgae for
CO2 bio-capture. Applied Energy, 276, 115502, https://doi.org/10.1016/j.apenergy.2020.115502
Yangali-Quintanilla V., Li Z., Valladares R., Li Q. and Amy G. (2011). Indirect desalination of Red Sea water with
forward osmosis and low pressure reverse osmosis for water reuse. Desalination, 280(1–3), 160–166, https://
doi.org/10.1016/j.desal.2011.06.066
Yue X., Koh Y. K. K. and Ng H. Y. (2018). Membrane fouling mitigation by NaClO-assisted backwash in anaerobic
ceramic membrane bioreactors for the treatment of domestic wastewater. Bioresource Technology, 268, 622–
632, https://doi.org/10.1016/j.biortech.2018.08.003
Zhang S., Wang P., Fu X. and Chung T.-S. (2014). Sustainable water recovery from oily wastewater via forward
osmosis-membrane distillation (FO-MD). Water Research, 52, 112–121, https://doi.org/10.1016/j.
watres.2013.12.044
Zhang W., Liang W., Zhang Z. and Hao T. (2020). Aerobic granular sludge (AGS) scouring to mitigate membrane
fouling: performance, hydrodynamic mechanism and contribution quantification model. Water Research,
188, 116518, https://doi.org/10.1016/j.watres.2020.116518
Zhou Y., Huang M., Deng Q. and Cai T. (2017). Combination and performance of forward osmosis and membrane
distillation (FO-MD) for treatment of high salinity landfill leachate. Desalination, 420, 99–105, https://doi.
org/10.1016/j.desal.2017.06.027
Chapter 12
Natural treatment systems and
integrated watershed management
for decarbonization
Hannah R. Molitor and Jerald L. Schnoor*
Department of Civil and Environmental Engineering, University of Iowa, Iowa City, IA, USA
*Correspondence: [email protected]
12.1 INTRODUCTION
Water and wastewater utilities use tremendous amounts of energy and emit copious quantities of
greenhouse gases (GHGs) through direct emissions from the facilities themselves (Scope 1 emissions),
purchased electricity and energy from outside suppliers (Scope 2 emissions), and emissions related
to the usage of water by customers (e.g., hot water) and energy/transportation costs associated with
moving equipment and personnel to the site each day (Scope 3 emissions).
For common emission estimates that cover Scopes 1 and 2, water and wastewater utilities are
estimated to emit 3–7% of all global greenhouse gas emissions (Trommsdorff, 2015). California is the
most extreme example and uses 20% of all the state’s electricity just in supplying water (Loge, 2016).
Water is a heavy commodity and moving it around via pumping uses tremendous amounts of electricity.
If the electricity is supplied by highly polluting coal-fired power plants, this especially increases the
carbon footprint of utilities. Aeration of wastewater by blowers and sparged air represents another huge
energy investment and a concomitant emission of greenhouse gases. Finally, treatment of carbonaceous
biochemical oxygen demand (BOD) and nitrogen in wastewater results in direct emissions of carbon
dioxide (CO2), methane (CH4) and nitrous oxide (N2O) that are potent biogenic greenhouse gases. Flaring,
the purposeful burning of digester gas, poses another highly-emitting CO2 operation by some utilities.
As alternatives to or subsequent processes to engineered systems, natural treatment systems (NTS)
have great potential to decarbonize the wastewater treatment by leveraging microbial and/or plant
communities. NTS can be used as tertiary (polishing) operations at wastewater treatment plants; or
they can be utilized separately to sequester carbon from the atmosphere (negative emissions), treat
stormwater runoff for removal of metals and organic contaminants (green infrastructure), or prevent
erosion and runoff of nutrients and pesticides in agricultural applications. NTS are engineered to
use a minimal amount of energy and mechanization. Instead, they utilize soil, plants/algae, bacteria,
and fungi to achieve sequestration of metals, carbon storage in wood and soils, or biodegradation
of organic contaminants to innocuous end-products (e.g., H 2O, CO2, HCl). They tend to be low cost
with low energy consumption, low emissions, and aesthetically pleasing to the public. Consequently,
judicious incorporation of NTS into the water and wastewater treatment sectors can mitigate GHGs
with low-energy, economical technologies that are adaptable to most regions of the world.
© 2022 The Editors. This is an Open Access eBook distributed under the terms of the Creative Commons Attribution Licence
(CC BY-NC-ND 4.0), which permits copying and redistribution for non-commercial purposes with no derivatives, provided the
original work is properly cited (https://creativecommons.org/licenses/by-nc-nd/4.0/). This does not affect the rights licensed or
assigned from any third party in this book. The chapter is from the book Pathways to Water Sector Decarbonization, Carbon Capture
and Utilization, Zhiyong Jason Ren and Krishna Pagilla (Eds.)
Johnston and Karanfil (2013) estimated the greenhouse gas emissions associated with seven utilities
in the southeastern US and determined an average of 1240 kg carbon dioxide equivalents (CO2-eq)/
million gallons (MG) when Scopes 1 and 2 emissions were evaluated. On an energetic basis, they
found energy use in the water cycle to vary from 1250 to 6500 kWh/MG with wastewater treatment
processes using the most energy. Energy required to treat and distribute drinking water ranged from
250 to 3500 kWh/MG.
Indicative of operations which may cause greenhouse gas emissions in the utility industry, Figure
12.1 shows the conveyance of water in drinking water and wastewater processes. Total energy
utilization for all the operations ranges from 1030 to 36 200 kWh/MG (Griffiths-Sattenspiel & Wilson,
2009). Pumping, customer usage (Scope 3 emissions) and wastewater treatment represent large energy
inputs and GHG emissions, thus, opportunities for GHG reductions. Meanwhile sludge handling
and natural treatment systems offer the potential for large energy savings – even net zero emissions
by virtue of utility-generated electricity, combined heat and power (CHP), and carbon sequestration
(negative emissions). Obviously, one of the first considerations should be to power the entire system
with renewable energy to the maximum extent possible (solar, hydro, biomass and wind) or low-
carbon electricity provided by nuclear power. If space is available onsite, solar panels for powering the
facility avoids Scope 2 emissions, a highly attractive investment. Water conservation, fewer chemicals,
less pumping, and more gravity flow offer other options to reduce GHG emissions (Erickson et al.,
2008). In wastewater processes, anaerobic operations result in less sludge production and more
methane production which can be used for process heating or microturbine electricity generation.
Process modifications such as anaerobic membrane bioreactors can be run in colder climates than
previously thought, which offers an alternative to achieve net zero emissions (McCarty et al., 2011).
Burning biosolids (biomass) to make district heat and electricity is another possibility in addition
to the application of biosolids onto nearby agricultural land for co-benefits of soil conditioning and
carbon sequestration.
One of the keys to reducing greenhouse gas emissions in the water industry is to view the entire
water cycle holistically – the One Water movement (One Water Hub|US Water Alliance). When we
pump-up groundwater or surface water as the source for drinking water, we begin the cycle. Drinking
water becomes wastewater (used water), and then used water is returned to streams or groundwater
after appropriate treatment, thus completing the cycle. Drinking water becomes wastewater, and
wastewater becomes drinking water.
Wastewater
Sludge handling
Source Water Intake Treatment (aeration,
(energy and nutrient
(pumping) pumping, algae CO2
recovery)
uptake)
Natural Treatment
Water Treatment Sewage Collection (created wetlands,
Processes and Conveyance land systems, C-
sequestration)
Figure 12.1 Water and wastewater utilities general flowscheme. Energy is required for each step in the process
with concomitant greenhouse gas emissions.
Figure 12.2 Water and wastewater treatment as a part of the One water cycle (withdrawal, drinking water treatment,
wastewater treatment, recovery of nutrients/water/energy, fertigation onto crops, and recharge of groundwater and
surface water).
In Integrated Watershed Management, drinking water treatment plants and wastewater treatment
and recovery are designed within the context of the entire system, the One Water Cycle. We can create
a circular economy around the use of water. Figure 12.2 depicts the water cycle, illustrating some
opportunities for conservation, integrated management, and reduction of carbon emissions. Under
integrated watershed management, wastewater treatment is more properly termed ‘water reclamation
and reuse’ where nutrients, energy, and water are recovered, and greenhouse gas emissions are
mitigated. By reusing water, the withdrawals of source water are decreased, thus reducing the energy
and carbon footprint of the entire water cycle. Water is reused for graywater applications, irrigation/
fertigation, aquifer recharge, and even for direct potable reuse. By treating and infiltrating stormwater,
aquifers are replenished and more water is available for reuse. Some treated wastewater is also
available for aquifer storage and recovery via infiltration basins. Excess nutrients from wastewater
can be applied as irrigation water onto crops (fertigation) as illustrated in Figure 12.2.
In this chapter, we describe natural treatment system technologies, their advantages and
disadvantages, and their potential to decarbonize the water sector when incorporated in integrated
watershed management. We emphasize phytoremediation and microalgal cultivation in two engineering
research examples.
The US EPA defines natural treatment systems as those having minimal dependence on mechanical
elements to support the wastewater treatment process. Instead, NTS systems use plants, bacteria,
archaea, fungi and/or algae to break down and neutralize pollutants in wastewater. Often, these
natural components work symbiotically with each other. For example, the bacteria on roots of plants
may break down wastewater organics while supplying nutrients for the plant and allowing the plant
to fix carbon dioxide from the atmosphere which becomes incorporated into woody tissue and soils
(carbon sequestration). Natural systems may include composting of biosolids and employ other small
animals in the treatment scheme like nematodes, earthworms, or fly larvae.
Through site remediation and wastewater treatment, NTS protect public and environmental
health. NTS include free water surface wetlands (FWS), subsurface flow (SF), vertical subsurface
flow (VSSF), and horizontal subsurface flow (HSSF) constructed wetlands (CW), biofilters (BF),
waste stabilization ponds (WSP), and land irrigation of wastewater onto plantations of trees and
grasses (PHYTO). With the goal of conserving water and nutrients, groundwater can be recharged
through infiltration basins, wastewater nutrients are applied onto cropland (fertigation), and biosolids
applications onto land provide soil conditioning and carbon sequestration. Such NTS systems can be
employed in-series following primary and secondary treatment for small communities and developing
countries, or they can be used as ‘polishing’ or tertiary treatment on the back-side of conventional
wastewater treatment facilities. As a tertiary treatment or effluent polishing step, NTS can remove
antimicrobial drugs that may otherwise be released by conventional wastewater treatment systems in
microdoses that can confer antibiotic-resistance to pathogenic bacteria (Ryan et al., 2011). In the same
way, NTS can also sequester or degrade anthropogenic compounds that are toxic to aquatic systems.
Where NTS use photosynthetic organisms, waters or polluted soils can be simultaneously treated and
oxygenated, thereby improving water quality or supporting aerobic environments for further pollutant
degradation.
Beyond the value of low-cost, low-energy, decarbonizing treatment, certain NTS provide marketable
products that offset treatment costs or energy requirements. Atmospheric carbon dioxide is first fixed
as biomass, and then can be used for biofuels, fertilizer, feed, biochar feedstock, fiber for pulp or paper,
or burned directly to generate power. However, the expense of biomass transport and access to quality
storage options can be prohibitive. Biomass can also be converted to biochar, a stable product for
storage, through pyrolysis. As a soil additive, biochar increases the health of soil through increased
water and nutrient capacity and subsequently improves the carbon sequestration abilities of the soil
(Saeid & Chojnacka, 2019).
Natural treatment systems can be simpler, more cost-effective, efficient, and reliable than
conventional treatment infrastructure. NTS are particularly attractive as a means to meet wastewater
treatment standards, remove nutrients and micropollutants, and sequester or utilize carbon because
they require less capital and operational investment than conventional methods (Mahmood et al.,
2013). The systems are relatively inexpensive to install and rarely rely upon the chemical inputs or
mechanical parts necessary for ‘gray infrastructure,’ making them effective for areas that have limited
access to power, specialized equipment, or skilled workers. These qualities make NTS appealing to
small communities or those in developing countries, as it facilitates effective wastewater treatment
and avoids a substantial investment that many communities cannot afford.
Of course, limitations exist such as the high degree of treatment specified in some effluent discharge
permits that may be difficult to achieve. Cold weather, temperate climates, and seasonal events (floods,
droughts) may also limit the application of NTS. However, NTS have great potential and adaptability
to various environments. The wide range of technologies included in the NTS umbrella naturally
cover a wide range of treatment scenarios; NTS are used to treat wastewater, stormwater, agricultural
runoff, and contaminated sites. The ability of NTS biomass to utilize nitrogen and phosphorous makes
it extremely useful for remediating high nutrient concentrations in wastewater and agricultural runoff.
Additionally, specialized NTS can sequester or transform certain hazardous organic species and heavy
metals, which makes them suitable to treat contaminated sites. NTS can also act as a stormwater filter
and remove pollutants from stormwater runoff before it enters rivers, lakes, or groundwater. For each
of these treatment scenarios, NTS employ mechanisms long-established in the natural world that
today’s engineers now channel into designed systems.
Figure 12.3 A typical horizontal flow constructed wetland treats conventional pollutants like BOD and nutrients
following primary treatment in rural settings or as a tertiary/polishing step following secondary treatment. Toxic trace
organics from pharmaceuticals, consumer-care products, industrial chemicals, and pesticides may be degraded in
such systems by biological processes and photolysis. Adapted from https://waterpurificationengineering.weebly.
com/constructed-wetlands.html.wetlands.html
In a wide literature analysis, Mander et al. (2014) found that free water surface (FWS) constructed
wetlands had substantially lower CO2 emissions than subsurface flow CWs, 95.8–137.0 mg m−2 h−1,
respectively. Methane emissions ranged from 3.0 to 6.4 mg m−2 h−1, while N2O release rates were
small but significant at 0.09–0.13 mg m−2 h−1. From this reference, it can be estimated that constructed
wetlands have lower greenhouse gas emissions than conventional forms of wastewater treatment like
activated sludge. They contribute a smaller net source of total greenhouse gases on a CO2-equivalent
basis: 191–332 mg CO2-eq m−2 h−1. Roughly 41–50% of global warming potential (GWP) was due to
CO2, 44–54% due to CH4, and 5–6% due to N2O emissions.
Overall, created wetlands may constitute a source or a sink of greenhouse gas emissions. It depends
on the wastewater influent concentrations (BOD, NH4+, NO3−), flow design and size of unit operations,
and the timescale of the analysis. In one particularly careful study, measured methane emissions were
highly variable but indicated a significant relationship with temperature and the density of vegetation.
Average methane emissions in the vegetation were 7.8 at 15°C and 24.5 mg m−2 h−1 at 24°C, respectively.
Nitrous oxide emissions ranged from 0.5 to 1.9 g m−2 y−1. Net carbon dioxide sequestration was
measured as 0.27–2.4 kg m−2 y−1 which represented 12–67% of the CO2 photosynthesized into biomass.
N2O emissions were a significant fraction of total GHG emissions (12–29%) according to de Klein and
van der Werf (2014). For this example, the constructed wetland was a net sink for greenhouse gases
expressed on a CO2-equivalent basis of 30.8–274 mg CO2-eq m−2 h−1. CO2 was photosynthesized into
plant biomass while N2O and CH4 represented GHG emissions, which were more than offset by carbon
sequestration. Such a study demonstrates the potential for created wetlands to serve as a net sink for
greenhouse gases resulting in ‘negative emissions’.
Floating mats of plants can also be used to create superior removal of pollutants from free water
surface wetlands and ponds. According to Pavlineri et al. (2017), floating wetlands removed 58%
of total nitrogen, 48.75% of total phosphorus, 72.8% total NH4 -H, and 57.8% of chemical oxygen
demand (COD). Floating wetlands may also sequester CO2 into plant biomass, but a net greenhouse
gas analysis was not performed in this study.
An example of a horizontal created wetland is shown at the Iowa Army Ammunition Plant in
Middletown, Iowa, in Figure 12.4. Contaminated groundwater and effluent from the nearby munitions
factory were treated by a 2-acre free-surface wetland planted with native vegetation, especially
Sagittaria spp. (common name, arrowhead). The munitions factory was making C4-explosive and
wastewater contained ppm quantities of RDX, TNT, and HMX. RDX was the most problematic of the
toxic chemical pollutants due to its high water solubility, persistence, and mobility in groundwater.
After constructing the controlled outlet through a dam and release structure, the constructed wetland
was successful in meeting the discharge permit for the wastewater plant of just 2 µg/L (ppb) due to
photolysis of RDX and biological degradation by plants and associated microorganisms. Removal
was attributed to photolysis of RDX and biological degradation by wetland plants and associated
microorganisms. Carbon dioxide was sequestered into plant biomass and phytoplankton.
Figure 12.4 Aerial view (left) and photo of the inlet to the horizontal flow, free surface pond at the iowa army
ammunition plant in middletown, iowa. It is planted with arrowhead (native vegetation) and resulted in the treatment
of RDX contaminant to meet required effluent discharge permit.
harvesting) (Steinmann et al., 2003), sludge buildup, uncontrolled effluent ammonia concentrations,
strong odors during spring or fall inversions, and the creation of habitat for insect vectors (e.g.,
mosquitos). Uncovered anaerobic lagoons for animal manure treatment are particularly problematic
as a major source of CH4 and a small contributor of N2O; in fact, manure management accounted for
9.7% of all United States anthropogenic CH4 emissions in 2018 (Desai & Camobreco, 2020).
In consideration of the shortcomings of lagoon systems for wastewater treatment, research is
ongoing to optimize alternative wastewater treatment strategies that will meet stringent nutrient
discharge regulations yet still be affordable for small communities and farms. This is an opportunity
to implement technologies that will reduce GHG emissions, produce salable goods, and better protect
receiving waters.
heavy metals. Activated carbon’s removal efficacy is often predicted by a compound’s hydrophobicity
but can also depend on pore diffusivity (molecular volume), electrostatic/pi-pi interactions, and
hydrogen donor/acceptor interactions with specific surface groups (Webb et al., 2020). Biofiltration
is also inadvertently used in drinking water treatment where pollutant-degrading microbes colonize
granular or powdered activated carbon for contaminant removal.
Figure 12.5 Phytoremediation schematic for the processes of photosynthesis and storage of carbon from the
atmosphere into woody biomass and soils via root turnover. Other processes include rhizosphere biodegradation, plant
uptake, translocation, plant phytotransformation, phytovolatilization, and photodegradation of organic chemicals.
grasses. Dense, shallow roots of the grasses and deep penetrating roots of the trees can facilitate
effective treatment of nutrients, fine particles, and organic chemicals to very low target concentrations.
Organic chemicals are degraded by rhizosphere bacteria and by the plant itself; nutrients (N,P) are
taken-up and removed by the plant; and carbon dioxide is sequestered from the atmosphere into
woody biomass and as organic carbon in soils (Figure 12.5). Volatilization may also play a treatment
role in any NTS that employs plants, but it is especially prominent in those applications for treatment
of volatile organic contaminants with high rates of transpiration regulated by climate, soil and plant
types. Under these conditions, contaminants are absorbed through the roots, along with water and
nutrients, and carried up the xylem to the plant’s stomata where volatile organic compounds can
change from the liquid to gas phase. Upon volatilization, some species are photochemically degraded
while others may persist for hours or more in the atmosphere.
Regenerative agriculture refers to the use of best management practices (BMPs) to restore and
preserve biodiversity and soil quality – it is a form of sustainable farming that values the long-term
productivity of soil organic carbon and fertility. BMPs include no tillage, diverse cover crops, multiple
crop rotations and legumes, intercropping, recycling of manure through beneficial grazing, minimizing
synthetic fertilizers and pesticides, use of perennial crops such as energy crops, silvopasturing
(integrating trees, forage, and grazing livestock), and restoring and creating wetlands. Such practices
have the potential to reduce erosion and add soil organic carbon where it has been badly depleted in
the past. Due to the large area of arable land in the world, it is possible to sequester highly significant
amounts of carbon dioxide from the atmosphere each year through regenerative agriculture. Minx
et al. (2018) estimate 2.3 Gigatons CO2/yr (2.3 billion metric tons/yr = 2.3 × 1015 g CO2/yr) could be
removed from the atmosphere by soil carbon sequestration which represents roughly 6% of total
annual anthropogenic emissions currently.
Water utilities could partner with farmers in their watershed or urban landowners in the sewershed
to create negative emissions and sequester carbon dioxide from the atmosphere. Such partnerships
could involve utilities supplying irrigation water and fertigation of crops resulting in the removal of
nitrogen and phosphorus to meet stringent discharge permits. This form of ‘water quality trading’
Table 12.1 Potential to reduce CO2 and other greenhouse gas (GHG) emissions from the water and wastewater
utility industry by negative emissions from natural treatment systems (or solar panels).
within the watershed is also possible whereby farmers sequester carbon into agricultural soils much
more cheaply than utilities can remove or reduce it. A direct payment to farmers for the service of
negative emissions is also possible. The goal of the utility is to recover as much resource from their
wastewater as possible including dollars, water, nutrients, and energy while avoiding greenhouse gas
emissions or creating negative emissions as off-sets.
Typical CO2 emissions from one million gallons (1.0 MG) of water withdrawn, used, treated, and
discharged by utilities (as in Figure 12.1) could be off-set by 1680–62 000 m 2 of farmland practicing
regenerative agriculture for the sequestration of carbon into soils (Lal, 2004, 2015; Minx et al., 2018;
Tellatin & Myers, 2018). Only a relatively small portion of the CO2 emitted by utilities would be
considered biogenic emissions, and the majority is from energy requirements to pump the water, run
the plants and treat the water. Based on the results in Table 12.1, it is clear that considerable land
is required to create enough negative emissions to off-set greenhouse gas emissions from water and
wastewater utilities. The most land would be required for regenerative agriculture and the least for
hybrid poplar buffer strips. Created wetlands show promise also, but the literature is mixed as to their
net benefits (carbon sinks or sources).
One million gallons of water produced by the water utility is a rather small volume of water by utility
standards. Even a small community of 5000 people using a total of 200 gallons per capita-day would
generate 1.0 million gallons per day (MGD) or 365 MG per year. Multiplying the areas in Table 12.1
(third column for hybrid poplar buffers) by 365 yields a required spatial area of 15.4–2260 ha – quite
a lot of land for a small utility to control. It is not practical to off-set 100% of the greenhouse gas
emissions from water utilities by NTS; rather it suggests one option for utilities to consider reducing
their carbon footprint.
To illustrate, the last row in Table 12.1 supposes that the utility purchases or partners to obtain
solar power for electrifying its operations or, alternatively, to off-set their Scope 2 greenhouse gas
emissions. The relatively small area (4.8 m 2) of solar panels required to off-set emissions from 1 MG
of water suggests a viable option for the utility to reduce its carbon footprint. For the 1 MGD water
utility (365 MG per year), only 1750 m 2 (0.175 ha) of solar panels would be required to off-set 100%
of the carbon footprint. Utilizing NTS would require much more land. Thus, most water utilities
which have reduced their carbon footprints have done so through: (1) electrifying pumping and other
operations with solar power and battery storage; (2) upgrading anaerobic digestion of wastewater
solids; (3) producing power from methane digester gas by microturbines; (4) producing combined heat
and power (CHP) by internal combustion engines using digester gas; (5) improving sludge stabilization
and biosolids applications or reducing landfilling; (6) implementing demand side management with
water conservation by smart water metering and variable pricing campaigns.
Indeed, Wong and Law-Flood (2011) have provided examples of several water utilities which have
successfully reduced their carbon footprints, mainly by anaerobic digestion of wastewater solids and
utilization of the gas (Wong & Law-Flood, 2011). Examples include: Sheboygan, Wisconsin achieved
35–50% reduction in energy; Nashua, New Hampshire saved $750 000 in annual electricity and
landfill costs from digester and biosolids improvements and 20% reduction in energy; Gloversville-
Johnstown, New York combined heat and power (CHP) system generated 100% of total energy needs
on-site; Essex Junction, Vermont used microturbines and CHP to provide 37–39% of its total energy
needs; Pittsfield, MA employed digester gas for CHP internal combustion engines and microturbines
to save 29% of energy needs; East Bay Municipal Utility District, California used internal combustion
engines for CHP and saved 90%; and Fairhaven, MA increased their organic solids loading to augment
volume of digester gas resulting in CHP to save 73% of energy.
Figure 12.6 Hybrid poplar plantation after seven years of growth at Amana, iowa, to treat runoff of pesticides and
nutrients from agriculture row crops while sequestering carbon dioxide into woody biomass and soils.
biomass would require a CO2-sequestration rate from the atmosphere of 200–208 metric tons CO2
per hectare over seven years. On average annually, the CO2-sequestration into woody biomass was
therefore equivalent to 27.5–29.3 metric tons CO2 ha−1 y−1 as shown in Table 12.1. Forests in the US
sequester 3.57–5.03 metric tons carbon per hectare per year in above-ground biomass. (Ney et al., 2005)
Unmanaged forests in Iowa averaged 5.06 metric tons C ha−1 y−1, (Ney et al., 2002) with a range of of
3.7–7.0 metric tons CO2 ha−1 y−1 (Table 12.1). However, most of the carbon stock in the riparian zone
buffer strip shown in Figure 12.6 is stored in the soils (below ground). Average soil organic carbon in
the top 30 cm of soil was 5.35%. (Ney et al., 2005) However, the amount sequestered into below-ground
soils on an annual basis is relatively small compared to the above-ground woody biomass sequestration.
In native forests of Iowa, carbon stocks are estimated to be 137.3 metric tons C per ha, of which
60.8% is in soils (top 30 cm), 24.5% stored in above-ground biomass in trees, 3.8% resides in trees
below-ground (roots), 10.4% lies in carbon on the forest floor, and 0.5% is comprised of understory
vegetation (Ney et al., 2002). By far the most carbon in native Iowa forests resides in below-ground
soils and roots, 83.6–90.7 metric tons C ha−1. Above-ground carbon content (in trees, understory, and
the forest floor) ranges from 35.2 to 61.9 metric tons C ha−1 with the greatest amount in native oak-
hickory forests. Soil carbon is a large pool which builds-up and oxidizes very slowly but can be lost
by erosion relatively quickly over decades to centuries. Net benefits of reforestation of native forests
in Iowa include the possibility to remove 7.0 metric tons CO2 ha−1 y−1 of which 6.21 metric tons CO2
ha−1 y−1 goes into above-ground forest biomass and 0.79 metric tons CO2 ha−1 y−1 sequesters into below-
ground soil and roots (Ney et al., 2002).
12.4 CASE STUDY: POWER PLANT FLUE GAS AND FERTILIZER WASTEWATER
TREATMENT BY NUTRITIOUS MICROALGAE
Microalgae are a promising alternative livestock feed source, with greater resource-to-biomass
conversion efficiency than conventional agricultural crops (corn, soy, wheat, etc.). High-protein
microalgae are of particular promise because they are able to remove nitrogen and phosphorus from
wastewater, fix CO2 and other pollutants from flue gas, tolerate saline water, and occupy less land
footprint per unit of produced biomass. This NTS resource recovery scheme simultaneously treats
waste, produces valuable biomass, and decarbonizes agriculture and waste treatment systems.
Table 12.2 Nutrient and flue gas component utilization rates by S. obliquus grown with simulated coal-fired
power plant flue gas and triple-nitrogen Bold’s Basal Medium during the exponential growth phase.
In our research, we seek to inform technology scale-up for power plant and industrial flue gas
treatment, and wastewater treatment, by microalgae. At bench scale, we modeled substrate inhibition
of microalgal biomass productivity by CO2, stimulated microalgal growth with simulated coal-fired
power plant emissions, and enhanced microalgal settleability with simulated emissions through
increased production of extracellular polymeric substances (EPS) (Table 12.2).
Figure 12.7 Schematic of treatment of 1 MGD fertilizer plant wastewater and CO2 sequestration from power
plant flue gas by microalgal photobioreactors. Treatment processes (clarification, PBRs, and final clarification) and
followed by biomass harvesting.
The wastewater is treated through primary clarification, a vertical manifold PBR with biomass
recycle, and final clarification while a co-located power plant (889 800 m3/d; 12% CO2) supplies CO2
to the PBR. Biomass dewatering and drying is achieved through centrifugation and drum drying. The
influent solids concentration is first reduced using two rectangular clarifiers, each 12 × 4 m and 4 m
deep with 2.4 h hydraulic residence time (Metcalf & Eddy, 2003). The clarifier effluent is then fed to the
PBR where nitrogen and phosphorus are removed by S. obliquus (dry biomass composition: 50.7 ± 0.1%
C, 6.44 ± 0.04% N, 1.0 ± 0.1% P) (Molitor & Schnoor, 2020). The PBR system was designed to meet
generic wastewater P effluent limits of 1.0 mg/L, as the system is barely N-limited, with maximum
culture density of 2 g/L (Acién et al., 2012). The designed solids retention time and hydraulic retention
time are 6.7 and 1.6 d, respectively. Consequently, the PBR requires 6100 m3 of working volume and a
recycle rate of 310 m3/d of 20.5 g/L biomass from the final clarifier.
To achieve a surface overflow rate of 27 m3/m 2/d, a final clarifier is 14 m in diameter with side
water depth of 3.7 m because the microalgae have a relatively rapid bulk settling rate of 0.058 m/
min (measured settling rate of 1 m/min adjusted to account for a fraction of the biomass settling at
that rate). The possibility of bulk settling in this scenario, facilitated by high concentrations of EPS,
is unique and advantageous to efficient biomass harvesting. The settled biomass that is harvested
(90 m3/d), rather than recycled to the PBRs, is dewatered through centrifugation in two centrifuges
with capacities of 4 m3/d and operating at 11 h/d (Tredici et al., 2016). The centrifuges produce
microalgal paste of approximately 20% solids, which is then dried to approximately 5% solids through
four drum dryers (Tang et al., 2003).
The treatment scheme has effluent concentrations of 5.6 mg/L N and 1.0 mg/L P, produces
1840 kg/d dried microalgal biomass, and sequesters 4550 kg/d CO2.
Microalgae are well suited to treat high nitrate wastewaters, without carbon source supplementation,
which is generally considered challenging for other microbes (Pinar et al., 1997). Fortunately, there are
thousands of sources of domestic secondary effluent, some of which are co-located with CO2-emitting
power plants. Beyond municipal wastewater, industrial and agricultural wastewater streams with low
organic carbon and high nitrogen concentrations (such as explosives factory wastewater, fertilizer
plant wastewater, agricultural run-off, and irrigation return waters) would be strategic options for
microalgal substrate (Ji et al., 2018). However, heavy metal and pathogen contamination from certain
wastewaters could impede production of microalgae for animal feed.
remainder of the SO2 (1140 kg/d) will be rapidly oxidized and accumulate at a rate of approximately
200 mg/L/d in the wastewater to a non-inhibitory concentration of 280 mg/L SO42−. If electricity were
to be provided by wind turbines, at a GHG emission rate of 1.8 × 10−2 kg/kWh CO2 (Alsaleh & Sattler,
2019), the projected net GHG emissions from the proposed scheme would be −4300 kg/d CO2-eq.
The corresponding land footprint is 11.6 acres based on a literature-sourced land-footprint-to-
cultivation-volume ratio of 7.7 m 2/m3, the value for a full-scale 1300 m3 -system of vertical manifold
PBRs designed by A4F-AlgaFuel, S.A. to treat cement plant flue gases (Torzillo & Chini Zittelli,
2015). Systems external to the PBRs are included in the area-to-cultivation-volume ratio, though the
PBRs account for the vast majority of the treatment facility footprint. As technologies for wastewater
treatment with microalgae progress, it can be expected that areal productivity will increase significantly
through improved understanding of cultivation conditions for increased nutrient uptake rates and
optimal HRT and SRT.
Table 12.3 Summary of natural treatment system mechanisms, advantages, disadvantages, and means of decarbonization.
required
Natural treatment systems and integrated watershed management for decarbonization 223
increase nutrient uptake rates). Combining NTS with conventional wastewater treatment is an
attractive method for improving treatment, meeting more stringent permit requirements for N and P,
and at the same time capturing carbon via photosynthesis.
Natural treatment systems incorporated into traditional water and wastewater operations can be
further integrated with watershed management to reduce the carbon footprint of water utilities with
the ultimate goal of ‘net zero emissions’. Figure 12.8 illustrates an integrated watershed approach
where the water cycle is monitored to reduce costs and greenhouse gas emissions, restore soil carbon,
enhance water quality, and conserve water while replenishing aquifers. Small-scale distributed water
treatment facilities and wastewater treatment plants are instrumented with sensors to continuously
monitor the state of the system in concert with the One-Water concept. NTS are integrated into the
water and wastewater operations to provide for carbon sequestration, the uptake of excess nutrients
and biodegradation of xenobiotic organic chemicals.
REFERENCES
Acién F. G., Fernández J. M., Magán J. J. and Molina E. (2012). Production cost of a real microalgae production
plant and strategies to reduce it. Biotechnology Advances, 30(6), 1344–1353, https://doi.org/10.1016/j.
biotechadv.2012.02.005
Alsaleh A. and Sattler M. (2019). Comprehensive life cycle assessment of large wind turbines in the US. Clean
Technologies and Environmental Policy, 21(4), 887–903, https://doi.org/10.1007/s10098-019-01678-0
Bowman R. H., Gloyna E. F., Middlebrooks E. J., Pearson G. F., Reed S. and Reid L. C. (2002). Wastewater
Technology Fact Sheet: Facultative Lagoons. Office of Water, United States Environmental Protection
Agency, Washington, DC, USA.
Crites R. W., Middlebrooks E. J., Bastian R. K. and Reed S. C. (2014). Natural Wastewater Treatment Systems. CRC
Press, Boca Raton, FL, USA.
de Klein J. J. M. and van der Werf A. K. (2014). Balancing carbon sequestration and GHG emissions in a constructed
wetland. Ecological Engineering, 66, 36–42, https://doi.org/10.1016/j.ecoleng.2013.04.060
eGRID. (2018). University of Iowa Main Power Plant. Emissions & Generation Resource Integrated Database
(eGRID). Office of Atmospheric Programs, Clean Air Markets Division, Washington, DC, USA.
Desai M. and Camobreco V. (2020). Inventory of US Greenhouse Gas Emissions and Sinks: 1990–2018. United
States Environmental Protection Agency, Washington, DC, USA.
Erickson D., Orrett E. B. and Rosenblum J. (2008). Greenhouse Gas Emissions Related to Water and Wastewater
Services: Baseline, Reduction Strategies, and Recommendations. City of Santa Rosa Utilities Department,
Santa Rosa, CA, USA, p. 167.
Fasaei F., Bitter J. H., Slegers P. M. and van Boxtel A. J. B. (2018). Techno-economic evaluation of microalgae
harvesting and dewatering systems. Algal Research, 31, 347–362, https://doi.org/10.1016/j.algal.2017.
11.038
Gassan H., Eugenia M. M. and Sebastián S. (2010). Influence of temperature on growth of scenedesmus obliquus
in diluted olive mill wastewater as culture medium. Engineering in Life Sciences, 10(3), 257–264, https://doi.
org/10.1002/elsc.201000005
Goldstein R. and Smith W. (2002). Water & Sustainability (Volume 4): U.S. Electricity Consumption for Water
Supply & Treatment – the Next Half Century. Electric Power Research Institute, Palo Alto, California,
USA.
Griffiths-Sattenspiel B. and Wilson W. (2009). The Carbon Footprint of Water. A River Network Report. River
Network, Portland, OR, USA.
Iasimone F., De Felice V., Panico A. and Pirozzi F. (2017). Experimental study for the reduction of CO2 emissions
in wastewater treatment plant using microalgal cultivation. Journal of CO2 Utilization, 22, 1–8, https://doi.
org/10.1016/j.jcou.2017.09.004
Ji F., Yin H., Zhang H., Zhang Y. and Lai B. (2018). Treatment of military primary explosives wastewater
containing lead styphnate (LS) and lead azide (LA) by mFe0-PS-O3 process. Journal of Cleaner Production,
188, 860–870, https://doi.org/10.1016/j.jclepro.2018.04.029
Johnston A. H. and Karanfil T. (2013). Calculating the greenhouse gas emissions of water utilities. Journal AWWA,
105(7), E363–E371, https://doi.org/10.5942/jawwa.2013.105.0073
Karigar C. S. and Rao S. S. (2011). Role of microbial enzymes in the bioremediation of pollutants: A review.
Enzyme Research, 2011, 11, https://doi.org/10.4061/2011/805187
Khan M. I., Shin J. H. and Kim J. D. (2018). The promising future of microalgae: current status, challenges, and
optimization of a sustainable and renewable industry for biofuels, feed, and other products. Microbial Cell
Factories, 17(1), 36, https://doi.org/10.1186/s12934-018-0879-x
Lal R. (2004). Soil carbon sequestration to mitigate climate change. Geoderma, 123(1), 1–22, https://doi.
org/10.1016/j.geoderma.2004.01.032
Lal R. (2015). Cover cropping and the 4 per thousand proposal. Journal of Soil and Water Conservation, 70,
141A–141A, https://doi.org/10.2489/jswc.70.6.141A
Lemar P. and de Fontaine A. (2017). Energy Data Management Manual for the Wastewater Treatment Sector.
Better Buildings. Office of Energy Efficiency and Renewable Energy, United States Department of Energy,
Washington, DC, USA.
Loge F. (2016). Is Using Less Water the Secret to Cutting our Greenhouse gas Emissions? The Guardian. Guardian
Media Group, London, UK.
Mahmood Q., Pervez A., Zeb B. S., Zaffar H., Yaqoob H., Waseem M., Zahidullah and Afsheen S. (2013).
Natural treatment systems as sustainable ecotechnologies for the developing countries. BioMed Research
International, 19, 796373.
Mander Ü., Dotro G., Ebie Y., Towprayoon S., Chiemchaisri C., Nogueira S. F., Jamsranjav B., Kasak K., Truu
J., Tournebize J. and Mitsch W. J. (2014). Greenhouse gas emission in constructed wetlands for wastewater
treatment: a review. Ecological Engineering, 66, 19–35, https://doi.org/10.1016/j.ecoleng.2013.12.006
Matassa S., Verstraete W., Pikaar I. and Boon N. (2016). Autotrophic nitrogen assimilation and carbon capture
for microbial protein production by a novel enrichment of hydrogen-oxidizing bacteria. Water Research, 101,
137–146, https://doi.org/10.1016/j.watres.2016.05.077
McCarty P. L., Bae J. and Kim J. (2011). Domestic wastewater treatment as a net energy producer–can this be
achieved? Environmental Science and Technology, 45(17), 7100–7106, https://doi.org/10.1021/es2014264
Metcalf & Eddy. (2003). Physical Unit Processes. In: Wastewater Engineering: Treatment and Reuse, 4th edn, G.
Tchobanoglous, F. L. Burton and H. D. Stensel (eds), McGraw-Hill, Boston, pp. 393–394.
Minx J. C., Lamb W. F., Callaghan M. W., Fuss S., Hilaire J., Creutzig F., Amann T., Beringer T., de Oliveira Garcia
W., Hartmann J., Khanna T., Lenzi D., Luderer G., Nemet G. F., Rogelj J., Smith P., Vicente J. L., Wilcox J.
and del Mar Zamora Dominguez M. (2018). Negative emissions – part 1: research landscape and synthesis.
Environmental Research Letters, 13(6), 063001, https://doi.org/10.1088/1748-9326/aabf9b
Molitor H. R. and Schnoor J. L. (2020). Using simulated flue gas to rapidly grow nutritious microalgae with
enhanced settleability. ACS Omega, 5(42), 27269–27277, https://doi.org/10.1021/acsomega.0c03492
Myhre G., Shindell D., Bréon F. M., Collins W., Fuglestvedt J., Huang J., Koch D., Lamarque J. F., Lee D., Mendoza
B., Nakajima T., Robock A., Stephens G., Takemura T. and Zhang H. (eds) (2013). Anthropogenic and
Natural Radiative Forcing. Cambridge University Press, Cambridge, UK.
Nahlik A. M. and Fennessy M. S. (2016). Carbon storage in US wetlands. Nature Communications, 7(1), 13835,
https://doi.org/10.1038/ncomms13835
Ney R. A., Schnoor J. L. and Mancuso M. A. (2002). A methodology to estimate carbon storage and flux in
forestland using existing forest and soils databases. Environmental Monitoring and Assessment, 78(3), 291–
307, https://doi.org/10.1023/A:1019939003210
Ney R. A., Meyers T. A., Espina A. and Schnoor J. L. (2005). How many samples are required? Evaluating a
model for verification of carbon sequestration in a hybrid poplar buffer strip. Environmental Monitoring and
Assessment, 102(1–3), 375–388, https://doi.org/10.1007/s10661-005-6393-8
Paterson K. G. and Schnoor J. L. (1992). Fate of alachlor and atrazine in a riparian zone field site. Water
Environment Research, 64(3), 274–283, https://doi.org/10.2175/WER.64.3.13
Paterson K. G. and Schnoor J. L. (1993). Vegetative alteration of nitrate fate in unsaturated zone. Journal of
Environmental Engineering, 119(5), 986–993, https://doi.org/10.1061/(ASCE)0733-9372(1993)119:5(986)
Pavlineri N., Skoulikidis N. T. and Tsihrintzis V. A. (2017). Constructed floating wetlands: a review of research,
design, operation and management aspects, and data meta-analysis. Chemical Engineering Journal, 308,
1120–1132, https://doi.org/10.1016/j.cej.2016.09.140
Pinar G., Duque E., Haidour A., Oliva J., Sanchez-Barbero L., Calvo V. and Ramos J. L. (1997). Removal of high
concentrations of nitrate from industrial wastewaters by bacteria. Applied and Environmental Microbiology,
63(5), 2071–2073, https://doi.org/10.1128/aem.63.5.2071-2073.1997
Rebolloso-Fuentes M. M., Navarro-Pérez A., García-Camacho F., Ramos-Miras J. J. and Guil-Guerrero J. L. (2001).
Biomass nutrient profiles of the microalga nannochloropsis. Journal of Agricultural and Food Chemistry,
49(6), 2966–2972, https://doi.org/10.1021/jf0010376
Rosli F. A., Lee K. E., Choo Ta G., Mokhtar M., Latif M. T., Goh T. and Simon N. (2017). The use of constructed
wetlands in sequestrating carbon: an overview. Nature Environment and Pollution Technology, 16, 813–819.
Ryan C. C., Tan D. T. and Arnold W. A. (2011). Direct and indirect photolysis of sulfamethoxazole and trimethoprim
in wastewater treatment plant effluent. Water Research, 45, 1280–1286, https://doi.org/10.1016/j.watres.2010.
10.005
Saeid A. and Chojnacka K. (2019). Organic Farming. Elsevier B.V., The Netherlands.
Silkina A., Ginnever N. E., Fernandes F. and Fuentes-Grünewald C. (2019). Large-scale waste bio-remediation
using microalgae cultivation as a platform. Energies, 12(14), 2772–2788, https://doi.org/10.3390/en12142772
Steinmann C. R., Weinhart S. and Melzer A. (2003). A combined system of lagoon and constructed wetland for an
effective wastewater treatment. Water Research, 37, 2035–2042, https://doi.org/10.1016/S0043-1354(02)00441-4
Tang J., Feng H. and Shen G.-Q. (2003). Drum dryers. In: Encyclopedia of Agricultural, Food, and Biological
Engineering, D. R. Heldman and C. I. Moraru (eds), Marcel Dekker Inc., New York, NY, USA, pp. 211–214.
Tellatin S. and Myers R. (2018). Cover Crops and Carbon Sequestration. United States Department of Agriculture,
National Institute for Food and Agriculture, Sustainable Agriculture Research & Education, College Park,
MD, USA.
Torzillo G. and Chini Zittelli G. (2015). ‘Tubular photobioreactors.’ Volume 2: products and refinery design. In:
Algal Biorefineries, A. Prokop, R. K. Bajpai and M. E. Zappi (eds), Springer International Publishing AG
Switzerland, Cham, Switzerland, pp. 187–212.
Tredici M. R., Rodolfi L., Biondi N., Bassi N. and Sampietro G. (2016). Techno-economic analysis of microalgal
biomass production in a 1-ha Green wall panel (GWP ®) plant. Algal Research, 19, 253–263, https://doi.
org/10.1016/j.algal.2016.09.005
Trommsdorff C. (2015). Can the Water Sector Deliver on Carbon Reduction? The Source. International Water
Association, London, UK.
Van Den Hende S., Vervaeren H., Desmet S. and Boon N. (2011). Bioflocculation of microalgae and bacteria
combined with flue gas to improve sewage treatment. New Biotechnology, 29(1), 23–31, https://doi.org/10.
1016/j.nbt.2011.04.009
Webb D. T., Nagorzanski M. R., Powers M. M., Cwiertny D. M., Hladik M. L. and LeFevre G. H. (2020). Differences
in neonicotinoid and metabolite sorption to activated carbon are driven by alterations to the insecticidal
pharmacophore. Environmental Science and Technology, 54(22), 14694–14705, https://doi.org/10.1021/acs.
est.0c04187
Wenk J., Nguyen M. T. and Nelson K. L. (2019). Natural photosensitizers in constructed unit process wetlands:
photochemical characterization and inactivation of pathogen indicator organisms. Environmental Science
and Technology, 53(13), 7724–7735, https://doi.org/10.1021/acs.est.9b01180
Wong S. C. and Law-Flood A. (2011). Tapping the Energy Potential of Municipal Wastewater Treatment: Anaerobic
Digestion and Combined Heat and Power in Massachusetts. Massachusetts Department of Environmental
Protection, Boston, MA, USA.
Zhu Y., Jones S. B. and Anderson D. B. (2018). Algae Farm Cost Model: Considerations for Photobioreactors.
Pacific Northwest National Laboratory, United States.
Chapter 13
Microbial electrochemical
communication in carbon and
electron flow for CO2 methanation
Aijie Wang1*, Bo Wang2, Zechong Guo3, Weiwei Cai4 and Wenzong Liu1
1State Key Lab of Urban Water Resource and Environment, School of Civil and Environmental Engineering, Harbin Institute of
Technology Shenzhen, Shenzhen 518055, China
2Center for Electromicrobiology, Section of Microbiology, Department of Biology, Aarhus University, Aarhus C 8000, Denmark
3School of Environmental and Chemical Engineering, Jiangsu University of Science and Technology, Zhenjiang 212100, China
*Correspondence: [email protected]
13.1 INTRODUCTION
Currently, sustainable technologies have been widely explored and developed to capture/fix or utilize
CO2 via chemical, electrochemical, photochemical, and biotechnological approaches accompanied
by less or green energy input (Centi & Perathoner, 2009; Kondratenko et al., 2013). Anaerobic
digestion (AD) involves bacterial fermentation of organic wastes in the absence of free oxygen, which
consequently supports methanogens to produce CH4 as fuel (Yu et al., 2021). Due to various biological
consortia and different by-products involved in complex substrates biodegradation, CH4 production is
determined by two main pathways of acetoclastic and hydrogenotrophic methanogensis. Importantly,
hydrogenotrophic methanogens can convert CO2 to CH4 by utilizing H 2 produced by acetogenesis.
However, H 2 is usually not accumulated much in the system because of the sensitive feedback of H 2
to acetogens.
Recently, electrochemical CO2 reduction is developed into an effective method for the activation
and transformation of stable CO2 molecules, which is driven by electron flow from anode to cathode
(Jhong et al., 2013). Importantly, the cathode surface chemistry of CO2 reduction always requires
efficient and special catalysts. The current technique of microbial electrocatalysis CO2 reduction
offers promising prospects for reducing carbon levels in a sustainable manner, taking full advantage
of CO2-derived chemical commodities (Bajracharya et al., 2017). Bioelectrocatalytic reduction of CO2
employs versatile microbes to achieve carbon reduction. On the anode side, biological electrons can
be collected by exoelectrogens from waste organics in wastewater, waste biomass, and so on., and
on the cathode side, the cathode can supply H 2 and electrons for communities, which can directly
utilize CO2 as a final electron acceptor for their metabolism, to produce high-value-added chemicals,
containing one or more carbons, like CH4, acetate, and so on. Therefore, carbon cycling driven by
various functional bacteria has been investigated on the acceleration and regulation process via
electron transfer.
© 2022 The Editors. This is an Open Access eBook distributed under the terms of the Creative Commons Attribution Licence
(CC BY-NC-ND 4.0), which permits copying and redistribution for non-commercial purposes with no derivatives, provided the
original work is properly cited (https://creativecommons.org/licenses/by-nc-nd/4.0/). This does not affect the rights licensed or
assigned from any third party in this book. The chapter is from the book Pathways to Water Sector Decarbonization, Carbon Capture
and Utilization, Zhiyong Jason Ren and Krishna Pagilla (Eds.)
The carbon cycle, as one of the crucial natural processes on the earth, includes a large number
of organisms, in which diverse species of microorganisms play joint efforts in organic degradation,
energy utilization, and biosynthesis. Different types of interspecies interactions work and connect
for mutualism, in which two or more distinct species living in close proximity rely on each other
for nutrients, protection, and/or other life functions (Kouzuma et al., 2015). Efficient interspecies
communication for electromethanogenesis (EM) has been disclosed through direct interspecies
electron transfer (DIET) in the form of electric currents in syntrophic consortia (Kouzuma et al.,
2015; Lovley, 2017b). Typically, a representative example of such syntrophy is found in an integrated
system of electrochemically enhanced AD for accelerating CH4 production, where the pathway of
CO2 reduction by reducing equivalents, for example H 2 and electrons, can be dominant between
syntrophic partners of exoelectrogens and hydrogenotrophic methanogens (Rotaru et al., 2014b).
Different types of electrochemically enhanced AD systems have been developed under external
applied voltages (Huang et al., 2022), indicating improved properties on microbial system stability
(Liu et al., 2016c; Cai et al., 2019), gas production (Cai et al., 2016b; Liu et al., 2016b), and low-
temperature adaption (Liu et al., 2016a). The direct biological conversion of electrical current and
CO2 into CH4 by electromethanogenesis has been observed (Cheng et al., 2009).
It has been ascertained that external energy can power microbes and microbial productions
(Lovley, 2011; Lovley & Nevin, 2013). EM requires a relatively low energy input (0.2–0.8 V) to provoke
CO2 reduction and CH4 recovery from wastewaters or wastes and simultaneously enhance organics
degradation when couples of an AD system (Cai et al., 2016b). Nonetheless, the energy consumption
is always evaluated through electricity input, moreover, the electricity delivery is restricted over
the long-term operation in rural or remote areas (Wang et al., 2020c). Therefore, renewable energy
powering electromethanogenesis is crucial to widespread application. Solar energy is regarded as one
of the most feasible options to conform to the energy demand worldwide because of its availability and
enormous solar energy (120,000 TW) striking the earth daily (Kamat, 2007). However, natural solar
energy is a day-night intermittent power, which is supposed to be a limitation aspect to overcome.
The utilization of electricity from renewable sources for CO2 reduction can also alleviate the existing
challenges associated with the intermittent output of renewable energy by storing the electricity in
chemical forms.
EM takes place in a niche with a plentiful supply of H 2 or electrons, like H 2-producing fermenters,
cathodic surface, or conductive carriers, and CO2 can be reduced directly or indirectly to CH4 by
methanogens through various pathways. Firstly, hydrogenotrophic methanogens can be enhanced
through mediated electron transfer with molecular H 2 as an electron carrier. Besides fermentative H 2,
which usually has very low pressure in an AD system, more H 2 can be further supplied on cathode
via electrochemical reactions (H 2 evolution under cathodic catalysts) or bioelectrochemical processes
(produced by exoelectrogens which are capable of accepting electrons from solid cathode) (Equation
(13.1)) and consumed by hydrogenotrophic methanogens to produce CH4 (Equation (13.2)) (Villano
et al., 2010). H 2 evolution theoretically requires a cathode potential of −0.41 V vs. SHE, but the more
negative potential is practically needed due to various electrochemical losses (Wang et al., 2020b):
2H+ + 2e− → H2 (13.1)
Secondly, EM can also directly be conducted without an electron carrier. Some hydrogenotrophic
methanogens (e.g. Methanobacterium species) can accept electrons directly from the cathode and
reduce CO2 to CH4 with the participation of protons produced by the anode (Equation (13.3)) (Lohner
et al., 2014). This process requires a cathode potential of −0.24 V vs. SHE:
CO2 + 8H+ + 8e− → CH4 + 2H2O (13.3)
Thirdly, except for H 2 , some other organic compounds (such as acetate, ethanol, butyrate,
succinate etc.) derived from electric current and CO2 via cathodic electrosynthesis also provide
more available substrates for methanogens (Steinbusch et al., 2010). Acetate synthesized on the
cathode with acetogenic microorganism (e.g., Sporomusa ovata) as catalysts (Equation (13.4))
(Nevin et al., 2011) can be subsequently converted into CH4 by acetotrophic methanogens (Equation
(13.5)) (Jiang et al., 2013), while other compounds require syntrophic metabolism via interspecies
electron transfer:
H2 + CO2 → Acetate (13.4)
Compared with anaerobic methanogenesis, EM has its own particularities in carbon conversion
and electron flow routes. Carbon conversion routes in electromethanogenesis are more diverse as a
result of the involvement of electrochemical processes, which is generally beneficial for the stability of
the system. In addition, organic substrate degradation and CH4 production are separated in different
regions, making it easier to create a favorable microenvironment for sensitive methanogens.
In AD, the electron transfer between different microbial populations is usually through intermediate
products (electron carriers), which is implicit and spontaneous. In the EM, the external voltage is
like a pump, extracting electrons from the source (organic substrates) and delivering them into the
product (CH4). Due to the existence of external voltage and circuit, a portion of the electron flow
becomes explicit and mandatory, which substantially improves the monitorability and controllability
of the system. In light of the above reasons, EM has great potentials to overcome the unfavorable
factors that plague the conventional methanogenic pathway in anaerobic bioreactors and improve
CH4 production efficiency.
common exoelectrogenic bacteria (e.g., Geobacter, Shewanella, and Desulfovibrio) are found related
to c-type cytochrome genes in these systems, moreover, a relatively high functional and phylogenetic
diversity of microorganisms can be developed despite the feedback of a single substrate, like acetate-
alone. For example, functional genes related to substrate degradation accounted for 15–25% of
carbon degradation and fixation in all gene categories, while functional genes related to complex
carbon utilization accounted for ca. 10% in all detected genes (Liu et al., 2010). A variety of carbon
degradation genes, including amylase, xylanase, and endochitinase, were varied considerably among
reactor operations. Accordingly, bioreactors with high coulombic efficiency and energy harvest
(H2 yield) have the greatest capability for using a variety of complex carbon sources. A significant
correlation is reported between coulombic efficiency and community composition (r = 0.84, P = 0.025),
and COD removal and carbon degradation (r = 0.84, P = 0.035) with community structure (Liu et al.,
2010). However, carbon degradation communities on anode are not significantly related to the cathode
terminal products.
When fermentative substrates (glucose) are used for reactors, functional genes with high diversities
involve in complex carbon degradation, including carbon degradation genes of cellulose, hemicellulose,
lignin, starch, pectin, and chitin (Varrone et al., 2014). More microorganisms in anode biofilm are
detected from cytochrome genes, such as cytochromes derived from Geobacter metallireducens for
metal reduction, Bradyrhizobium sp. involved in the oxidation of organic contaminants, OmcA/MtrC
from Geobacter sulfurreducens, Shewanella sediminsis, S. oneidensis, S. amazonensis, S. loihica, and S.
pealeana. It indicates that the reactors with the highest energy recovery showed a higher (total) amount
of cytochrome genes. Meanwhile, microorganisms related to carbon fixation genes for rubisco, carbon
monoxide dehydrogenase, and propionyl-CoA carboxylase are detected. For example, propionyl-CoA
carboxylase genes derived from Roseiflexus sp., Nitrobacter hamburgensis, Chloroflexus aggregans.
Also, a part of detected bacteria are uncultured, like carbon monoxide dehydrogenase (CODH) from
an uncultured bacterium (lab clone), and rubisco derived from an uncultured bacterium.
A comprehensive community structure is supposed to be well developed from various substrates,
though higher functional diversity can be detected when utilizing complex carbon than simple sole
carbons from acetate to glucose. Genes from all major functional categories indicate that microbial
communities are able to perform a large variety of functions. Necessarily, the variability of community
functions is not (only) related to the presence of exoelectrogens, as a fraction of carbon degradation
functions can be played by non-exoelectrogens of fermentation and or symbiotic relationships with
other bacteria. Accordingly, the extracellular electron transfer pathway has an important impact
on the methanogenic community structure in the reactors. Applied voltages increased the relative
abundance of hydrogenotrophic methanogens over acetoclastic methanogens. In control reactors with
open circuit, a relatively high abundance of acetoclastic methanogens are presented. However, six
genera were detected with the ability of H 2 utilization and CH4 production in closed circuit, including
Methanobacterium, Methanococcus, Methanoculleus, Methanocorpusculum, Methanospirillum, and
Candidatus Methanoregula, but only one genus of Methanosarcina was defined as acetate utilization
methanogens.
processes when the system is unstable. The microbial electrocatalytic process, in contrast, is able to
exert a directional intervention to organic degradation pathways through external electric energy
input, which can be an effective complementary approach for AD. Therefore, an electrochemically
enhanced AD is proposed in the expectation of integrating the advantages of these two technologies
and providing an ideal choice for carbon fixation.
Since the concept of the electrochemically enhanced AD is proposed by Willy Verstraete in 2006
(Pham et al., 2006), many research works have been conducted to verify the feasibility of coupling
technology. It was reported that CH4 production performance and the stability of AD systems can be
promoted by introducing a microbial electrocatalytic process (Malaeb et al., 2013; Wang et al., 2022b).
The potential mechanism for these promoting effects has been analyzed in recent studies, and it is
believed that microbial community distribution and carbon conversion pathways in the anaerobic
bioreactor are positively regulated by bioelectrochemical processes (Guo et al., 2017a). Applying
voltage in a complicated anaerobic environment is beneficial for the creation of suitable habitat for
anaerobic and facultative populations by reducing electrode potential and boosting hydrogenotrophic
methanogens by providing additional H 2 (Wang et al., 2009). The influence of electrochemistry includes
both instant direct contribution and long-term indirect effect (such as affecting the distribution of
anaerobic community during the start-up stage) (Liu et al., 2016c; Zhao et al., 2015). It is disclosed
by sequencing techniques in some researches that external voltage, as a positive growth condition
or selective pressure, has substantial regulatory effects on microbial community structures. In
particular, external voltage leads to an increase in electrogenic microorganisms and hydrogenotrophic
methanogens (Liu et al., 2016b).
Besides microorganism community distribution, the external voltage can also influence the
organic degradation pathway positively in an anaerobic system. According to an electron-balance
analysis of the glucose digestion process in our case (Figure 13.1) (Guo et al., 2017a), more electrons
Figure 13.1 Electron balance analysis for glucose digestion in absence and presence of electrochemical process
(Guo et al., 2017a).
obtained from the oxidation of glucose were transferred and finally stored in propionate and butyrate
(34.9%) in an AD system without electrochemical regulation. Hydrogenotrophic methanogenesis
was unrestricted, but acetotrophic methanogenesis was inadequate due to relatively low amounts of
acetotrophic methanogens, and only a small amount of acetate was converted into CH4 (3.6%). The
waste of electron flow in fermentation products (propionate, butyrate, and even acetate) resulted
in unsatisfactory CH4 -producing performance. With the regulation of bioelectrochemistry, however,
more electrons were transferred from glucose to H 2 (23.4%) and acetate (48.6%), instead of butyrate
and propionate, and the bioelectrolysis reaction created an additional pathway between acetate and
H 2, about 10.6% of the total electrons were transferred from acetate into H2 through circuit current,
which alleviated the limitation of acetotrophic methanogens and led to a significant enhancement of
CH4 production.
The influence of electrochemistry on the anaerobic degradation process is largely determined by
the electrochemical efficiency of electrodes (e.g., current density, cathodic catalysis efficiency, etc.)
and the complexity of the system. Generally, the ratio of bioelectrochemical reactions in the total
electron flow and the direct contribution of electrochemistry on biogas production will be promoted by
increasing the electrochemical efficiency of electrodes (current density) through various methods (such
as increasing electrode areas, using efficient electrode materials, optimizing reactor configuration,
etc.) (Wang et al., 2017). Supplying simple substrates (such as acetate and ethanol) is beneficial for
establishing an efficient anode biofilm and obtaining good electrochemical performances. With
complex compounds (such as wastewater or waste sludge) as carbon substrates, however, the biofilm
communities and anaerobic digestion processes became more complex (Zhang et al., 2011). The
relative abundance of electrode-respiring bacteria and the electron transfer efficiency were generally
lower than those in the systems using acetate, which may weaken the effects of electrochemistry on
the AD process (Zhang et al., 2011).
Methanogenesis is generally the rate-limiting step of AD with soluble substrates, and methanogens
are the most important but also the most vulnerable members in the CH4 -producing functional
microbial community. In fact, many bottlenecks in AD technology (such as long lag time, poor stability)
can be ultimately attributed to the high sensitivity and slow growth of methanogens. Considering the
positive effect of electrochemistry on the growth and metabolism of methanogens, the introduction of
electrochemical components may be a feasible solution to the existing problems of the AD system and
may play a greater role in constructing carbon-neutral treatment process of organic waste streams.
However, the metabolism type of methanogens was rarely monotonous in AD due to the substrate
limitation for methanogens. Commonly, the H 2/CO2 and acetate will be utilized as the substrate for
CH4 production, which are termed hydrogenotroph and acetotroph, respectively. Although there are
methylotrophs in natural ecosystems for methanogenesis, the lower abundances of methylotrophs
result in less attention in AD reactors.
Among the nutrient type of methanogens in AD of WAS, the hydrogenotrophic methanogens,
utilizing H 2 as an electron donor and CO2 as an electron acceptor, are an important adjuster of H 2.
As the fermentative species need to reduce NADH into NAD+, and the regeneration of NAD+ ensures
the completion of glycolysis, H 2 will be sensitive as its accumulation will thermodynamically affect
the NAD+ regeneration (Stams & Plugge, 2009). The H 2 formation in the fermentation stage also
provides an insight into biological H 2 production in dark fermentation. However, recent progress
in bioelectrochemistry which combined dark fermentation and microbial electrolysis led to a
breakthrough in H 2 production from glucose (Varanasi et al., 2019; Wang et al., 2020b). Naturally,
the hydrogenotrophic methanogens will capture H 2 as a precursor for CH4 production, therefore, the
CH4 is considered as an inevitable product in a hybrid system (Wang et al., 2009).
Recently, a hybrid model for electrochemically enhanced AD was proposed to enable an
accelerated CH4 production rate from waste activated sludge (WAS) (Liu et al., 2016b). Electrode H 2
evolution provides a unique niche for hydrogenotrophic methanogens as a continuous H2 supplier.
A novel electrochemically enhanced AD reactor was carried out to test its performances in CH4
production which was fed with WAS. The CH4 production rate was enhanced by the introduction of
electrodes, approaching approximately 3.2 times the control AD. Based on the electron balance, the
sum of control CH4 production and current contribution is highly aligned with the enhanced CH4
production. The current value can reach as high as 12 mA (the maximum), which is higher than that
of lab-scale bioelectrochemical reactors. A further dynamic model of methanogenesis was built in
electrochemically enhanced AD, indicating the CH4 production rate would be improved by 1.4 times
feeding with glucose (Guo et al., 2017a). The configuration of electrochemically enhanced AD will
affect the performance in the CH4 production rate. The higher ratio of cathode/anode enhanced the
CH4 production rate (increased by 56–180%) in our previous study (Guo et al., 2017b). Furthermore,
the independent cathode also caused an enhancement in CH4 production in a continuous model
with glucose as substrate for cathode and acetate for anode (Cai et al., 2016a). The electrochemically
enhanced AD provided a promising way to accelerate the CH4 production from WAS, which is valuable
for AD as the long sludge retention time is the bottleneck in its implementation. The improvement
of the CH4 production rate potentially enables the AD to shorten the sludge retention time with the
similar performance of CH4 yield.
proving the Methanosaeta was capable of accepting electrons directly from Geobacter. This finding
induced considerable attention on the DIET (Rotaru et al., 2014a). A mutant of Methanococcus
without hydrogenase was subsequently cultured as working species, confirming the direct electron
transfer without H 2 dependence (Lohner et al., 2014). Moreover, Methanosarcinales was verified to
perform multiple modes (hydrogenase-mediated and free extracellular enzyme-independent modes)
of electrode interactions on cathodes (Rowe et al., 2019), which expands the underlying electron
transfer mechanisms at the cathode. Recently, a defined coculture of Methanobacterium strain YSL
and Geobacter metallireducens can grow via DIET (Zheng et al., 2020). The main classifications of
methanogens were putative to have the ability to achieve extracellular electrons to produce CH4 (Gao
& Lu, 2021). Therefore, the DIET in methanogens is more broadly distributed than we expected.
Although the pure culture was found to be able to directly acquire electrons from the cathode, the
electron transfer in microbial electrolysis assisted anaerobic digestion normally occurred within the
biofilm which consists of diverse species. There are two questions that should be noticed: firstly, why the
Methanobacterium or Methanobrevibacter always dominated in the biofilm rather than Methanosaeta
or Methanosarcina that both can utilize H 2/electrons from the cathode; secondly, whether the DIET
will support the holistic biofilm performances. A further study revealed the microenvironment of
the cathode is differing from the bulk solution as the rapid consumption of proton (Cai et al., 2020).
The mcrA sequencing technology provided a higher resolution into the classification of methanogens
at the species level, the basophilic methanogens was confirmed as the enriched Methanobacterium
genus at the cathode with the extreme alkaline microenvironment (at the µm-scale) (Cai et al., 2018a).
Therefore, the extreme condition may induce the enrichment of special methanogens, such as the
Methanobacterium genus. In the mixed culture, H 2 was firstly confirmed as the electron carriers in
the syntrophy between fermentative species and methanogens, then the formate was found to be an
alternative to H 2 with extremely high efficiency (Stams & Plugge, 2009). The novel DIET was proved
as one kind of electron carrier, a recent study suggested it can be an option for diverse syntrophs as
the presence of e-pili in Syntrophus resembled the type IV pili of Geobacter (Walker et al., 2020). A
modeling confirmed that the formate can reach 317 × 103 e− cp−1 s−1 for electron transfer in syntrophy,
whose the DIET ability will be 44.9 × 103 e − cp−1 s −1 that is higher than that of H 2 -mediate (5.24 ×
103 e− cp−1 s−1) (Storck et al., 2016). Clearly, the formate will be the ideal mediates for electron transfer,
which normally is not the main pathway for syntrophy as the formate is thermodynamically reversible
to form H 2 (Cai et al., 2020). Besides, the growth of methanogens without cytochrome was restricted
on formate, as the H 2 is still an intermediate inside of methanogens through coenzyme F420 -dependent
formate dehydrogenase, the formate-dependent will lose an advantage to compete with others who
own lower H 2 thresholds in the natural environments (Thauer et al., 2008). In typical AD reactors,
H 2 is still the main intermediate for electron transfer within syntrophic bacteria and methanogens.
H 2 was detected in our subsequent test in cathodic biofilm (Cai et al., 2020). A clear gradient of H 2
concentration implied it will be the electron carriers at the cathode for methanogenesis. Meanwhile,
the cyclic voltammetry (CV) test was used to couple with the microsensor, which provided semi-
quantitative results of H 2 contribution in total electron transfer. It was surprising to discover only less
than 50% of electrons would be transported via H 2, thus, the DIET may contribute significantly to
cathodic methanogenesis. However, the extracellular electron transfer pathways have been extensively
expanded according to recent progress, not only the pili, flavin, cytochrome, but also exDNA, vesicle,
extracellular polymer substances (EPS) have been involved (Liu et al., 2020; Lovley, 2017a; Saunders
et al., 2020; Xiao et al., 2017). The H 2-free electron pathway also may include multi-mediates rather
than only direct electron transfer, which still needs further study.
Figure 13.3 Enhanced hydrogenotrophic methanogenesis via the extracellular electron transfer pathway.
Figure 13.4 Microbial network of acidogenic bacteria, electro-respiring bacteria and methanogens.
communities are highly enriched in dominant functional groups related to Proteobacteria (∼60%)
and Firmicutes (20–30%), which conduct substrate degradation and electron transfer with significant
potential on complicated carbon utilization, as core communities in the context of extracellular
electron transfer. They are supposed to play a very important role in carbon recycling as it has been
reported that Firmicutes may have a symbiotic relationship with anode respiring bacteria (Zhang
et al., 2011). Representative anaerobic fermentative bacteria account for 1–10% overall communities,
consisting of Citrobacter (class Gammaproteobacteria), Macellibacteroides (class Bacteroidia)
and two genera of class Clostridia (Proteiniclasticum and Sedimentibacter) were enriched after
prefermentation and became the predominant bacteria. Proteiniclasticum, responsible for the
degradation of proteins to produce HAc, propionic acid (HPr) and iso-butyric acid (iso-HBu) (Zhou
et al., 2018), had an abundance of 6%. Macellibacteroides with abundance of 5–7% can metabolize
various carbohydrates for HAc, HBu, and iso-HBu generation (Jabari et al., 2012). The genera
Citrobacter and Sedimentibacter can degrade organics to produce VFAs and H 2 during acidogenesis
and acetogenesis. Compared to the original anode biofilm communities, some bacteria also shift
during the connection to sludge fermentation process. Four genera of the class Clostridia (namely
Acetoanaerobium, Acetobacterium, Anaerovorax and Fusibacter) decreased in all integrated reactors
feeding sludge fermentation liquid.
Methanogens are powered and developed in all biofilms in integrated systems. Meanwhile,
the abundance of the fermentation bacteria in the sludge can be significantly improved, but the
acetotrophic methanogen shows a decrease (Wang et al., 2019). The acetotrophic methanogens can
be enriched in anode biofilm in all conditions with enough substrate carbons, though the substrate
competition exists between anode microorganisms and methanogens on acetate utilization, while
hydrogenotrophic methanogens will be further enriched with higher abundance in cathode biofilm with
continuous applied voltages (Cai et al., 2019), including Methanocorpusculum, Methanosphaerula,
Methanoregula, Methanospirillum, Methanobacterium and Methanobrevibacter, which closely
related to the H 2 evolution of cathodic reaction, supplying favor substrates for hydrogenotrophic
methanogens in anaerobic condition. Community structure and the methanogen production
pathway will be directed by the biocathode materials with high H 2 evolution reaction (HER) activity.
A high HER tends to select H 2 instead of electrons for methanogenesis. As a result, the dominating
microorganisms at the cathode with HER capacity can be selected similarly, usually including
hydrogenotrophic Methanobacterium and Methanospirillum (Ferry et al., 1974; Rotaru et al., 2014a),
and the Methanobacterium could not only directly convert H 2 into CH4 but also electrons (Cheng
et al., 2009).
Inoculuma Reactor Temperature Cathode Applied Inorganic Current CH4 Current Reference
Configuration (°C) Material Voltage (V) Carbon Density Production to CH4
and Source/ (A·m−2) Rate (mol Efficiency
Operation conc. CH4·m−3 (%)
Mode (mol·L−1) reactor·d−1)
Anaerobic sludge from DC 35 Carbon paper Cathode CO2/ Cathode 0.173 ± 0.026 76 ± 7 Villano et al.
a packed bed biofilm Batch potential Quantity current (2010)
reactor fed with a −0.75 V vs. sufficient density –0.69
synthetic mixture SHE
of fatty acids and
alcohols
Activated sludge DC 30 Graphite felt Cathode HCO3−/0.06 3.8 0.268 51.3 Van Eerten-
Continuous potential Jansen et al.
−0.55 V vs. (2012)
NHEb
Enriched activated DC 30 ± 1 Carbon felt Cathode CO2/ 3.37 0.024 89.2 Jiang et al.
sludge Batch potential Quantity (2013)
IM1
An enriched DC 22 ± 2 Porous carbon Cathode CO2 Ca.15 0.03 ± 0.015 39 Dykstra and
hydrogenotrophic Batch felt potential (1.65 atm) (Coulombic Pavlostathis
methanogenic culture −0.8 V vs. efficiency) (2017)
SHE
by guest
Anaerobic sludge TC c 30 Heat-treated 3.5 ± 0.3 HCO3−/0.06 7.1 HSSF: 12.86 HSSF: 60.8 Sangeetha
Batch stainless steel felt SSF: 9.19 SSF: 56.9 et al. (2017)
(HSSF), stainless GF: 15.72 GF: 69.4
steel felt (SSF),
and graphite felt
(GF)
Anaerobic mixed DCBatch 30 Granular Cathode CO2/ 35 4.3 66 Liu et al.
sludge including activated carbon potential Quantity (2018)
granular sludge from −0.58 V sufficient
the paper industry
WWTP b and sludge
from the municipal
WWTP b
Anaerobic sludge DC 30 Graphite granules 2.8 ± 0.1 35 4.1 67
Continuous
Activated sludge from DC 30 ± 1 Carbon felt Cathode HCO3−/0.06 2.442 ± 0.484 0.24 ± 0.05 44.27 ± 4.01 Yang et al.
a full-size, enlarged Continuous potential (2020)
Neutral red- 7.622 ± 1.436 1.39 ± 0.17 58.90 ± 11.47
granular sludge-bed −1.0 V vs. Ag/
modified carbon
reactor for starch AgCl
felte
(a) (b)
Suspension
Methanobacterium
Acetoanaerobium
Petrimonas
Anaerocella
ShewanellaDechlorosoma
Methanosarcina
Arcobacter Acinetobacter
Azoarcus
Geobacter Fusibacter
Dechlorobacter
Acetobacteroides
Comamonas
Pseudomonas
Lentimicrobium
Acetobacterium
Methanobrevibacter
Acetoanaerobium
Acetobacterium Petrimonas
Anaerovorax Azoarcus
Acinetobacter
Dechlorobacter Desulfovibrio
Rhodopseudomonas Fusibacter
Acinetobacter Geobacter
Lentimicrobium Azoarcus
Methanobrevibacter Anaerocella Geobacter
Methanobacterium
Comamonas Dechlorosoma
Rhodopseudomonas
Desulfovibrio Pseudomonas
PseudomonasComamonas
Anaerovorax
Acetobacterium
Acetoanaerobium Methanobrevibacter
Methanosarcina
Figure 13.5 (a) schematic diagram of the single-chambered membrane-less bioelectrochemical system driven by
the intermittent electric field applied by manual on-off (group A) or natural solar power (group B) and continuous
electrical field (group C). (b) molecular ecological networks (MEN) visualization of OTUs from functional microbial
consortia in the anode biofilm, cathode biofilm, and bulk solution, respectively. Each node represents an OTU
(species), and different colors and shapes of nodes signifier categories of specific functional genera: electroactive
microorganisms (sky blue ellipse), methanogens (yellow round rectangle), acetogens (pink diamond) and anaerobic
fermentative bacteria (purple triangle). A red edge indicates a positive interaction between two individual nodes,
while a Green edge indicates a negative interaction (reproduced from Wang et al., 2020d).
in the bioelectrodes were presented in the acetogens. Fermentative bacteria (FB), with the ability
to degrade organic matters, showed tight affiliations with methanogens and acetogens attached
to the biocathode. More complex and diverse connectivity of EAMs, methanogens, acetogens
and FB appeared in the bulk solution and suggested obvious mutualistic symbiosis, cooperation,
and competition. Together, electrode biofilms had a more positive association and the planktonic
microbial community showed the opposite connectivity. Hence, either microbial assembly in the
anode or the bulk solution were closely linked to carbon source conversion, whereas the cathode
to biosynthesis CH4 was relatively independent, basically relying on electron transfer to bridge the
communication with the anode and suspension.
ACKNOWLEDGEMENTS
The authors would like to acknowledge the support from the Shenzhen Overseas High-level Talents
Research Startup Program from Harbin Institute of Technology, Shenzhen. This research was
financially supported by the Natural Science Foundation of Guangdong Province for Distinguished
Young Scientists (No. 2021B1515020084), the National Natural Science Foundation of China
(No. 51778607), the National Natural Science Foundation of China NSFC-ISF Joint Program (No.
41961144024), and the Shenzhen Science and Technology Innovation Program (No. KQTD2019092917
2630447).
REFERENCES
Ailijiang N., Chang J., Liang P., Li P., Wu Q., Zhang X. and Huang X. (2016). Electrical stimulation on
biodegradation of phenol and responses of microbial communities in conductive carriers supported
biofilms of the bioelectrochemical reactor. Bioresource Technology, 201, 1–7, https://doi.org/10.1016/j.
biortech.2015.11.026
Appels L., Baeyens J., Degrève J. and Dewil R. (2008). Principles and potential of the anaerobic digestion of waste-
activated sludge. Progress in Energy and Combustion Science, 34(6), 755–781, https://doi.org/10.1016/j.
pecs.2008.06.002
Bajracharya S., Srikanth S., Mohanakrishna G., Zacharia R., Strik D. P. B. T. B. and Pant D. (2017). Biotransformation
of carbon dioxide in bioelectrochemical systems: state of the art and future prospects. Journal of Power
Sources, 356, 256–273, https://doi.org/10.1016/j.jpowsour.2017.04.024
Cai W., Han T., Guo Z., Varrone C., Wang A. and Liu W. (2016a). Methane production enhancement by an
independent cathode in integrated anaerobic reactor with microbial electrolysis. Bioresource Technology,
208, 13–18, https://doi.org/10.1016/j.biortech.2016.02.028
Cai W. W., Liu W. Z., Yang C. X., Wang L., Liang B., Thangavel S., Guo Z. C. and Wang A. J. (2016b). Biocathodic
methanogenic community in an integrated anaerobic digestion and microbial electrolysis system for
enhancement of methane production from waste sludge. ACS Sustainable Chemistry and Engineering, 4(9),
4913–4921, https://doi.org/10.1021/acssuschemeng.6b01221
Cai W., Liu W., Zhang Z., Feng K., Ren G., Pu C., Sun H., Li J., Deng Y. and Wang A. (2018a). mcrA sequencing
reveals the role of basophilic methanogens in a cathodic methanogenic community. Water Research, 136,
192–199, https://doi.org/10.1016/j.watres.2018.02.062
Cai W. W., Liu W. Z., Sun H. S., Li J. Q., Yang L. M., Liu M. J., Zhao S. L. and Wang A. J. (2018b). Ni5P4-NiP2
nanosheet matrix enhances electron-transfer kinetics for hydrogen recovery in microbial electrolysis cells.
Applied Energy, 209, 56–64, https://doi.org/10.1016/j.apenergy.2017.10.082
Cai W. W., Liu W. Z., Zhang Z. J., Feng K., Ren G., Pu C. L., Li J. Q., Deng Y. and Wang A. J. (2019). Electro-driven
methanogenic microbial community diversity and variability in the electron abundant niche. Science of the
Total Environment, 661, 178–186, https://doi.org/10.1016/j.scitotenv.2019.01.131
Cai W., Liu W., Wang B., Yao H., Guadie A. and Wang A. (2020). Semiquantitative detection of hydrogen-associated
or hydrogen-free electron transfer within methanogenic biofilm of microbial electrosynthesis. Applied and
Environmental Microbiology, 86(17), e01056-20.
Cai W., Wang B., Liu W., Yao H., Deng Y. and Wang A. (2021). Sessile methanogens dominated cathodic biofilm:
distribution and network in physiological transitions. Science of the Total Environment, 795, 148724, https://
doi.org/10.1016/j.scitotenv.2021.148724
Centi G. and Perathoner S. (2009). Opportunities and prospects in the chemical recycling of carbon dioxide to
fuels. Catalysis Today, 148(3–4), 191–205, https://doi.org/10.1016/j.cattod.2009.07.075
Cheng S., Xing D., Call D. F. and Logan B. E. (2009). Direct biological conversion of electrical current into
methane by electromethanogenesis. Environmental Science and Technology, 43(10), 3953–3958, https://doi.
org/10.1021/es803531g
Cho S. K., Lee M. E., Lee W. and Ahn Y. (2019). Improved hydrogen recovery in microbial electrolysis cells
using intermittent energy input. International Journal of Hydrogen Energy, 44(4), 2253–2257, https://doi.
org/10.1016/j.ijhydene.2018.07.025
Ding A., Yang Y., Sun G. and Wu D. (2016). Impact of applied voltage on methane generation and microbial
activities in an anaerobic microbial electrolysis cell (MEC). Chemical Engineering Journal, 283, 260–265,
https://doi.org/10.1016/j.cej.2015.07.054
Dykstra C. M. and Pavlostathis S. G. (2017). Methanogenic biocathode microbial community development and
the role of bacteria. Environmental Science and Technology, 51(9), 5306–5316, https://doi.org/10.1021/acs.
est.6b04112
Ferry J. G., Smith P. H. and Wolfe R. (1974). Methanospirillum, a new genus of methanogenic bacteria, and
characterization of Methanospirillum hungatii sp. nov. International Journal of Systematic and Evolutionary
Microbiology, 24(4), 465–469.
Fu Q., Kuramochi Y., Fukushima N., Maeda H., Sato K. and Kobayashi H. (2015). Bioelectrochemical analyses
of the development of a thermophilic biocathode catalyzing electromethanogenesis. Environmental Science
and Technology, 49(2), 1225–1232, https://doi.org/10.1021/es5052233
Gao K. and Lu Y. (2021). Putative extracellular electron transfer in methanogenic Archaea. Frontiers in
Microbiology, 12, 611739, https://doi.org/10.3389/fmicb.2021.611739
Gao T., Zhang H., Xu X. and Teng J. (2021). Integrating microbial electrolysis cell based on electrochemical
carbon dioxide reduction into anaerobic osmosis membrane reactor for biogas upgrading. Water Research,
190, 116679, https://doi.org/10.1016/j.watres.2020.116679
Guo Z., Liu W., Yang C., Gao L., Thangavel S., Wang L., He Z., Cai W. and Wang A. (2017a). Computational and
experimental analysis of organic degradation positively regulated by bioelectrochemistry in an anaerobic
bioreactor system. Water Research, 125, 170–179, https://doi.org/10.1016/j.watres.2017.08.039
Guo Z., Thangavel S., Wang L., He Z., Cai W., Wang A. and Liu W. (2017b). Efficient methane production from
beer wastewater in a membraneless microbial electrolysis cell with a stacked cathode: the effect of the
cathode/anode ratio on bioenergy recovery. Energy and Fuels, 31(1), 615–620, https://doi.org/10.1021/acs.
energyfuels.6b02375
Huang Q., Liu Y. and Dhar B. R. (2022). A critical review of microbial electrolysis cells coupled with anaerobic
digester for enhanced biomethane recovery from high-strength feedstocks. Critical Reviews in Environmental
Science and Technology, 52(1), 50–89, https://doi.org/10.1080/10643389.2020.1813065
Hussain S. A., Perrier M. and Tartakovsky B. (2018). Long-term performance of a microbial electrolysis cell
operated with periodic disconnection of power supply. RSC Advances, 8(30), 16842–16849, https://doi.
org/10.1039/C8RA01863D
Jabari L., Gannoun H., Cayol J. L., Hedi A., Sakamoto M., Falsen E., Ohkuma M., Hamdi M., Fauque G.,
Ollivier B. and Fardeau M. L. (2012). Macellibacteroides fermentans gen. nov., sp. nov., a member of
the family Porphyromonadaceae isolated from an upflow anaerobic filter treating abattoir wastewaters.
International Journal of Systematic and Evolutionary Microbiology, 62(Pt 10), 2522–2527, https://doi.
org/10.1099/ijs.0.032508-0
Jhong H. R., Ma S. C. and Kenis P. J. A. (2013). Electrochemical conversion of CO2 to useful chemicals: current
status, remaining challenges, and future opportunities. Current Opinion in Chemical Engineering, 2(2),
191–199, https://doi.org/10.1016/j.coche.2013.03.005
Jiang Y., Su M., Zhang Y., Zhan G., Tao Y. and Li D. (2013). Bioelectrochemical systems for simultaneously
production of methane and acetate from carbon dioxide at relatively high rate. International Journal of
Hydrogen Energy, 38(8), 3497–3502, https://doi.org/10.1016/j.ijhydene.2012.12.107
Kamat P. V. (2007). Meeting the clean energy demand: nanostructure architectures for solar energy conversion.
Journal of Physical Chemistry C, 111(7), 2834–2860, https://doi.org/10.1021/jp066952u
Kondratenko E. V., Mul G., Baltrusaitis J., Larrazabal G. O. and Perez-Ramirez J. (2013). Status and perspectives
of CO2 conversion into fuels and chemicals by catalytic, photocatalytic and electrocatalytic processes.
Energy and Environmental Science, 6(11), 3112–3135, https://doi.org/10.1039/c3ee41272e
Kouzuma A., Kato S. and Watanabe K. (2015). Microbial interspecies interactions: recent findings in syntrophic
consortia. Frontiers in Microbiology, 6, 477.
Liu W., Wang A., Cheng S., Logan B. E., Yu H., Deng Y., Nostrand J. D., Wu L., He Z. and Zhou J. (2010). Geochip-
based functional gene analysis of anodophilic communities in microbial electrolysis cells under different
operational modes. Environmental Science and Technology, 44(19), 7729–7735, https://doi.org/10.1021/
es100608a
Liu D. D., Zhang L., Chen S., Buisman C. and ter Heijne A. (2016a). Bioelectrochemical enhancement of methane
production in low temperature anaerobic digestion at 10 degrees C. Water Research, 99, 281–287, https://
doi.org/10.1016/j.watres.2016.04.020
Liu W., Cai W., Guo Z., Wang L., Yang C., Varrone C. and Wang A. (2016b). Microbial electrolysis contribution
to anaerobic digestion of waste activated sludge, leading to accelerated methane production. Renewable
Energy, 91, 334–339, https://doi.org/10.1016/j.renene.2016.01.082
Liu W., He Z., Yang C., Zhou A., Guo Z., Liang B., Varrone C. and Wang A.-J. (2016c). Microbial network for waste
activated sludge cascade utilization in an integrated system of microbial electrolysis and anaerobic fermentation.
Biotechnology for Biofuels, 9(1), 1–15, https://doi.org/10.1186/s13068-015-0423-8
Liu D., Roca-Puigros M., Geppert F., Caizán-Juanarena L., Na Ayudthaya S. P., Buisman C. and
Ter Heijne A. (2018). Granular carbon-based electrodes as cathodes in methane-producing
bioelectrochemical systems. Frontiers in Bioengineering and Biotechnology, 6, 78, https://doi.org/10.3389/
fbioe.2018.00078
Liu W., Wang L., Gao L. and Wang A.-J. (2019). Hydrogen and methane production in bioelectrochemical system:
biocathode structure and material upgrading, Chapter 5.10. In: Microbial Electrochemical Technology, S. V.
Mohan, S. Varjani and A. Pandey (eds), Elsevier, The Netherlands, pp. 921–953.
Liu X., Jing X., Ye Y., Zhan J., Ye J. and Zhou S. (2020). Bacterial vesicles mediate extracellular electron transfer.
Environmental Science and Technology Letters, 7(1), 27–34, https://doi.org/10.1021/acs.estlett.9b00707
Lohner S. T., Deutzmann J. S., Logan B. E., Leigh J. and Spormann A. M. (2014). Hydrogenase-independent uptake
and metabolism of electrons by the archaeon Methanococcus Maripaludis. ISME Journal, 8(8), 1673–1681,
https://doi.org/10.1038/ismej.2014.82
Lopez-Garcia P. and Moreira D. (2020). The syntrophy hypothesis for the origin of eukaryotes revisited. Nature
Microbiology, 5(5), 655–667, https://doi.org/10.1038/s41564-020-0710-4
Lovley D. R. (2011). Powering microbes with electricity: direct electron transfer from electrodes to microbes.
Environmental Microbiology Reports, 3(1), 27–35, https://doi.org/10.1111/j.1758-2229.2010.00211.x
Lovley D. R. (2017a). Happy together: microbial communities that hook up to swap electrons. ISME Journal, 11(2),
327–336, https://doi.org/10.1038/ismej.2016.136
Lovley D. R. (2017b). Syntrophy goes electric: direct interspecies electron transfer. Annual Review of Microbiology,
71, 643–664, https://doi.org/10.1146/annurev-micro-030117-020420
Lovley D. R. and Nevin K. P. (2013). Electrobiocommodities: powering microbial production of fuels and
commodity chemicals from carbon dioxide with electricity. Current Opinion in Biotechnology, 24(3), 385–
390, https://doi.org/10.1016/j.copbio.2013.02.012
Lu L., Li Z. D., Chen X., Wang H., Dai S., Pan X. Q., Ren Z. Y. J. and Gu J. (2020). Spontaneous solar syngas
production from CO2 driven by energetically favorable wastewater microbial anodes. Joule, 4(10), 2149–
2161, https://doi.org/10.1016/j.joule.2020.08.014
Lyu Z., Shao N., Akinyemi T. and Whitman W. B. (2018). Methanogenesis. Current Biology, 28(13), R727–R732,
https://doi.org/10.1016/j.cub.2018.05.021
Malaeb L., Katuri K. P., Logan B. E., Maab H., Nunes S. P. and Saikaly P. E. (2013). A hybrid microbial fuel
cell membrane bioreactor with a conductive ultrafiltration membrane biocathode for wastewater treatment.
Environmental Science and Technology, 47(20), 11821–11828, https://doi.org/10.1021/es4030113
Mallapaty S. (2020). How China could be carbon neutral by mid-century. Nature, 586(7830), 482–483, https://doi.
org/10.1038/d41586-020-02927-9
Mao Z., Sun Y., Zhang Y., Ren X., Lin Z. and Cheng S. (2021). Effect of start-up process using different
electrochemical methods on the performance of CO2-reducing methanogenic biocathodes. International
Journal of Hydrogen Energy, 46(4), 3045–3055, https://doi.org/10.1016/j.ijhydene.2020.02.002
McCarty P. L. and Smith D. P. (1986). Anaerobic wastewater treatment. Environmental Science & Technology,
20(12), 1200–1206, https://doi.org/10.1021/es00154a002
Nevin K. P., Hensley S. A., Franks A. E., Summers Z. M., Ou J. H., Woodard T. L., Snoeyenbos-West O. L. and
Lovley D. R. (2011). Electrosynthesis of organic compounds from carbon dioxide is catalyzed by a diversity
of acetogenic microorganisms. Applied and Environmental Microbiology, 77(9), 2882–2886, https://doi.
org/10.1128/AEM.02642-10
Patil S. A., Arends J. B. A., Vanwonterghem I., van Meerbergen J., Guo K., Tyson G. W. and Rabaey K. (2015).
Selective enrichment establishes a stable performing community for microbial electrosynthesis of acetate
from CO2. Environmental Science and Technology, 49(14), 8833–8843, https://doi.org/10.1021/es506149d
Perona-Vico E., Blasco-Gomez R., Colprim J., Puig S. and Baneras L. (2019). NiFe -hydrogenases are constitutively
expressed in an enriched Methanobacterium sp. population during electromethanogenesis. Plos One, 14(4),
e0215029, https://doi.org/10.1371/journal.pone.0215029
Pham T. H., Rabaey K., Aelterman P., Clauwaert P., De Schamphelaire L., Boon N. and Verstraete W. (2006).
Microbial fuel cells in relation to conventional anaerobic digestion technology. Engineering in Life Sciences,
6(3), 285–292, https://doi.org/10.1002/elsc.200620121
Rotaru A.-E., Shrestha P. M., Liu F., Shrestha M., Shrestha D., Embree M., Zengler K., Wardman C., Nevin K.
P. and Lovley D. R. (2014a). A new model for electron flow during anaerobic digestion: direct interspecies
electron transfer to Methanosaeta for the reduction of carbon dioxide to methane. Energy and Environmental
Science, 7(1), 408–415, https://doi.org/10.1039/C3EE42189A
Rotaru A. E., Shrestha P. M., Liu F., Markovaite B., Chen S., Nevin K. and Lovley D. (2014b). Direct interspecies
electron transfer between Geobacter metallireducens and Methanosarcina barkeri. Applied and
Environmental Microbiology, 80(15), 4599–4605, https://doi.org/10.1128/AEM.00895-14
Rowe A. R., Xu S., Gardel E., Bose A., Girguis P., Amend J. P. and El-Naggar M. Y. (2019). Methane-linked
mechanisms of electron uptake from cathodes by Methanosarcina barkeri. Mbio, 10(2), e02448-18.
Sangeetha T., Guo Z. C., Liu W. Z., Gao L., Wang L., Cui M. H., Chen C. and Wang A. J. (2017). Energy recovery
evaluation in an up flow microbial electrolysis coupled anaerobic digestion (ME-AD) reactor: role of
electrode positions and hydraulic retention times. Applied Energy, 206, 1214–1224, https://doi.org/10.1016/j.
apenergy.2017.10.026
Saunders S. H., Tse E. C. M., Yates M. D., Otero F. J., Trammell S. A., Stemp E. D. A., Barton J. K., Tender L. M. and
Newman D. K. (2020). Extracellular DNA promotes efficient extracellular electron transfer by pyocyanin in
Pseudomonas aeruginosa biofilms. Cell, 182(4), 919–932, https://doi.org/10.1016/j.cell.2020.07.006
Siegert M., Yates M. D., Spormann A. M. and Logan B. E. (2015). Methanobacterium dominates biocathodic
archaeal communities in methanogenic microbial electrolysis cells. ACS Sustainable Chemistry and
Engineering, 3(7), 1668–1676, https://doi.org/10.1021/acssuschemeng.5b00367
Silvestre G., Fernandez B. and Bonmati A. (2015). Significance of anaerobic digestion as a source of clean energy in
wastewater treatment plants. Energy Conversion and Management, 101, 255–262, https://doi.org/10.1016/j.
enconman.2015.05.033
Stams A. J. M. and Plugge C. M. (2009). Electron transfer in syntrophic communities of anaerobic bacteria and
archaea. Nature Reviews Microbiology, 7(8), 568–577, https://doi.org/10.1038/nrmicro2166
Steinbusch K. J. J., Hamelers H. V. M., Schaap J. D., Kampman C. and Buisman C. J. N. (2010). Bioelectrochemical
ethanol production through mediated acetate reduction by mixed cultures. Environmental Science and
Technology, 44(1), 513–517, https://doi.org/10.1021/es902371e
Storck T., Virdis B. and Batstone D. J. (2016). Modelling extracellular limitations for mediated versus direct
interspecies electron transfer. Isme Journal, 10(3), 621–631, https://doi.org/10.1038/ismej.2015.139
Tang Y.-Q., Shigematsu T., Morimura S. and Kida K. (2015). Dynamics of the microbial community during
continuous methane fermentation in continuously stirred tank reactors. Journal of Bioscience and
Bioengineering, 119(4), 375–383, https://doi.org/10.1016/j.jbiosc.2014.09.014
Thauer R. K., Kaster A.-K., Seedorf H., Buckel W. and Hedderich R. (2008). Methanogenic archaea: ecologically
relevant differences in energy conservation. Nature Reviews Microbiology, 6(8), 579–591, https://doi.
org/10.1038/nrmicro1931
Van Eerten-Jansen M. C. A. A., Heijne A. T., Buisman C. J. N. and Hamelers H. V. M. (2012). Microbial electrolysis
cells for production of methane from CO2: long-term performance and perspectives. International Journal of
Energ Research, 36(6), 809–819, https://doi.org/10.1002/er.1954
Van Eerten-Jansen M. C. A. A., Jansen N. C., Plugge C. M., de Wilde V., Buisman C. J. N. and ter Heijne A.
(2015). Analysis of the mechanisms of bioelectrochemical methane production by mixed cultures. Journal of
Chemical Technology and Biotechnology, 90(5), 963–970, https://doi.org/10.1002/jctb.4413
Varanasi J. L., Veerubhotla R., Pandit S. and Das D. (2019). Biohydrogen production using microbial electrolysis
cell: recent advances and future prospects, chapter 5.7. In: Microbial Electrochemical Technology, S. V.
Mohan, S. Varjani and A. Pandey (eds), Elsevier, The Netherlands, pp. 843–869.
Varrone C., Van Nostrand J. D., Liu W., Zhou B., Wang Z., Liu F., He Z., Wu L., Zhou J. and Wang A.
(2014). Metagenomic-based analysis of biofilm communities for electrohydrogenesis: from wastewater
to hydrogen. International Journal of Hydrogen Energy, 39(9), 4222–4233, https://doi.org/10.1016/j.
ijhydene.2014.01.001
Villano M., Aulenta F., Ciucci C., Ferri T., Giuliano A. and Majone M. (2010). Bioelectrochemical reduction of
CO(2) to CH(4) via direct and indirect extracellular electron transfer by a hydrogenophilic methanogenic
culture. Bioresource Technology, 101(9), 3085–3090, https://doi.org/10.1016/j.biortech.2009.12.077
Walker D. J. F., Nevin K. P., Holmes D. E., Rotaru A.-E., Ward J. E., Woodard T. L., Zhu J., Ueki T., Nonnenmann S.
S., McInerney M. J. and Lovley D. R. (2020). Syntrophus conductive pili demonstrate that common hydrogen-
donating syntrophs can have a direct electron transfer option. ISME Journal, 14(3), 837–846, https://doi.
org/10.1038/s41396-019-0575-9
Wang A., Liu W., Cheng S., Xing D., Zhou J. and Logan B. E. (2009). Source of methane and methods to control
its formation in single chamber microbial electrolysis cells. International Journal of Hydrogen Energy, 34(9),
3653–3658, https://doi.org/10.1016/j.ijhydene.2009.03.005
Wang L. Y., Nevin K. P., Woodard T. L., Mu B. Z. and Lovley D. R. (2016). Expanding the eiet for DIET:
electron donors supporting direct interspecies electron transfer (DIET) in defined co-cultures. Frontiers in
Microbiology, 7, 236.
Wang L., Liu W., He Z., Guo Z., Zhou A. and Wang A. (2017). Cathodic hydrogen recovery and methane conversion
using Pt coating 3D nickel foam instead of Pt-carbon cloth in microbial electrolysis cells. International
Journal of Hydrogen Energy, 42(31), 19604–19610, https://doi.org/10.1016/j.ijhydene.2017.06.019
Wang H., Cai W.-W., Liu W.-Z., Li J.-Q., Wang B., Yang S.-C. and Wang A.-J. (2018). Application of sulfate radicals
from ultrasonic activation: disintegration of extracellular polymeric substances for enhanced anaerobic
fermentation of sulfate-containing waste-activated sludge. Chemical Engineering Journal, 352, 380–388,
https://doi.org/10.1016/j.cej.2018.07.029
Wang L., He Z. W., Guo Z. C., Sangeetha T., Yang C. X., Gao L., Wang A. J. and Liu W. Z. (2019). Microbial
community development on different cathode metals in a bioelectrolysis enhanced methane production
system. Journal of Power Sources, 444, 227306, https://doi.org/10.1016/j.jpowsour.2019.227306
Wang B., Liu W., Cai W., Li J., Yang L., Li X., Wang H., Zhu T. and Wang A. (2020a). Reinjection oilfield wastewater
treatment using bioelectrochemical system and consequent corrosive community evolution on pipe material.
Journal of Bioscience and Bioengineering, 129(2), 199–205, https://doi.org/10.1016/j.jbiosc.2019.09.001
Wang B., Liu W., Varrone C., Yu Z. and Wang A. (2020b). Hydrogen and methane generation from biowaste:
enhancement and upgrading via bioelectrochemical systems. In: Bioelectrochemical Systems: Vol. 2 Current
and Emerging Applications, P. Kumar and C. Kuppam (eds), Springer Singapore, Singapore, pp. 83–130.
https://doi.org/10.1007/978-981-15-6868-8_5
Wang B., Liu W., Zhang Y. and Wang A. (2020c). Bioenergy recovery from wastewater accelerated by solar power:
intermittent electro-driving regulation and capacitive storage in biomass. Water Research, 175, 115696,
https://doi.org/10.1016/j.watres.2020.115696
Wang B., Liu W., Zhang Y. and Wang A. (2020d). Intermittent electro field regulated mutualistic interspecies
electron transfer away from the electrodes for bioenergy recovery from wastewater. Water Research, 185,
116238, https://doi.org/10.1016/j.watres.2020.116238
Wang B., Liu W., Zhang Y. and Wang A. (2022a). Natural solar intermittent-powered electromethanogenesis
towards green carbon reduction. Chemical Engineering Journal, 432, 134369, https://doi.org/10.1016/j.
cej.2021.134369
Wang X. T., Zhang Y. F., Wang B., Wang S., Xing X., Xu X. J., Liu W. Z., Ren N. Q., Lee D. J. and Chen C. (2022b).
Enhancement of methane production from waste activated sludge using hybrid microbial electrolysis
cells-anaerobic digestion (MEC-AD) process-a review. Bioresource Technology, 346, 126641, https://doi.
org/10.1016/j.biortech.2021.126641
Wu W.-M., Jain M. K., De Macario E. C., Thiele J. H. and Zeikus J. G. (1992). Microbial composition and
characterization of prevalent methanogens and acetogens isolated from syntrophic methanogenic granules.
Applied Microbiology and Biotechnology, 38(2), 282–290, https://doi.org/10.1007/BF00174484
Xiao Y., Zhang E., Zhang J., Dai Y., Yang Z., Christensen H. E. M., Ulstrup J. and Zhao F. (2017). Extracellular
polymeric substances are transient media for microbial extracellular electron transfer. Science Advances,
3(7), e1700623, https://doi.org/10.1126/sciadv.1700623
Xu H., Wang K. and Holmes D. E. (2014). Bioelectrochemical removal of carbon dioxide (CO2): an innovative method
for biogas upgrading. Bioresource Technology, 173, 392–398, https://doi.org/10.1016/j.biortech.2014.09.127
Yang H.-Y., Wang Y.-X., He C.-S., Qin Y., Li W.-Q., Li W.-H. and Mu Y. (2020). Redox mediator-modified biocathode
enables highly efficient microbial electro-synthesis of methane from carbon dioxide. Applied Energy, 274,
115292, https://doi.org/10.1016/j.apenergy.2020.115292
Yu Z., Liu W., Shi Y., Wang B., Huang C., Liu C. and Wang A. (2021). Microbial electrolysis enhanced bioconversion
of waste sludge lysate for hydrogen production compared with anaerobic digestion. Science of the Total
Environment, 767, 144344, https://doi.org/10.1016/j.scitotenv.2020.144344
Zhang Y. F., Min B., Huang L. P. and Angelidaki I. (2011). Electricity generation and microbial community
response to substrate changes in microbial fuel cell. Bioresource Technology, 102(2), 1166–1173, https://doi.
org/10.1016/j.biortech.2010.09.044
Zhao Z., Zhang Y., Quan X. and Zhao H. J. B. T. (2015). Evaluation on direct interspecies electron transfer in
anaerobic sludge digestion of microbial electrolysis cell. Bioresource Technology, 200, 235–244, https://doi.
org/10.1016/j.biortech.2015.10.021
Zheng S., Liu F., Wang B., Zhang Y. and Lovley D. R. (2020). Methanobacterium capable of direct interspecies
electron transfer. Environmental Science and Technology, 54(23), 15347–15354, https://doi.org/10.1021/acs.
est.0c05525
Zhou A., Liu Z., Varrone C., Luan Y., Liu W., Wang A. and Yue X. (2018). Efficient biorefinery of waste activated
sludge and vinegar residue into volatile fatty acids: effect of feedstock conditioning on performance and
microbiology. Environmental Science Water Research and Technology, 4, 1819, https://doi.org/10.1039/
C8EW00266E
Ziels R. M., Nobu M. K. and Sousa D. Z. (2019). Elucidating syntrophic butyrate-degrading populations in
anaerobic digesters using stable-isotope-informed genome-resolved metagenomics. Msystems, 4(4), e00159-
19, https://doi.org/10.1128/mSystems.00159-19
Chapter 14
Thermal energy from wastewater
James McQuarrie*
Tetra Tech Engineering Services Company, Denver, CO, USA
*Correspondence: [email protected]
14.1 INTRODUCTION
Governments, power utilities, public institutions, and private corporations are beginning to establish
goals and putting plans into action to decarbonize their footprints. Thankfully, costs for renewable
electricity generation and grid-scale battery energy storage are now competitive with the cost of fossil-
based thermoelectricity and forecasted to trend lower through 2030 (See Chapter 2). This trend in grid
electrical power generation along with other innovations in batteries is allowing other historically
energy intensive societal needs like personal automobile transportation to piggyback onto the low-
carbon energy trend via electrification. Unlike electricity, there is no trend of similar scale to replace
natural gas and other fossil-based combustion fuels for indoor heating and hot water with low-carbon
energy. Yet, energy consumption for indoor heating and hot water is substantial. For scale, consider
Figure 14.1 below that shows in most of the United States, more than half of total residential energy
consumption is for indoor heating and hot water. In temperate climates like the northeastern United
States, more than two-thirds of total household energy consumption is expressly for indoor heat and
hot water. Similarly, most residential consumption for natural gas is for indoor heating and hot water.
Electrical resistance heating and application of heat pumps are two transition alternatives to
the current heavy reliance on combustion-based heating systems discussed above. With electrical
resistance heating, 100% of the electrical energy input is converted to heat output giving the local
system a coefficient of performance (CoP) of 1. Heat-pump based systems can perform at a CoP
commonly in the range of 2–6, depending largely on the temperature and mass flow characteristics of
the connected thermal source. Heat pumps achieve these levels of efficiency since they are designed
to transfer thermal energy from a local source to a sink (i.e., residential or office space) rather than
generate the heat itself from a primary energy source. Figure 14.2 illustrates the range of emission
reduction that can be achieved through application of heat pumps compared with electrical resistance
heating and how water-source heat pumps can provide the most favorable CoP in terms of unit heat
(or cooling) provided per unit of primary energy put into the system. All of the systems summarized in
Figure 14.2 are electrical-based heating systems. Therefore, if the electricity is generated from 100%
renewables, all types of electrical heating systems would be low carbon. However, the figure shows
© 2022 The Editors. This is an Open Access eBook distributed under the terms of the Creative Commons Attribution Licence
(CC BY-NC-ND 4.0), which permits copying and redistribution for non-commercial purposes with no derivatives, provided the
original work is properly cited (https://creativecommons.org/licenses/by-nc-nd/4.0/). This does not affect the rights licensed or
assigned from any third party in this book. The chapter is from the book Pathways to Water Sector Decarbonization, Carbon Capture
and Utilization, Zhiyong Jason Ren and Krishna Pagilla (Eds.)
100%
90%
80%
70%
60%
50%
40%
30%
20%
10%
0%
Northeast Midwest South West Pacific
Fracon of total residenal energy consumpon
Fracon of total residenal natural gas consumpon
Figure 14.1 Fraction of total residential energy consumption and natural gas consumption used for indoor heating
and hot water (adapted from US Energy Information Association, 2018).
that heat pump-based systems reduce the burden of electrical energy consumption compared with
electrical resistance heating.
To achieve deep decarbonization, transition of indoor heating and hot water systems away from
combustion of fossil-based fuels is necessary. In addition, much greater system-level efficiencies are
needed within the built environment that go beyond individual building systems. Rather, over the
next few decades, a new generation of modern district heating (DH) and district energy systems
(DES) are necessary in the major cities throughout the world that connect and couple buildings (i.e.,
demands or loads) with non-primary thermal energy (i.e., sources). Lily Riahi, Policy Unit, Climate,
Water-Sourced (CoP = 3)
Water-Sourced (CoP = 6)
Figure 14.2 Range of CO2 emission reductions achieved through application of heat pump systems based on the
portfolio makeup of the electrical generating plant that supplies the grid.
Energy and Environmentally Sound Technologies UNEP (2015) identifies modern district energy
as the most effective approach for many cities to transition to sustainable heating and cooling, by
improving energy efficiency and enabling higher sharing of renewable thermal energy. Countries
such as Denmark have made modern district energy the cornerstone of their energy policy to reach
their goal of 100% renewable energy, and, similarly, other countries, such as China, are exploring
synergies between high levels of wind production and district heating. Like the modern electrical
grid that consists of more (but smaller) distributed electrical generation stations that put renewable
electricity into the grid, modern DH and DES systems link thermal energy sources to a loop that
distributes thermal energy to meet connector energy load requirements. Cities are responsible for 70%
of the global energy demand. At the same time, cities provide the population density and potential
for economic growth necessary to support and sustain modern DH and DES systems. For obvious
reasons where there is population density, there is also wastewater. Figure 14.3 shows that in addition
to other sources, public investments in wastewater infrastructure can provide value-added service to
society in providing a local source for thermal energy to a campus or district scale DH or DES system.
where Q is heat transfer (watts); m is the mass flow rate of water (g/sec); c is the specific capacity of
water (4.186 J/g-C); ΔT is the temperature drop incurred during heat transfer, Tf minus Ti (C).
As an example, the quantity of heat transferred from a wastewater flow rate of 3.785 million
liters per day (mld) is calculated below and assumes a temperature differential of 4°C across the heat
exchanger:
43.8 L 1000 g 4.186 J
Q= × × × (17C − 13C ) = 733 kilowatts
sec L g −C
In this example, 733 kilowatts (KW) of heat are drawn from the system. Figure 14.5 helps to depict
the range in quantity of heat that could be recovered from an interceptor, at a lift station or from
effluent at a treatment plant. Due to the specific capacity of water and the relatively favorable initial
temperature of wastewater, a considerable quantity of heat can be transferred to meet indoor heating
and hot water needs.
Figure 14.3 Illustration of a modern district energy system where a variety of non-primary thermal energy sources are used in conjunction with a
common system to provide low-carbon heating (and cooling) to end-users (from UNEP, 2016).
Thermal energy from wastewater 253
Figure 14.4 Temperature of sewer water and of the ambient air for two years, sewer is the Upper Schuylkill River
East Side Interceptor Sewer, Philadelphia Water Department (Kohl, 2019).
Figure 14.5 Range of recoverable heat (MW) as a function of flow rate at DH/DES access point and dT drop during
contact with the heat exchanger or heat pump. A ΔT of 2°C is reasonable for an ambient loop DES type system
and ΔT of 4°C is a reasonable upper range for a DH system using a direct heat pump. Two existing DH systems are
shown for reference.
Figure 14.6 Schematic illustration of a fourth generation DH system sourced to a wastewater interceptor and
serving district buildings with building heat and hot water.
The advent of fourth generation district and campus heating systems which operate at much lower
temperature hot water than previous generation systems has allowed for low-grade thermal energy
sources like wastewater to enter the portfolio of sources to supply DH and DES loops (Lund et al., 2018)
Building on the low carbon performance capabilities of fourth generation systems, fifth generation
district energy systems (DES) operate at even lower temperatures and seek to provide a campus with
year-round low carbon heating and cooling (also known as eco-loops). These systems operate on the
premise of an ambient temperature loop that circulates tepid water year-round through a network
of buildings. For example, during winter the loop may circulate 17°C water in the main loop to each
building, a heat pump extracts and amplifies heat from the loop. In mixed building use environments,
some buildings may be doing the opposite where the heat pump is rejecting heat back (e.g., a data
center) into the loop for other buildings to pick up and utilize. During summer, the loop may circulate
at 22°C and within each building the heat pump is rejecting heat into the loop. Throughout, the
wastewater in the interceptor provides the system-level sink or source as needed to balance the overall
needs of the system. Figure 14.7 illustrates an example DES using a wastewater interceptor as heating
source or sink. Within the campus utility plant, wastewater is passed through a heat exchanger where
the heat is exchanged with the ambient circulation loop. A secondary loop in each building utilizes a
local heat pump to provide heating or cooling as required by the building. In the built environment,
DES systems such as these can provide groups of buildings with extremely efficient and low carbon
heating and cooling. An air sourced or ground sourced heat pump (refer to Figure 14.2) can be
installed at the campus utility plant if needed to provide peaking capacity during extreme cold or
warm periods of the year.
Figure 14.7 Schematic illustration of a fifth generation DES system sourced to a wastewater interceptor and
serving district buildings with heating and cooling. Heat is exchanged from sewage to separated ambient loop in
the Central Utility Plant.
Figure 14.8 Energy model output of building campus energy demands (a) and overlay of thermal energy recovered
from a nearby wastewater interceptor (b). The red shows seasonal heating demands, and the blue shows seasonal
cooling demands. The yellow shows the coverage of heating and cooling requirements provided through the
interceptor.
For a given application or site, it is necessary to understand the potential for thermal energy from
wastewater to meet or match building requirements. The assessment must consider the building energy
demand profile and compare it to the diurnal and seasonal profile of the thermal energy capabilities
of the wastewater (or effluent). An overlay of the building energy model demands with the recoverable
thermal energy in the wastewater flow help to determine fit. Figure 14.8(a) provides an example of an
hour-by-hour one-year output of a building energy model (eQuest software, Department of Energy).
The model incorporates local climate data, building details and architectural treatments for energy
efficiency. Figure 14.8(b) provides an overlay of the heat transfer capacity of a nearby interceptor
based on the initial wastewater temperature (Ti) and the estimated dry weather flow pattern in the
interceptor. In this example, a campus DES coupled to the nearby interceptor can satisfy nearly all
the heating requirements and about two-thirds of its summertime cooling requirements. The electrical
input for the pumps can be satisfied with grid or on-site PV. The CoP of the water-sourced heat pumps
helps reduce the overall electrical requirement of the system. The need for natural gas for heating in
winter has been minimized.
Figure 14.9 A large scale heat pump (∼4 MW) extracts heat directly from screened raw wastewater to serve the
Southeast False Creek District (courtesy city of Cancouver).
operation since 2010 and as of 2019 provided space heating and hot water to roughly 5,750,000 ft 2 of
residential and commercial space.
Figure 14.10 provides a schematic illustration of how a heat pump is applied in this instance to
extract heat and boost the temperature in the supply loop. In the source loop shown on the left,
warm wastewater is passed through the low-pressure evaporator side of the heat pump. In this side
of the heat pump cycle, the refrigerant in the pump draws heat out of the passing fluid. The cooled
Figure 14.10 Schematic illustration of a heat pump applied to wastewater in a fourth generation DH system.
wastewater exits the heat pump evaporator and is returned to the sewer. The warmed refrigerant
gas is then compressed and transferred to the supply loop where under high pressure, the now hot
refrigerant gas in the condenser exchanges the heat into the hot water supply line where incoming
return flow from the neighborhood takes heat from the hot refrigerant and exits as hot water in the
supply line. An expansion valve releases the refrigerant back to the low-pressure evaporator side to
begin the cycle again and absorb heat.
The National Western Center (NWC) in Denver, Colorado is a mixed-use redevelopment that
includes event space, arenas, research, and education spaces. The thermal energy from the wastewater
system at the NWC provides an example of a 5th generation ambient loop DES system that supplies
indoor heating during winter and some cooling in summer. The campus utility plant for the site has a
nominal capacity of 4 MW and draws heat (or rejects) from an interceptor that runs under the site. A
two-pipe loop extends across the site and provides a source of indoor heating and cooling to 119,000
square meters of event space and research space.
Figure 14.11 Schematic framework for enabling the integration of low-temperature sources into district energy
system (Bertelsen et al., 2021).
illustrated with these two examples, help to align end-user needs (e.g., building owner) with potential
local low-grade thermal energy source owners (e.g., data center or wastewater sanitation district). In
support of these broader goals, wastewater utilities can establish sewer heat recovery policies which
establish the utility’s position on thermal energy recovery from its infrastructure as well as other
technical, administrative and/or financial requirements.
resource mapping allows the developer to contemplate campus energy options at the early master
planning stages of design. For cities with defined strategic plans in place (e.g., Defining Strategic
Planning), a wastewater utility can provide mapping layers of interceptor systems with wastewater
flows to provide a high-level understanding of available thermal energy available. Under mandates
that require alternatives to fossil natural gas, resource mapping and site proximity to available heating
sources could become a factor in the overall benefit, attraction and appraised value of a build site to
a developer.
Contracted DES 1 2
Partner National Wester Center Wastewater Utility
Design, Build, Finance, Authority Interceptor Owner
Operate, Maintain Overall Campus Operator
(DBFOM)
3
Figure 14.12 General structure of relationships that form the basis of the thermal energy recovery system at NWC.
REFERENCES
Bertelsen N., Mathiesen B. V., Djørup S. R., Schneider N. C. A., Paardekooper S., Sánchez García L., Thellufsen
J. Z., Kapetanakis J., Angelino L. and Kiruja J. (2021). Integrating Low-Temperature Renewables in District
Energy Systems: Guidelines for Policy Makers. International Renewable Energy Agency, Masdar City P.O.
Box 236, Abu Dhabi, United Arab Emirates, Available at https://vbn.aau.dk/en/publications/d68fe575-
1109-40a3-b758-b48af9939f80 (Accessed September, 2021).
Kohl P. (2019). State of the Science and Issues Related to Heat Recovery from Wastewater, Project Number
ENER10C13/4788, Water Research Foundation, Denver, CO, USA.
Lily Riahi, Policy Unit, Climate, Energy and Environmentally Sound Technologies UNEP (2015). District Energy
in Cities: Unlocking the Potential of Energy Efficiency and Renewable Energy, UNEP, Copenhagen Centre
for Energy Efficiency (C2E2), ICLEI – Local Governments for Sustainability and UN-Habitat, United
Nations Environment Programme, Gigiri Nairobi, Kenya.
Local Law 97 (2019). Local Laws for the City of New York. Available at https://www1.nyc.gov/assets/buildings/
local_laws/ll97of2019.pdf
Lund H., Ostergaard P. A., Chang M., Werner S., Svendsen S., Sorknaes P., Thorsen J. E., Hvelplund F., Ole
Gram Mortensen B., Vad Mathiesen B., Bojesen C., Duic N., Zhang X. and Moller B. (2018). The status
of 4th generation district heating: research and results. Energy, 164, 147–159. https://doi.org/10.1016/j.
energy.2018.08.206 (Accessed September 2021).
New York City Mayor’s Office of Sustainability (2016). New York City’s Roadmap to 80×50. Available at
https://www1.nyc.gov/assets/sustainability/downloads/pdf/publications/New%20York%20City’s%20
Roadmap%20to%2080%20x%2050_Final.pdf (Accessed October 2021).
Schmid F. (2008). Sewage Water: Interesting Heat Source for Heat Pumps and Chillers. Proceedings of the 9th
International IEA Heat Pump Conference, Zürich, Switzerland, 20–22 May 2008, pp. 1–12.
US Energy Information Association (2018). Annual household site end-use consumption by fuel in the U.S. – totals,
2015. Available at https://www.eia.gov/consumption/residential/data/2015/index.php?view=consumption
(accessed 29 December 2020)
Wilson M. P. and Worall F. (2021). The heat recovery potential of ‘wastewater’: a national analysis of sewage
effluent discharge temperatures. Environmental Science: Water Research and Technology, 7, 1760–1777.
https://doi.org/10.1039/D1EW00411E, https://doi.org/10.1039/D1EW00411E
Chapter 15
Concept wastewater treatment
plants in China
Jiuhui Qu1,2*, Hongqiang Ren2,3, Hongchen Wang2,4, Kaijun Wang1,2, Gang Yu1,2, Bing Ke1,5,
Han-Qing Yu1,6, Xingcan Zheng1,7 and Ji Li1,8
1School of Environment, Tsinghua University, Beijing 100084, China
2Expert Committee for China’s Concept WWTPs, Beijing 100044, China
3School of the Environment, Nanjing University, Nanjing 212013, China
4School of Environment and Nature Resources, Renmin University of China, Beijing 100872, China
5Administrative Centre for China’s Agenda 21, Ministry of Science and Technology, Beijing 100038, China
6CAS Key Laboratory of Urban Pollutant Conversion, University of Science and Technology of China, Hefei 230026, China
7North China Municipal Engineering Design & Research Institute, Tianjin 300074, China
8School of Environment and Civil Engineering, Jiangnan University, Wuxi 214122, China
*Correspondence: [email protected]
15.1 INTRODUCTION
China has witnessed fast growth in public infrastructure over the past several decades. In order to
maintain the rapid development of cities and industries, the government and experts are seeking to
solve the environmental problems that are far behind its economic development. The total number
of wastewater treatment plants in Chinese cities increased by 10-fold, from 481 to 5640, during the
period from 2000 to 2018. In the meantime, the construction of thousands more wastewater treatment
plants (WWTPs) in the near future is being planned. To explore the suitable wastewater treatment
paradigm for China, the China Concept WWTP Committee (CCWC), initiated by several academic
leaders in 2014, was formed. The committee laid out a grand vision to ponder the goals of wastewater
management in 21st century China. They proposed a new concept for these future WWTPs: sustainable
water quality, resource recovery, energy neutrality, and environmental friendliness, which they called
the ‘Concept Plant’. China’s Concept WWTP is expected to lead a national paradigm shift of the
wastewater industry.
Water pollution control is currently one of the most pressing challenges faced by China (Lu et al.,
2015; Yu et al., 2019). In the battle against environmental pollution, wastewater treatment plays a
pivotal role. Although China has the largest wastewater treatment capacity and market of the world,
the development history of its wastewater industry is actually very short. Wastewater management in
China was almost in blank until 40 years ago when several public environmental incidents raised the
urgency of water environment protection. In the 1980s, the National Environmental Protection Bureau
was set up, and the first large-scale WWTP with a treatment capacity of 260,000 m3/d was constructed
in Tianjin. Since then, accompanied with the rapid economic development and urbanization, the
© 2022 The Editors. This is an Open Access eBook distributed under the terms of the Creative Commons Attribution Licence
(CC BY-NC-ND 4.0), which permits copying and redistribution for non-commercial purposes with no derivatives, provided the
original work is properly cited (https://creativecommons.org/licenses/by-nc-nd/4.0/). This does not affect the rights licensed or
assigned from any third party in this book. The chapter is from the book Pathways to Water Sector Decarbonization, Carbon Capture
and Utilization, Zhiyong Jason Ren and Krishna Pagilla (Eds.)
Figure 15.1 Growth of municipal WWTPs number in China during 2007–2017 (Qu et al., 2019).
amount of municipal wastewater increased drastically, and the wastewater composition became
increasingly complicated because of the entering of more industrial wastewater into the sewers.
To address these challenges, China started to build more centralized WWTPs and supplementary
facilities. The construction speed and WWTPs scale have been increasing continuously over the years
till the end of the ‘Twelfth Five-Year Plan’ period (Figure 15.1).
The water environmental pollution in China has also aggravated the shortage of water resources,
especially in North China regions which face more severe water deficiency. To overcome this
limitation, wastewater reclamation and reuse presents a key pathway. Beijing has been pioneering in
this direction and has achieved great progress in constructing water reclamation infrastructures. In
2016, the Beijing Gaobeidian WWTP, with the treatment capacity of 1 million m3/d, was successfully
upgraded into a reclaimed water plant, announcing a transition of wastewater management in China
from simply treatment to reclamation. However, the overall water reclamation ratio in China is still
very low, and the reclaimed water was mainly reused as landscape water due to relatively low quality.
Currently, the reclaimed wastewater still cannot compete in price with conventional water supply,
and establishment of water reuse infrastructures and program is at slow pace.
After nearly 40 years of remarkable development, China’s wastewater industry has grown into the
largest one in the world. It now possesses more than 5000 municipal WWTPs with a daily treatment
capacity of nearly 200 million m3/d (Figure 15.1). Accordingly, the wastewater treatment ratio has
increased substantially, reaching 90% by 2018. These WWTPs play key roles in water environmental
protection by reducing the environmental release of pollutants. The wastewater management mode
has also changed from the unitary government-dominated construction and operation into a multiple
system that involves both the government and enterprises. Such a transition not only lessened the
financial burden of the government to a certain extent, but also improved the construction and
operation efficiencies of wastewater facilities.
Figure 15.2 (left) The geographic distribution of influent COD and NH3-N concentrations of WWTPs in China; (right)
proportion of WWTPs implementing class 1A effluent standards and the energy consumption intensity of WWTPs
in China (Qu et al., 2019).
considerable gaps in the design and operation performance of the treatment facilities compared to
those in the developed countries. For example, most plant designs and operations did not consider the
sustainability development demand, instead they overstress the pollutants abatement in order to meet
the stringent national Class 1A effluent standards. Therefore, in most WWTPs primary sedimentation
tanks were omitted while delayed aeration and additional biofiltration were widely implemented,
resulting in overtreatment and significantly increased energy/chemical consumption (Figure 15.2).
This was aggravated by the lagged development of a sewer system, especially at county level. As a
consequence, China’s wastewater management suffer from insufficient wastewater collection on one
hand and a lower operating ratio of the WWTPs on the other. Such insufficient municipal wastewater
collection, plus the dilution by stormwater, significantly lowers the organic strength of wastewater
while complicates operation (Figure 15.2).
Consequently, the organic loading in wastewater is typically too low to support efficient
denitrification, and to solve this problem external electron donors such as methanol have to be applied.
In addition, the low organic content and high sand composition of wastewater sludge also prohibits
its anaerobic digestion, a common practice for bioenergy recovery worldwide. It is estimated that less
than 3% of WWTPs in China are equipped with anaerobic digestion, among which a large fraction are
in poor operation (Jin et al., 2014; Zhang et al., 2016). Therefore, there is almost no energy recovery in
China’s WWTPs currently, not to mention the recovery of the nutrient resources. How to improve the
sustainability of wastewater treatment in China remains a critical issue to be addressed.
15.3 WASTEWATER CONCEPT PLANT PROVIDES A VISION AND EXAMPLE FOR FUTURE
DEVELOPMENT
Remarkable progress has been made by China’s wastewater industry in infrastructure construction
and technology innovation. However, with continued population growth and urbanization in the
future, the water shortage will become more severe and urban ecology may become more vulnerable.
Thus, the target of wastewater management is shifting from solely pollutant abatement to water reuse,
resource recovery and water ecology restoration. This has been reflected by the recent policy changes
in China (Wang & Gong, 2018).
For many years, China has been enforcing an end-of-pipe pollution control strategy, that is
emphasizing wastewater treatment and water environmental remediation. However, the overall
environmental quality has shown no obvious improvement. In 2015, the Chinese government issued
the Action Plan on Water Pollution Control (The State Council, 2016), opening a new age of water
environmental protection aimed at improving the quality of the overall water ecological environment
instead of simple water quality control (Hansen et al., 2018; Holdgate, 1987). This means that the
frontier of pollution control will extend from WWTPs to the upstream sewer networks and the
downstream rivers and wetlands.
With the vision of turning a WWTP from a site of pollutant removal into a plant of energy, water
and fertilizer and an integrated part of urban ecology, in 2014 several experts from top institutes,
universities and authority in China jointly proposed the program for constructing a brand-new
‘concept plant’ (Jin et al., 2014). This China Concept WWTP Committee (CCWC) envisioned that the
concept plants will be implemented in 2030–2040, practicing low-carbon concepts, and intensively
applying and demonstrating global advanced technologies that have been and will be engineered,
so as to fully meet China’s sustainable development goal and hopes to become the benchmark of
municipal WWTPs in the world. Over the past few years, the CCWC has gathered global insights
and established cooperation with many domestic institutions. Discussion and exchanges, visits,
collaborative research, formulation of plans, work on engineering practice, and gathering of feedback
have been carried out. In 2015, the CCWC initiated and hosted the ‘Urban Sewage Treatment Concept
Plant’ campus creative design competition under the theme of ‘Concept Plant – Water Future – My
Heart’. The competition was attended by nearly 1000 students from more than 100 universities across
the country, effectively enlightening the thinking of the wastewater industry and conveying the ideas
to society, especially the younger generation (Figure 15.3).
The CCWC has successfully promoted the implementation of the concept WWTP into practice.
The first attempt was completed in Suixian, Henan, and another plant is under construction in Yixing,
Jiangsu. In the near future, the Yixing concept plant will be the most instructive plant, leading the
WWTPs upgrading to a large-scale, sustainable wastewater treatment plant. In 2018, the first concept
plant was built in Suixian County, Henan Province, via a public-private-partnership (PPP) model
(Figure 15.4).
Figure 15.3 International competition of campus creative design ‘the concept WWTP for the future of water –
concept WWTP in My mind’.
Figure 15.4 Sui county NO.3 WWTP aerial view and scenery.
The completed Sui County No. 3 WWTP serves a population of ∼900,000 and treats wastewater
at an average flowrate of 40,000 cubic meters per day (CMD). In the first phase, the average design
flowrate is 20,000 CMD. The plant includes a liquid treatment area, an organic waste processing area,
a constructed wetland, agriculture and sponge city demonstration areas, and an office building and
education center. Wastewater is treated by preliminary treatment (screens and aerated grit chamber),
primary clarification and fermentation, and a step feed activated sludge process with biological
nutrient removal. Secondary effluent is polished by denitrification filters and disinfected by ozonation,
which is also effective for the destruction of trace levels of emerging contaminants. Treated effluent
passes through a constructed wetland, replenishing local surface water bodies. The good effluent
quality makes it possible for potential reuse in industrial applications.
The organic waste processing system is designed to treat 100 tons/day of organic waste. Sludge
produced from wastewater treatment is co-digested with manure collected from livestock and poultry
farms and straw from agricultural operations throughout the county. This uses the DANAS (Dry
ANAerobic System) process, a dry anaerobic digestion technology developed by CSDWS. Design
capacity for the first phase project is 50 tons/day. The central load rate of organic matter is more than
85%. Co-digestion not only mitigates the non-point source pollution problems in the county, but also
produced 510,000 m3 of biogas, 438,765 kWh of electricity and 4500 tons of fertilizer in 2020. The
constructed wetland, agricultural demonstration area (using the organic fertilizer produced onsite),
and sponge city demonstration area together constitute the ecological park, which creates a synergy
between wastewater treatment and the surrounding environment. The office building houses a modern
control center and an exhibition hall that displays the treatment technologies employed at the plant.
It also serves as an education center to demonstrate the importance of environmental protection and
how various resources can be recovered from wastewater and beneficially reused.
The Sui County Concept Factory is combined with the local organic fertilizer factory, adopting
the method of ‘bartering things, leaving biogas in the middle’, that is, the organic fertilizer factory is
responsible for the collection, storage and transportation of livestock and poultry manure, and the
concept factory produces organic fertilizer raw materials to supply the organic fertilizer plant, which
produces biogas for power generation. This method maintains the healthy operation of the organic
matter center and realizes the resource recovery and harmless utilization of sludge. The products
comply with the ‘China Organic Fertilizer Standard (NY525-2012)’ (Figure 15.5).
Having adopted the goals set out by the CCWC, the Sui County No. 3 WWTP project has received
national recognition. Achievements include an energy self-sufficiency of 50% (Figure 15.6).
Another example is Yixing Concept Water Resource Recovery Facility (WRRF). From the beginning
of 2017 to the end of 2019, after five changes in the draft, it was finally completed by the Beijing Municipal
Institute, SUP Atelier and THUPDI Architectural Design Branch. The construction of this plant will
be completed in 2021. Yixing concept plant will not only become a water resource recovery facility, but
also serve as a full-scale R&D center aiming at comprehensive research and verification of emerging
technologies. This brand-new demonstration integrates a pollutant reduction factory with energy, water,
and fertilizer factories, which will become a new type of environmental infrastructure that integrates
with the surrounding neighborhoods. It is hoped that through the Yixing concept plant, the concept of
‘sewage is a resource and sewage treatment plant is a resource factory’ will be clearly conveyed to the
whole society, and the public’s inherent perception of sewage treatment plants will be changed.
The urban sewage resource concept plant that was built in Yixing adopts ‘three-in-one’ construction,
which consists of a water purification center with a capacity of 20,000 tons/day, a collaborative
processing center for organic matter with a capacity of 100 tons/day, and a production-oriented R&D
center. The sewage treatment part has achieved superior nitrogen and phosphorus removal (TN
<3 mg/L, TP <0.1 mg/L, Table 15.1), and its cost performance is significantly better than the current
domestic sewage plants. The organic matter co-processing center can treat sludge and cyanobacteria,
kitchen waste and straw to produce energy (energy self-sufficiency rate >60%) and fertilizer, and the
production-oriented R&D center is composed of two state-of-the-art pilot facilities, displaying the
world’s most advanced sewage treatment technology in real time (Figure 15.7).
In the Yixing concept WWTP, there will be an innovation center for demonstration and
commercialization of the leading-edge technologies with great engineering potential. Those technologies
will be selected and demonstrated in the plant with a wastewater treatment capacity of ∼1000 t/d
for technology showcases. Yixing concept WWTP will serve as a great platform for those innovative
technologies to be applied and promoted in this industry. Another 4–5 concepts plants will be designed
and built in the next few years as well.
REFERENCES
Hansen M. H., Li H. and Svarverud R. (2018). Ecological civilization: interpreting the Chinese past, projecting the
global future. Global Environmental Change, 53, 195–203, https://doi.org/10.1016/j.gloenvcha.2018.09.014
Holdgate M. W. (1987). Our Common Future: The Report of the World Commission on Environment and
Development. Oxford University Press, Oxford & New York.
Jin L., Zhang G. and Tian H. (2014). Current state of sewage treatment in China. Water Research, 66, 85–98,
https://doi.org/10.1016/j.watres.2014.08.014
Lu Y., Song S., Wang R., Liu Z., Meng J., Sweetman A. J. and Wang T. (2015). Impacts of soil and water pollution
on food safety and health risks in China. Environment International, 77, 5–15, https://doi.org/10.1016/j.
envint.2014.12.010
Qu J., Wang H., Wang K., Yu G., Ke B., Yu H. Q. and Gong H. (2019). Municipal wastewater treatment in China:
development history and future perspectives. Frontiers of Environmental Science and Engineering, 13(6),
1–7.
The State Council. (2016). The 13th Five-Year Plan for the Construction of Urban Sewage Treatment and Recycling
Facilities. The State Council of the People’s Republic of China, Beijing (in Chinese).
Wang M. and Gong H. (2018). Imbalanced development and economic burden for urban and rural wastewater
treatment in China – discharge limit legislation. Sustainability, 10(8), 2597, https://doi.org/10.3390/
su10082597
Yu C., Huang X., Chen H., Godfray H. C. J., Wright J. S., Hall J. W. and Taylor J. (2019). Managing nitrogen to
restore water quality in China. Nature, 567(7749), 516–520, https://doi.org/10.1038/s41586-019-1001-1
Zhang Q. H., Yang W. N., Ngo H. H., Guo W. S., Jin P. K., Dzakpasu M. and Ao D. (2016). Current status of urban
wastewater treatment plants in China. Environment International, 92, 11–22, https://doi.org/10.1016/j.
envint.2016.03.024
Chapter 16
Data science tools to enable
decarbonized water and wastewater
treatment systems
Kathryn B. Newhart1*, Amanda S. Hering2 and Tzahi Y. Cath3
1United States Military Academy, West Point, NY, USA
2Baylor University, Waco, TX, USA
3Colorado School of Mines, Golden, CO, USA
*Correspondence: [email protected]
16.1 INTRODUCTION
Data science tools can leverage historical and currently generated data to inform and impact how
water and wastewater distribution and treatment systems are monitored and controlled. Despite
decades of advancement in data-driven modeling of engineered environmental systems, water and
wastewater treatment facilities continue to use basic monitoring, analysis approaches, and control
schema. Conventionally, models in water and wastewater treatment are derived from the fundamental
understanding of the phenomena responsible for the removal of contaminants (e.g., gravity separation
and settling, chemical and microbial kinetics). Due to the large size and complexity of full-scale
treatment facilities, these models are rarely sufficiently accurate for process monitoring and control.
Rather, control thresholds such as upper and lower limits for individual process variables are set based
on historical performance and on operators’ understanding of a specific system to define normal
operating conditions. Such values are static and include a large factor of safety to account for all
possible water quality, environmental, and operational conditions; ultimately substantially reducing
the efficiency of a system. An alternative to these static, physical, mathematical models are empirical,
data-driven models. These ‘intelligent’ models rely on the relationships between variables identified
within a data set without explicitly defining the relationship based on pre-existing knowledge.
Data-driven modeling (DDM) has intensified in recent years as the expense of data collection and
storage has decreased while processing speeds have exponentially increased. However, the water
and wastewater treatment sectors have not fully realized these technological advancements. Manesis
et al. (1998) presumed that the limiting factors to adopting DDM in the treatment industry were: (1)
the underdeveloped field of intelligent control; and (2) the lack of familiarity with DDM by engineers.
Despite increased interest in the scientific literature, full-scale application of DDM in water and
wastewater treatment systems is still limited due to the second of Manesis’ factors. The purpose of this
chapter is to familiarize water treatment engineers with DDM methods to achieve decarbonization
objectives in the water and wastewater treatment sectors.
© 2022 The Editors. This is an Open Access eBook distributed under the terms of the Creative Commons Attribution Licence
(CC BY-NC-ND 4.0), which permits copying and redistribution for non-commercial purposes with no derivatives, provided the
original work is properly cited (https://creativecommons.org/licenses/by-nc-nd/4.0/). This does not affect the rights licensed or
assigned from any third party in this book. The chapter is from the book Pathways to Water Sector Decarbonization, Carbon Capture
and Utilization, Zhiyong Jason Ren and Krishna Pagilla (Eds.)
DDM includes both statistical and machine learning (ML) methods, and while they may seem
similar, there is no single approach that is universally ‘better’ than the other due to differences in
their purpose and requirements. Statistical models are inherently probabilistic models, meaning
that a measure of uncertainty comes automatically with the model. Thus, when used to analyze,
summarize, and draw conclusions from data, statistical models include a margin of error that depends
on the noise in the data. Assumptions about the shape of the distribution of the noise in the data
or the functional form of the relationship between variables (e.g., linear, exponential, polynomial)
are needed for these models to be valid. On the other hand, ML models are incredibly flexible and
can handle modeling nonlinear and complex relationships among variables. They do not require any
assumptions regarding the sampling distribution or shape of relationships between variables. However,
uncertainty quantification cannot be as readily derived from ML models, a large number of internal,
tuning parameters must be selected, and they often require very large sample sizes to fit them. Both
approaches can be used to achieve the same goal and are agnostic to specific processes or systems,
but they differ philosophically with statistical models taking a stochastic approach and ML models
taking an algorithmic approach. A more in-depth discussion of the distinction between statistical and
ML models can be found in Boulesteix and Schmid (2014). Some may argue that ML models are ‘black
boxes’ compared to statistical models; indeed, both types of models are black boxes relative to physics-
based models. The importance of each variable in the model is automatically provided by a statistical
model and takes more work to obtain from a ML model, but both types require some interpretation to
understand each variable’s influence on the response of interest (Ljung, 2010).
The objective of DDM for decarbonization is to minimize energy consumption and inefficiencies,
and maximize resource and energy recovery to ultimately reduce direct and indirect greenhouse gas
(GHG) emissions. Sources of direct GHG emissions include oxidation of organic matter, byproducts
of the biological nitrogen removal processes, and biogas production from anaerobic digestion (AD)
and combustion. Indirect GHG emissions include emissions associated with electrical energy
consumption, external carbon for denitrification, and sludge disposal and recovery (Flores-Alsina et al.,
2011). Simulation studies have theorized the impact of different control schemes on GHG footprint for
water and wastewater utilities (Barbu et al., 2017), but are based on qualitative assumptions regarding
operating costs and effluent quality indexes. This is because simulation studies approximate, but do
not always accurately represent, the true multivariate relationships among variables at full-scale water
treatment plants (WTP) and wastewater treatment plants (WWTP). For example, Oppong et al. (2013)
compared the Pearson correlation coefficients between inputs and outputs of a full-scale AD and
the most popular simulation model (benchmark simulation model no. 2 or BSM2) (Jeppsson et al.,
2007). They found large differences in magnitude and/or direction for all variable pairs. Additionally,
the full-scale outputs did not match the BSM2 outputs. This discrepancy could be caused by many
factors, including the influence of other variables and process disturbances such as change to influent
composition, infrequent sampling of certain variables, and a different operating range at full-scale
than is possible in simulation. Ultimately, the simulation model for this case was overly simplistic
and unable to capture the true behavior of the AD process at full-scale. Dellana and West (2009)
compared the prediction performances of a statistical model and an ML model on simulated and real
WWTP data, showing conflicting results as to which model can ‘best’ predict effluent nitrogen and
phosphorus. While the statistical model had a lower prediction error for some simulated cases, the
ML model had the lowest prediction error for all cases using the real WWTP data. Ultimately, DDM
at full-scale for decarbonization must explicitly target features known to impact energy consumption
and GHG emissions; however, the actual impact may be difficult to extrapolate for an individual
facility.
The work presented in this chapter is intended to introduce the reader to DDM methods that can
be used within a larger decarbonization strategy and considerations to apply them appropriately. This
chapter is organized into five sections. In Section 16.2, data preparation, common DDM methods, and
metrics for comparing model performance are presented. In Section 16.3, unit processes in WTP and
WWTP are discussed in which the Section 16.2 methods may be used to maximize decarbonization.
In Section 16.4, recommendations for full-scale implementation of DDM are presented, and Section
16.5 includes concluding remarks.
distribution of error and differences in R 2 across multiple models. These plots can answer diagnostic
questions such if there are outliers, particular groups of observations that are over or underestimated,
or a range of y i values with greater variability. R 2 does have some limitations, such as being sensitive
to outliers, not providing a good measure of the magnitude of the differences, and not penalizing
more complex models that have more parameters to estimate. Thus, it is important to understand the
underlying differences between two models and their predictions before comparing their R 2 values, or
use a different metric for evaluation.
It is expected for a model to have a higher R 2 on the training data compared to the testing data.
However, a large difference in training and testing R 2 values can indicate that the model is overfit to
the data. When a model is overfit, there are more parameters in the model than necessary to capture
the overall pattern. Figure 16.2c is an example of overfitting where the model has too many parameters,
and consequently fits to error in the data. A model can also be underfit (Figure 16.2a), where there are
an insufficient number of model parameters to adequately capture changes in the dependent variable.
Given that R 2 will be higher for an overfit model than a balanced model, R 2 should always be
complemented with other measures of error. For example, model fitting criteria such as Akaike
Information Criterion (AIC, Equation (16.2)) (Akaike, 1974) and Bayesian Information Criterion
Figure 16.2 Example of an underfit model (orange), a robust/balanced model (blue), and an overfit model (Green).
(BIC, Equation (16.3)) (Schwarz, 1978) are metrics that balance model error against the number of
parameters in a model, as follows:
SSE (16.2)
AIC = n ⋅ ln +2⋅ p
n
SSE (16.3)
BIC = n ⋅ ln + ln (n) ⋅ p
n
where n is the number of observations, and p is the number of model parameters. The difference
between AIC and BIC is the penalty for the number of parameters. When comparing different models
(e.g., number of parameters for a given model type), the model with the lowest error and fewest
parameters will have the lowest AIC or BIC. When compared for the same model, AIC will choose
a model with more inputs than BIC. In this case, it is suggested that models are selected that are
generally favored by both AIC and BIC (Burnham & Anderson, 2004; Kuha, 2004; Vrieze, 2012). For
example, if AIC is minimized with five parameters and BIC is minimized with three parameters, then
a model with four parameters may be best.
To better understand the magnitude of model error, metrics that are not normalized (R 2) or
penalized (AIC, BIC) can be used to measure the difference (or squared difference) between the actual
and predicted observations. Mean absolute error (MAE, Equation (16.4)), mean squared error (MSE,
Equation (16.5)), and root mean squared error (RMSE, Equation (16.6)) are examples of such metrics:
∑
n
yi − yˆ i
MAE = i =1 (16.4)
n
SSE
MSE = (16.5)
n
SSE
RMSE = (16.6)
n
The individual metric selected depends on the desired sensitivity to large errors. For example, MAE
depends on absolute errors as opposed to the squared errors of MSE and RMSE, so it is less influenced
by large differences between actual and modeled values. An additional consideration is whether the
metric is applied to the training or testing data. When MAE, MSE, or RMSE are used on both training
and testing data in a single publication, some authors will use MAPE, MSPE, and RMSPE to indicate
the prediction or testing metric. However, it is also common to see AIC, BIC, and/or R 2 used as the
training metrics and MAE, RMSE, and/or RMSE as the testing metrics.
detecting faults, including noisy variables that do not change substantially over the monitoring period,
makes it more difficult to detect faults (Harrou et al., 2021).
Several methods to approach the problem of dimension reduction by feature selection are
described here. Statistical dimension reduction methods such as correlation coefficients and principal
component analysis (PCA) can be used prior to model building. Stepwise variable selection and lasso
modeling approaches are both oftentimes used in the modeling step to reduce the number of variables
in a model. A comparable stepwise approach for variable selection in ML models is also described.
where x and y are the sample mean of the x i and y i values, respectively. Variables that are not linearly
related may still be related but have a relatively low r. Spearman’s rank-order correlation coefficient
(rs, Equation (16.8)) is a nonparametric variation of r that is able to measure the strength and direction
of a monotonic relationship between two variables. For example, as X increases, Y increases for all
X and Y, but not necessarily linearly. The Spearman coefficient achieves this by comparing the ranks
(i.e., position when observations are ranked from smallest to largest) of each pair of observations, as
opposed to the values themselves, as follows:
2
6 ∑ ( xirank − yirank ) (16.8)
rs = 1 −
n (n2 − 1)
where x irank is the rank of x i, and similarly for y irank. Depending on the direction and magnitude of
the expected correlation as well as the sample size, different correlation coefficients may be more
appropriate. Figure 16.3 illustrates how Spearman is less sensitive to small sample size and nonlinear
behavior than Pearson. However, both are unable to quantify relationships that both increase and
decrease in direction. It is important to note that different correlation coefficient values should not
be directly compared. For example, Pearson’s r values should only be compared to other Pearson’s r
values to determine if one pair of features is more highly linearly correlated than another.
Figure 16.3 Examples of Pearson’s r and Spearman’s rs for different sample sizes, n = 10 (top) and n = 100 (bottom),
and data with and without noise (black dots) compared to a fitted linear regression (dashed line).
Figure 16.4 (a) Two nonlinear functions (y1 and y2) and a linear combination of the nonlinear functions (y 3) are
plotted as functions of t; (b) PC1 scores and PC2 scores are the scaled observations from (a) multiplied by the
variable loadings (i.e., rotation matrix) for PC1 and PC2, respectively. The percent of total variation captured by each
PC is in parenthesis.
treatment system, y1 could be taken as the ideal growth rate of algae, y 2 as solar irradiation, and y3
as the actual growth rate. Even though y1 and y 2 are nonlinear functions, Figure 16.4b shows how
their linear combination (y3) is captured in the first component. The remaining variation in the three-
variable system not explained by PC1 is captured by PC2.
The most popular applications of PCA are dimension reduction for multiple regression models
(Section 16.2.4.1) (Wallace et al., 2016; Wang et al., 2017) and general process insight (Corominas
et al., 2018). For modeling, the PCs with the largest loadings (i.e., magnitude of variance in a
particular direction) are retained for model building. The specific number of components retained
for model building depends on the percentage of variation described by the PCA subspace, and
usually a value in the range of 90–99% is used as a threshold for the number of components
retained.
Some limiting factors of conventional PCA, as well as the majority of standard statistical methods for
water and wastewater applications, are the assumptions of stationarity (constant mean and variance),
linearity, and independence over time. Modifications such as rolling training windows, nonlinear
dimension reduction methods, and lagging observations can help approximate the conditions required
for methods such as PCA (Kazor et al., 2016; Odom et al., 2018). Newhart et al. (2019) describes these
adaptations in detail for municipal wastewater treatment.
16.2.4.2 Lasso
Statistical models such as multiple regression are most often fit using ordinary least squares (OLS),
which estimates the parameters such that the SSE between the actual and predicted values are
minimized. While the OLS model fitting approach results in unbiased estimators, measurement error
in the training dataset can produce high variance, which makes the model difficult to generalize
(James et al., 2013). Another approach to statistical model fitting that introduces a small amount
of bias but that reduces variability and model complexity is lasso (Least Absolute Shrinkage and
Selection Operator) (Tibshirani, 1996). Lasso performs variable selection and parameter estimation
simultaneously, as opposed to performing these tasks in two separate steps.
Lasso shrinks unimportant predictors’ coefficients to zero, thereby selecting only those variables
with nonzero coefficients to remain in the model; however, conventional lasso does not always select
the correct subset of variables. For example, if two variables are highly correlated, they will both be
included in the final variable selection. To address this, Zou (2006) proposed adaptive lasso, which
has been used for fault detection in a sequencing-batch membrane bioreactor for municipal wastewater
treatment (Newhart et al., 2020). Fused lasso is another variation of lasso that can address timeseries
data where sequential time-varying coefficients are expected to be similar. In fused lasso, the difference
between adjacent coefficients are penalized as opposed to the coefficients themselves (Hastie et al.,
2015; Tibshirani et al., 2005). An example of fused lasso in WWTP for fault isolation can be found in
Klanderman et al. (2020). A variation of lasso that is useful in the biological sciences is group lasso.
Group lasso identifies groups of variables that are jointly in or out of the model (Yuan & Lin, 2006).
An example of group lasso in WWTP can be found in Bai et al. (2019).
When cyclic patterns are evident in variables observed over time and their corresponding
autocorrelation plots, a multivariate regression with sine and cosine terms (i.e., Fourier series) can be
used:
K K
2π kx 2π kx
f ( x) = ∑
k =1
αk cos
T
+ ∑β sin
k =1
k
T
+ ε (16.9)
where K is the number of cosine and sine pairs; x is the time in some period T; and α k and βk are
estimated model coefficients. For example, Newhart et al. (2020) used a linear combination of sine,
cosine, and process variables to model ammonia concentration in an activated sludge system, and
the sine and cosine terms captured the diurnal variation in ammonia concentration. In this case, T
was 1440 minutes (equating to 1 day for the length of a single cycle), and x was the minute of a day.
Variable selection methods can be used to choose K.
network configurations defined by the use of different activation functions in different layers. The
literature has not yet established the ‘best’ neural network for water and wastewater treatment; thus,
it is important to trial a range of options when developing a predictive model for a specific application.
Figure 16.5 Example of (a) triangular membership functions for fuzzification and (b) a Gaussian membership
function for defuzzification. A ‘++’ or ‘−−’ indicates a slight change to power and a ‘+++’ or ‘−−−’ indicates a
substantial change to power in the corresponding direction (increase or decrease).
are linguistic values of the membership functions. In contrast, TS fuzzy rules use similar ‘if’ logic and
a mathematical equation (e.g., constant, linear, nonlinear combination of input variables) (Takagi &
Sugeno, 1983). For example, ‘if the acid flow is low, then pH = k · flowz + c,’ where k, z, and c are fitted
model parameters.
The steps to develop an intelligent FL controller are (Manesis et al., 1998):
(1) Divide variables into manipulated and controlled. A controlled variable quantifies a
characteristic of a system (e.g., performance, water quality). A manipulated variable is adjusted
to keep a controlled variable at its set point. For example, a recirculation pump flow rate is an
example of a manipulated variable, while the concentration of suspended solids, which varies
with the recirculation pump flow rate, is a controlled variable.
(2) Establish a set of linguistic descriptors for each manipulated variable (e.g., high, normal, low),
which are understood by plant operators. The granularity is directly related to the number of
descriptors, although between three and five is appropriate for most control applications. For
each set, determine a membership function (Ross, 2010), but the individual function is less
important than the number of linguistic descriptors in a set (Sadollah, 2018).
(3) Define if-then rules to form the knowledge base using the linguistic descriptors for manipulated
and controlled variables. The form of if-then rules depends on whether the system is Mamdani
or TS.
(4) Select a method of weighted averaging for the membership functions of the manipulated and
controlled variables (e.g., max-min, center of gravity (COG)).
Adaptive neuro-fuzzy inference systems (ANFIS) are five-layer networks that combine the advantages
of ANN and TS FL by fuzzifying and defuzzifying the inputs and outputs to an ANN, respectively,
to improve prediction accuracy for noisy data (Abraham, 2005). Due to this hybrid structure, ANFIS
is considered to be a universal estimator (Jang et al., 1997). The same method of training ANN (e.g.,
back-propagation) is used to the tune the FL parameters; however, the membership function itself
must be defined. Additionally, the TS rules established in the first layer of ANFIS no longer have
the advantage of interpretability compared to conventional FL models, and alternative variable
importance approaches must be used to understand the input-output relationships. Applications of
ANFIS in water and wastewater treatment are described in Sections 16.3.2 and 16.3.4.
Figure 16.6 Example of a decision tree for determining which base (sodium hydroxide, NaOH, or calcium hydroxide,
CaOH) to add based on water quality characteristics. Each node (circle) represents a binary classifier. The orange
nodes represent the path taken to reach an action of ‘add NaOH’ for water with a pH of 7 and alkalinity of 150 mg/L.
variant of bagging in which a weighted average is used to aggregate the results for regression models,
and the re-sampling of a bootstrap sample changes with each model fitting iteration. By including
more incorrectly predicted observations in subsequent training steps, models are created that can
handle unique cases. Additional classifier algorithms have been developed since bagging and boosting
with the most popular being the AdaBoost algorithm, which uses a weighted average of a series of
single binary classifiers that are determined by more heavily weighting incorrectly classified samples
of previous classifiers (Freund & Schapire, 1997).
The most popular decision tree approach in water and wastewater treatment are random forests. A
random forest (RF) is the average of a large number of decision trees created by recursively subsetting
input variables and random resampling of training observations (i.e., bootstraping) (Breiman, 2001).
The weights of each binary node from the fitting algorithm (specifically boosting or AdaBoost) can
be used to determine the importance of a specific predictor in a model. However, input variable
weights will vary depending on how nodes are split and therefore can be an inconsistent indicator of
variable importance. Finally, to fit an RF model, the number of trees, the maximum depth of a tree,
the minimum number of samples to form a split at a node, and the maximum number of variables to
evaluate the best split are important hyperparameters that require tuning, resulting in a direct tradeoff
between computational burden and accuracy.
16.2.6 Optimization
Optimization algorithms have two major practical applications: optimization of a predictive model (i.e.,
lowest error by adjusting internal model parameters or hyperparameters) (Le et al., 2019) or finding the
optimum set of inputs for an existing predictive model. This chapter focuses on the latter application
of optimization as it is more relevant to the objectives of decarbonization. It is the combination of the
data-driven predictive model and optimization that constitute model predictive control.
Metaheuristic algorithms are non-exact frameworks designed to search a solution space for the global
optimum without calculating every possible solution. Given that there are very few exact mechanistic
models for water and wastewater processes that are sufficiently accurate for monitoring and control at
full-scale (Newhart et al., 2019), metaheuristic algorithms are used to identify the optimum solution
for the predictive models described in Section 16.2.4. The three most popular algorithms used in the
water and wastewater distribution and treatment literature are genetic algorithms, particle swarm
optimization, and simulated annealing, which will be presented in this section. For additional exact
and heuristic methods not listed here, Beheshti and Shamsuddin (2013) provide a more complete list
of optimization methods.
where W is the inertial weight; c1 is the influence of the local minimum on the velocity vector (i.e., self-
confidence factor); c2 is the influence of the global minimum on the velocity vector (swarm confidence
factor); and r1 and r 2 are randomly generated numbers between 0 and 1.
where ΔD is the change of distance between states, T is a synthetic temperature that represents the
range of solution space considered for a different state, and R(0, 1) is a random number between 0
and 1. The probabilistic approach is important to avoid being stuck at local minima by exploring a
reasonable space of solutions to find a global minimum. At each step of SA, a neighboring state (s*) is
compared to the current state (s) and probabilistically decides between moving the system to state s*
or staying in-state s. These probabilities ultimately lead the system to move to states of lower energy.
Typically, this step is repeated until the system reaches a state that is sufficient for the application,
or until a given computation budget (e.g., number of iterations) has been exhausted. To strategically
achieve a global optimum when the objective function is declining slowly, an ‘annealing schedule’ can
be used in which T is iteratively reduced when the objective function plateaus. However, if the initial
step size between states is not sufficiently small, then there is no guarantee that the global minimum
will be found. In practice, the computational requirements of such granularity generally exceed the
improvements in performance (Trosset, 2001). In conventional applications, SA continues to iterate
until no change in the objective function is found for 300 iterations (Prakash et al., 2008). If the search
space is generally smooth or if there are multiple local minimums, SA could terminate early or be
stuck in a local minimum. In these cases, PSO may be a better alternative. Section 16.3.3 describes
how SA can be used for aeration control in WWTP.
Table 16.1 Examples of data science water applications for pump optimization.
scientific literature are discussed below and summarized in Table 16.1. Commercial optimization
software for water distribution systems, the infrastructure (digital and physical) to support the
software, labor, and training for operators typically have a 2–5 year payback period, with energy cost
reductions ranging from 5–15% (Badruzzaman et al., 2014).
Torregrossa et al. (2017) developed an FL pump performance metric that monitors efficiency and
recommends preventative or immediate maintenance accounting for flow conditions. To do this, an
efficiency index based on the mass of water lifted and energy consumed is calculated, and a rolling
median is used to distinguish the long-term trend from fluctuations attributed to changing conditions.
Multiple consecutive days with negative short-term fluctuations is indicative of needed maintenance.
The immediacy of the maintenance response is determined by an FL system that weighs the long-
and short-term efficiency, and the economic consequences of maintenance versus replacement are
evaluated assuming maintenance is able to restore the pump to a baseline efficiency compared to a
new, more efficient pump.
Sadatiyan Abkenar et al. (2015) used a GA approach to optimize the pumping schedule for two
pumps in a hydraulic model of a moderate size water distribution system in Monroe, MI, USA,
simultaneously minimizing energy while including an additional penalty for high pressures. A
continuous approach that used pairs of start and stop times as genes produced infeasible solutions
(i.e., conflicting ON or OFF times). To mitigate this, any mutation that produced an infeasible solution
was ‘repaired’ prior to calculating the fitness of the solution. An alternative discrete approach that
used a binary ON or OFF indicator for 1-hour intervals, where each interval is a gene, only produced
feasible solutions.
Kebir et al. (2014) modeled a full-scale WWTP that relied on a sequential ON/OFF influent
pumping strategy, which is inherently inefficient, and proposed a new FL controller that adjusts a
pump’s VFD by the deviation from average height of an upstream reservoir; reporting a hypothetical
40% reduction in energy. Zhang et al. (2012) used an ANN to develop an energy consumption model
for a given flowrate, pump configuration for parallel operation, and upstream reservoir levels. They
then determined the optimum pump schedule for a given flowrate, desired reservoir level, and physical
constraints of the system using PSO; reporting a hypothetical 8–24% reduction in energy.
An important consideration for large WTP or WWTP is the cost of energy, especially if the cost
of energy changes throughout the day or if utilities are billed based on monthly energy consumption
peaks. Authors in scientific literature largely neglect changing energy costs with time. Rather, energy
consumption models are developed based on the proxy of a VFD frequency or the energy rating of
individual pieces of equipment. In most cases, minimizing energy consumption will result in the lowest
costs; however, forecasting models may need to incorporate a variation of a cost function that describes
true cost instead of using consumption as a proxy. For example, initiating pumping may not be the
most energy efficient action at a given time but may reduce pumping costs over the course of a day if
the immediate demand (when the tank levels must be lowered) coincides with increased energy cost.
16.3.1.1 Coagulation
Coagulation is the process in WTP (and in some instances WWTP) in which chemical (i.e., coagulant)
is added to destabilize colloidal and suspended particulate matter, allowing the particles to aggregate
(i.e., floc) and be more easily removed by gravity due to the larger, neutrally-charged aggregate mass.
The generation and transportation of chemicals for coagulation and flocculation can account for
5–20% of a WTP’s carbon footprint (Biswas & Yek, 2016); therefore, precision chemical treatment
could account for significant cost and carbon savings, depending on the size of the treatment facility
and initial water quality. To reduce the amount of chemical used to treat water, dose control strategies
must be designed that can adjust for non-ideal physiochemical reaction kinetics due to poor mixing
and changing water quality. However, this is rarely the case in full-scale treatment. In WTP, chemical
dosing is primarily flow-paced in which a concentration of chemical per unit volume of water is
maintained by adjusting the flow rate of a chemical dosing pump proportional to the flow of water.
The concentration setpoint is usually only adjusted when a major water quality change or process
upset occurs because the identification of an ‘ideal’ dose in a laboratory bench-scale experiment
(i.e., jar tests) is time-consuming and labor-intensive, and results can greatly differ from full-scale.
Therefore, the use of data-driven methods of chemical dosing could significantly improve treatment
stability and reduce the carbon footprint of treatment facilities. Examples in the scientific literature
are discussed below.
The application of ANN to predict coagulant dosing is not a novel concept. Van Leeuwen et al.
(1999) were able to predict alum dose for a given water quality using historical jar test data and an
ANN; although a multiple linear regression model achieved similar results. Ten years later, Maier
et al. (2009) used the same data as Van Leeuwen to predict treated water quality (turbidity, color,
pH, UV-254, residual alum) and optimal alum dose using a DNN (two-layer ANN) and was able to
reduce the standard deviation of the prediction error by 37%. Zangooei et al. (2016) used historical
jar test data to predict turbidity using pH, initial turbidity, temperature, type of coagulant (e.g., solid
or liquid poly aluminum chloride from different vendors), and concentration of coagulant. An MLP
with two hidden layers outperformed an RBF ANN and FL regression model and required less time
to train. Similarly, Wu and Lo (2008) found that an ANN outperformed an ANFIS prediction model
for treated water quality when influent water quality data were available. In the absence of real-time
water quality, the ANFIS model was able to more accurately predict treated water quality based on
historical trends and the present-day dose. When historical dosing data were available, Wu and Lo
(2010) found that the inclusion of the previous timestep’s coagulant dose (output variable of DNN
model) reduced testing error.
Chen and Hou (2006) observed that multiple regression models were able to predict coagulant dose
and pH adjustment dose for surface water using historical data. However, two models were developed
separately for low and high influent turbidity conditions. Chen and Hou furthered their work to
adjust feedback control parameters using Mamdani FL in order to minimize coagulant dosing while
simultaneously achieving effluent turbidity and pH goals. Bello et al. (2014) proposed a linearized TS
fuzzy model predictive control strategy to improve coagulant dose control stability by maintaining the
surface charge and pH of the treated water. Depending on the quality of the available data and online
instrumentation, FL controllers can improve precision and stability over conventional cascade control.
In order to minimize coagulant dosing, influent water quality, final water quality, and the coagulant
dose need to be aggregated to train a predictive model for treated water quality. Predictive model
options include multiple regression, ANN, or DNN. To bring the predictive power of the models to a
full-scale utility, the best predictive model can then be incorporated into a control strategy in a variety
of ways. The most basic control option is a standard cascade control in which the coagulant dose
is increased when a treated water quality variable such as turbidity exceeds a threshold. However,
this requires a well-understood dose-response relationship. A strict rule to increase the dose when
treated water quality goals are not met could increase the concentration of coagulant beyond the need
to satisfy electroneutrality, thereby causing effluent turbidity to worsen as particles re-stabilize in
suspension (Tchobanoglous et al., 2014). The proposed FL controllers could prevent such overdosing
by including rule sets that account for the worsening water quality, but this approach would require
more complex programming and a method for tuning consequent statement parameters. Adjustment
can be done manually within the existing structure of the existing DCS at WTP, but will need to
be done externally if an ANFIS is used. The same concern of programmatic complexity is raised if
additional ANN or DNN are developed to identify the required coagulant dose for a given effluent
water quality. In this case, the controller must operate on a separate server system, which can provide
outputs to an existing cascade control strategy.
16.3.2 Nitrification
Biological nutrient removal (BNR) is the most expensive, variable, and difficult-to-model process in
WWTP; yet it is required at the majority of modern facilities around the world in order to achieve the
required nitrogen and phosphorus removal. The difficulty in modeling and control is due to two factors
common to most WWTP processes: lack of reliable instrumentation and non-ideal process conditions
at full scale. The microbial solid–liquid matrix where the treatment takes place (i.e., activated sludge
or AS) interferes with common in-line instrumentation measurements due to biofilm growth on the
instrument itself and competing ions or solids interference. Instrumentation that utilizes light rather
than ion transfer, such as dissolved oxygen (DO) concentration, are sufficiently robust to provide
reliable measurements with less frequent cleaning and maintenance. DO is a critical water quality
parameter to measure in activated sludge systems because the availability of specific forms of oxygen
determines the active microorganisms and, consequently, specific contaminant transformation.
Aqueous oxygen available as free oxygen (O2) will increase the DO concentration and is an indicator
of aerobic conditions. When oxygen is only available in the form of nitrate (NO3), conditions are
anoxic. When no oxygen is available, conditions are anaerobic. It is the strategic alternation of these
oxidation conditions that transforms contaminants of concern (namely carbon, nitrogen, phosphorus,
and, to a lesser extent, sulfur) into the gas or solid phase, and thus reduces the aqueous concentration.
While DO sensors can ensure aeration conditions are met, the measurement itself is a proxy for
the completion of the contaminant transformation. For example, low strength wastewater (e.g., low
concentration of organic materials) will not require as much oxygen in order to achieve treatment
goals; however, aeration will continue to be provided to maintain a DO setpoint in the majority of
systems regardless of demand.
A sequencing-batch reactor (SBR) is a commonly used wastewater treatment technology in which
a single biologically-active, completely-mixed reactor undergoes a sequence of different operating
conditions to achieve contaminant removal. The most common control strategy for SBRs uses timed
sequences with distinct DO concentration setpoints for each phase in the treatment cycle. The DO
setpoints are determined by operator experience and a general knowledge of the environmental
conditions required at each phase, each of which activates a unique set of microorganisms. This
control strategy requires only one in-line instrument (the DO sensor) and can ensure the desired
treatment is achieved under stable influent conditions. However, DO is a surrogate for the actual
contaminants removed in the process and cannot guarantee that effluent water quality standards are
met. Historically, this uncertainty has been addressed by increasing DO setpoints to fully oxidize
chemical contaminants and ensure microbial processes are not substrate-limited. This approach
increases the energy consumption of the treatment process, accounting for 35–50% of a wastewater
utility’s total energy (Newhart et al., 2020) and the second largest operational cost behind labor
(Lindtner et al., 2008). To reduce the energy consumption associated with aeration in SBRs and other
secondary biological treatment systems (i.e., conventional and novel activated sludge configurations),
new intelligent monitoring and control strategies are needed. Examples from literature are discussed
below and summarized in Table 16.2.
Traoré et al. (2005) proposed an FL aeration control strategy for a step-fed, pilot-scale SBR treating
municipal wastewater. The rules for the FL DO controller determined the air flow to maintain a
DO setpoint from the measured DO and the cycle phase. Compared to an ON/OFF DO control
approach (when DO measurement exceeds setpoint, turn off air; when DO measurement is less
than the setpoint, turn on air) and conventional proportional-integral-derivative (PID) control, the
fuzzy controller was able to maintain the DO setpoint with greater precision over a wider range of
environmental conditions. The addition of pH and oxygen uptake rate (OUR) to the fuzzy rule set
could shorten aerated cycle times and further reduce energy consumption (Puig et al., 2006). Ferrer
et al. (1998) used a similar fuzzy DO controller for a pilot BARDENPHO activated sludge system;
showing similar improvement in precision compared to an ON/OFF approach with energy savings of
up to 40%. Du et al. (2018) developed an RBF NN to adjust cascade control parameters to improve
DO controller performance, including significantly reduced variability (67% for dry weather flow,
59–93% for wet weather) and slightly reduced aeration energy (100 kWh/d).
An alternative to adjusting individual DO controllers is to identify the optimum operating strategy
given a system-wide model. To accomplish this, Asadi et al. (2017) compared MARS, ANN, and RF,
among others, of DO in the Detroit Water and Sewerage Department’s secondary aeration basins.
MARS predicted DO and other effluent water quality variables better than ANN and RF (using
MAE and R 2). Using the MARS predictive model, they then compared two sets of weights, one that
emphasized the best treated water quality and one that emphasized energy consumption, using SA.
When optimizing for best water quality, they showed that it was possible to reduce air flow rate by
30% without compromising treated water quality. However, nutrients were not considered, which is
an important driver in aeration requirements and strategy for the majority of WWTP. Additionally,
Asadi et al. (2017) concluded that more frequent sampling of the influent variables was required for
ML models compared to statistical models. This comparison holds when the assumptions made about
the shape of the relationships among the variables are true. In general, the simpler assumptions made
by statistical models can more accurately fill gaps than ML models when data are sparse. In contrast,
at a minimum ML models require examples of conditions in training data for reasonably accurate
predictions of similar conditions in testing data.
retention times, DDM of the AD process could provide insight for more efficient operation. Examples
from the scientific literature are discussed below and summarized in Table 16.3.
Turkdogan-Aydınol and Yetilmezsoy (2010) developed a multiple input-multiple output (MIMO)
FL model to predict biogas production, which outperformed a multiple non-linear regression model.
Polit et al. (2002) used a mechanistic mass balance model with fuzzy pH and temperature coefficient
adjustments to predict biogas production, and this approach had the ability to track gas production
under load adjustments better than the mechanistic model alone. Holubar et al. (2003) used a
hierarchical system of ANN to predict volatile fatty acid (VFA) production and pH followed by a
biogas production and composition forecast of an AD during start-up and stabilization. Operating
parameters were adjusted by a one-at-a-time search algorithm to simultaneously maximize organic
loading rate (OLR) and methane production.
By applying PSO to an ANN model of biogas production, Akbaş et al. (2015) identified the
operational conditions that produced the highest percent methane and biogas production. Compared
to the average of the historical data, the optimum conditions increased the methane fraction in the
biogas by 5% and biogas production by 64%. To achieve this, daily averages of sludge loading rate,
temperature, pH, total solids, total volatile solids, VFA, alkalinity, solids retention time (SRT), and
OLR were used as inputs to prediction models for percent methane and biogas production. Dimension
reduction for input variable selection was applied using a boosting tree algorithm (Breiman, 1996),
which improved prediction performance.
Both FL and ANN can be used for prediction, but while ANN has been shown to be more precise,
FL is able to better handle variability in the inputs and outputs (Kambalimath & Deka, 2020; Özcan
et al., 2009). There is extensive research demonstrating the efficiency of ANN to predict biogas
production and AD performance (Levstek & Lakota, 2010). Therefore, there is a boom of hybrid
fuzzy models, such as ANFIS, to address AD systems (Abrahart et al., 2008). Perendeci et al. (2009)
showed that an ANFIS model was able to predict effluent COD of a seasonal anaerobic wastewater
treatment system, improving performance by adding input variables, including an indicator of
whether the system was under start-up or pseudo-steady-state conditions using 10days of historical
COD data.
Unlike the case of coagulation in WTP (Section 16.3.2), a WWTP has some control over the organic
loading rate, temperature, and SRT depending on the number and size of AD available. To utilize fully
DDM for decarbonization, predictive models should be fit for biogas production (both quantity and
quality, such as specific methane mass flow rate) using optimization methods such as PSO or GA to
identify the ideal operating conditions. Individual models can also be developed for variables that
cannot be quickly measured but are critical to understanding performance, such as VFA.
• goals and key performance indicators (KPI), both plant-wide and process-specific, that address
treatment and energy performance (e.g., kWh/MG, gCO2/MG);
• limiting technological factors, including operational constraints, instrumentation, data
management, control system structure, and cybersecurity constraints;
• unknowns, such as impacts on other processes or the rate of adoption of new technology.
Once the project constraints above are identified, DDM method selection begins in earnest. The
general steps include:
(1) Identify predictor (input) and response (output) variables that would integrate easily with the
existing control strategy and provide substantial benefit.
(2) Develop practical and sound data blending protocols that consider real world implementation
(i.e., merge observations when variables have different sampling frequencies). This includes
careful consideration for when laboratory data become available.
(3) Perform variable selection to minimize the amount of error introduced into the DDM by
irrelevant inputs.
(4) Pilot different modeling frameworks, both simple and advanced, to assess the best candidates
for the desired response. If optimization is the ultimate goal, follow the predictive model
development with an experiment that lends itself towards identifying the ‘best’ optimization
algorithm for the given problem.
(5) Tune predictive and optimization models by increasing and decreasing model complexity to
provide the most accurate performance on testing data, which is not used in model fitting.
(6) Program the model onto a server (using a programming language such as R or Python), which
can export data to the data archiving system used at the specific utility. The data archiving
system frequently has access to the SCADA system without posing a security risk.
(7) Monitor the stability of the prediction over time, and identify unforeseen contingencies that
need to be integrated with the new control strategy prior to full-scale deployment.
(8) Schedule monitoring periods for the model developer, control expert, and operations to
simultaneously observe full-scale implementation. These periods should span weeks to
thoroughly evaluate different environmental conditions and should then progressively increase
in runtime until the operational staff are comfortable with unsupervised operation.
(9) Compare the impact of the new DDM control strategy to the original strategy using pre-
established KPIs. If the new DDM control strategy meets or exceeds the KPIs of the original
strategy, then the previous steps can be repeated to incorporate additional predictions or rely
more directly on the predictions by eliminating layers in a control loop.
16.5 CONCLUSIONS
Interest in integrating DDM into WTP and WWTP is growing rapidly, but utilities are largely
overwhelmed by the task. Simultaneous development of good, internal data management protocols
and the application of DDM to water and wastewater treatment challenges could dramatically reduce
the carbon cost of clean water. Given that water treatment currently consumes 3% of electrical energy
generated in the US each year and is projected to increase to 6% due to increasing demand and
intensity of treatment processes (i.e., higher quality effluent) (Chaudhry & Shrier, 2010), data-driven
process optimization is the proverbial ‘low-hanging-fruit’ for carbon and cost reduction. There is
a large body of scientific literature in which conventional and novel DDM methods are applied to
engineered environmental systems like WTP and WWTP; however, the heuristic nature of many ML
approaches make declaring any one method the ‘best’ for a specific process application impossible.
Experimentation with an individual utility’s datasets using published literature as guidelines is truly
the ‘best’ framework. Fundamentally, given existing machinery and treatment technologies, it is the
investment in people and improved operational strategies that will help WTP and WWTP achieve
their full treatment potential with the smallest environmental impact.
ACKNOWLEDGEMENTS
This work was supported by the National Science Foundation PFI:BIC Award No: 1632227; the
National Science Foundation Engineering Research Center program under cooperative agreement
EEC-1028968 (ReNUWIt).
REFERENCES
Abraham A. (2005). Adaptation of fuzzy inference system using neural learning. In: Fuzzy Systems Engineering,
N. Nedjah and L. de Macedo Mourelle (eds), Springer, Berlin/Heidelberg, Germany, pp. 53–83.
Abrahart R. J., See L. M. and Solomatine D. P. (2008). Practical Hydroinformatics: Computational Intelligence and
Technological Developments in Water Applications. Springer Science & Business Media, Berlin/Heidelberg,
Germany.
Akaike H. (1974). A new look at the statistical model identification. IEEE Transactions on Automatic Control,
19(6), 716–723, https://doi.org/10.1109/TAC.1974.1100705
Akbaş H., Bilgen B. and Turhan A. M. (2015). An integrated prediction and optimization model of biogas
production system at a wastewater treatment facility. Bioresource Technology, 196, 566–576, https://doi.
org/10.1016/j.biortech.2015.08.017
Asadi A., Verma A., Yang K. and Mejabi B. (2017). Wastewater treatment aeration process optimization: a
data mining approach. Journal of Environmental Management, 203, 630–639, https://doi.org/10.1016/j.
jenvman.2016.07.047
Badruzzaman M., Cherchi C., Oppenheimer J., Gordon M., Bunn S. and Jacangelo J. G. (2014). Implementation
of energy and water quality management systems modified with a GHG module. Proceedings of the Annual
Conference & Exposition, Boston, MA, USA.
Bai H., Zhu R., An H., Zhou G., Huang H., Ren H. and Zhang Y. (2019). Influence of wastewater sludge properties
on the performance of electro-osmosis dewatering. Environmental Technology, 40(21), 2853–2863, https://
doi.org/10.1080/09593330.2018.1455744
Barbu M., Vilanova R., Meneses M. and Santin I. (2017). Global evaluation of wastewater treatment plants control
strategies including CO2 emissions. IFAC-PapersOnLine, 50(1), 12956–12961, https://doi.org/10.1016/j.
ifacol.2017.08.1800
Beale R. and Jackson T. (1990). Neural Computing – An Introduction. Taylor & Francis, UK.
Beheshti Z. and Shamsuddin S. M. (2013). A review of population-based meta-heuristic algorithm. International
Journal of Advances in Soft Computing and its Applications, 5, 1–35.
Bello O., Hamam Y. and Djouani K. (2014). Fuzzy dynamic modelling and predictive control of a coagulation
chemical dosing unit for water treatment plants. Journal of Electrical Systems and Information Technology,
1(2), 129–143, https://doi.org/10.1016/j.jesit.2014.08.001
Bishop C. M. (1995). Neural Networks for Pattern Recognition. Oxford University Press, Inc., USA.
Biswas W. K. and Yek P. (2016). Improving the carbon footprint of water treatment with renewable energy: a Western
Australian case study. Renewables: Wind, Water, and Solar, 3(1), 14, https://doi.org/10.1186/s40807-016-0036-2
Boulesteix A.-L. and Schmid M. (2014). Machine learning versus statistical modeling. Biometrical Journal, 56(4),
588–593, https://doi.org/10.1002/bimj.201300226
Breiman L. (1996). Bagging predictors. Machine Learning, 24(2), 123–140.
Breiman L. (2001). Random forests. Machine Learning, 45(1), 5–32, https://doi.org/10.1023/A:1010933404324
Burnham K. P. and Anderson D. R. (2004). Multimodel inference: understanding AIC and BIC in model selection.
Sociological Methods and Research, 33(2), 261–304, https://doi.org/10.1177/0049124104268644
Chaudhry S. and Shrier C. (2010). Energy sustainability in the water sector: challenges and opportunities.
Proceedings of Annual Conference and Exposition, 2010 Annual Conference and Exposition of the American
Water Works Association, Chicago, IL.
Chen C.-L. and Hou P.-L. (2006). Fuzzy model identification and control system design for coagulation chemical
dosing of potable water. Water Science and Technology: Water Supply, 6(3), 97–104, https://doi.org/10.2166/
ws.2006.782
Cherchi C., Badruzzaman M., Oppenheimer J., Bros C. M. and Jacangelo J. G. (2015). Energy and water quality
management systems for water utility’s operations: a review. Journal of Environmental Management, 153,
108–120, https://doi.org/10.1016/j.jenvman.2015.01.051
Corominas L., Garrido-Baserba M., Villez K., Olsson G., Cortés U. and Poch M. (2018). Transforming data into
knowledge for improved wastewater treatment operation: a critical review of techniques. Environmental
Modelling and Software, 106, 89–103, https://doi.org/10.1016/j.envsoft.2017.11.023
Dellana S. A. and West D. (2009). Predictive modeling for wastewater applications: linear and nonlinear approaches.
Environmental Modelling and Software, 24(1), 96–106, https://doi.org/10.1016/j.envsoft.2008.06.002
Du X., Wang J., Jegatheesan V. and Shi G. (2018). Dissolved oxygen control in activated sludge process using
a neural network-based adaptive PID algorithm. Applied Sciences, 8(2), 261, https://doi.org/10.3390/
app8020261
Eberhart R. and Kennedy J. (1995). A new optimizer using particle swarm theory. Proceedings of the Sixth
International Symposium on Micro Machine and Human Science, IEEE, Nagoya, Japan, pp. 39–43.
Ferrer J., Rodrigo M. A., Seco A. and Penya-Roja J. M. (1998). Energy saving in the aeration process by fuzzy logic
control. Water Science and Technology, 38(3), 209–217, https://doi.org/10.2166/wst.1998.0210
Fiter M., Güell D., Comas J., Colprim J., Poch M. and Rodríguez-Roda I. (2005). Energy saving in a wastewater
treatment process: an application of fuzzy logic control. Environmental Technology, 26(11), 1263–1270,
https://doi.org/10.1080/09593332608618596
Flores-Alsina X., Corominas L., Snip L. and Vanrolleghem P. A. (2011). Including greenhouse gas emissions
during benchmarking of wastewater treatment plant control strategies. Water Research, 45(16), 4700–4710,
https://doi.org/10.1016/j.watres.2011.04.040
Freund Y. and Schapire R. E. (1997). A decision-theoretic generalization of on-line learning and an application to
boosting. Journal of Computer and System Sciences, 55(1), 119–139, https://doi.org/10.1006/jcss.1997.1504
Friedman J. H. (1991). Multivariate adaptive regression splines. Annals of Statistics, 19(1), 1–67.
Hering A. S. (2021). Fault isolation, chapter 3. In: Statistical Process Monitoring Using Advanced Data-Driven
and Deep Learning Approaches, F. Harrou, Y. Sun, A. S. Hering, M. Madakyaru and A. Dairi (eds), Elsevier,
The Netherlands, pp. 71–117, https://doi.org/10.1016/B978-0-12-819365-5.00009-7.
Hassan R., Cohanim B., De Weck O. and Venter G. (2005). A comparison of particle swarm optimization and
the genetic algorithm. 46th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics and Materials
Conference, Austin, Texas. American Institute of Aeronautics and Astronautics.
Hastie T. and Tibshirani R. (1999). Generalized Additive Models. Chapman & Hall/CRC, Boca Raton, FL, USA.
Hastie T., Tibshirani R. and Wainwright M. (2015). Statistical Learning with Sparsity: The Lasso and
Generalizations. Routledge, United Kingdom.
Helsel D. and Hirsch R. (2002). Statistical methods in water resources. In: Techniques of Water-Resources
Investigations of the United States Geological Survey. Hydrologic Analysis and Interpretation, USGS, Reston,
VA, pp. 209–218. Available at https://pubs.er.usgs.gov/publication/twri04A3
Holubar P., Zani L., Hager M., Fröschl W., Radak Z. and Braun R. (2003). Start-up and recovery of a biogas-reactor
using a hierarchical neural network-based control tool. Journal of Chemical Technology and Biotechnology,
78(8), 847–854, https://doi.org/10.1002/jctb.854
Huang M., Han W., Wan J., Ma Y. and Chen X. (2016). Multi-objective optimisation for design and operation
of anaerobic digestion using GA-ANN and NSGA-II. Journal of Chemical Technology and Biotechnology,
91(1), 226–233, https://doi.org/10.1002/jctb.4568
Hüllermeier E. (2015). From knowledge-based to data-driven fuzzy modeling. Informatik-Spektrum, 38(6), 500–
509, https://doi.org/10.1007/s00287-015-0931-8
Jackson J. E. (1991). A User’s Guide To Principal Components. John Wiley & Sons, Inc.
James G., Witten D., Hastie T. and Tibshirani R. (2013). An Introduction to Statistical Learning. Springer, New
York, NY, USA.
Jang J.-S. (1993). ANFIS: adaptive-network-based fuzzy inference system. IEEE Transactions on Systems, Man,
and Cybernetics, 23(3), 665–685, https://doi.org/10.1109/21.256541
Jang J.-S. R., Sun C.-T. and Mizutani E. (1997). Neuro-fuzzy and Soft Computing: A Computational Approach to
Learning and Machine Intelligence. Prentice Hall, Upper Saddle River, NJ, USA.
Jeppsson U., Pons M. N., Nopens I., Alex J., Copp J. B., Gernaey K. V., Rosen C., Steyer J. P. and Vanrolleghem P.
A. (2007). Benchmark simulation model no 2: general protocol and exploratory case studies. Water Science
Technology, 56(8), 67–78, https://doi.org/10.2166/wst.2007.604
John G. H., Kohavi R. and Pfleger K. (1994). Irrelevant features and the subset selection problem. In: Machine
Learning Proceedings of the Eleventh International Conference, W. W. Cohen and H. Hirsh (eds), Morgan
Kaufmann Publishers, San Francisco, CA, pp. 121–129.
Kambalimath S. and Deka P. C. (2020). A basic review of fuzzy logic applications in hydrology and water resources.
Applied Water Science, 10(8), 191, https://doi.org/10.1007/s13201-020-01276-2
Kasabov N. K. (1996). Foundations of Neural Networks, Fuzzy Systems, and Knowledge Engineering. MIT Press,
Cambridge, MA, USA.
Kazor K., Holloway R. W., Cath T. Y. and Hering A. S. (2016). Comparison of linear and nonlinear dimension reduction
techniques for automated process monitoring of a decentralized wastewater treatment facility. Stochastic
Environmental Research and Risk Assessment, 30(5), 1527–1544, https://doi.org/10.1007/s00477-016-1246-2
Kebir F. O., Demirci M., Karaaslan M., Ünal E., Dincer F. and Arat H. T. (2014). Smart grid on energy efficiency
application for wastewater treatment. Environmental Progress & Sustainable Energy, 33(2), 556–563,
https://doi.org/10.1002/ep.11821
Khataee A. R. and Kasiri M. B. (2011). Modeling of biological water and wastewater treatment processes
using artificial neural networks. CLEAN – Soil, Air, Water, 39(8), 742–749, https://doi.org/10.1002/
clen.201000234
Kira K. and Rendell L. A. (1992). A practical approach to feature selection. In D. Sleeman and P. Edwards (eds),
Machine learning proceedings 1992. Morgan Kaufmann, pp. 249–256. Available at https://doi.org/10.1016/
B978-1-55860-247-2.50037-1
Kirkpatrick S., Gelatt C. D. and Vecchi M. P. (1983). Optimization by simulated annealing. Science, 220(4598),
671–680, https://doi.org/10.1126/science.220.4598.671
Klanderman M., Newhart K. B., Cath T. and Hering A. S. 2020 Case studies in real-time fault isolation in a
decentralized wastewater treatment facility. Journal of Water Process Engineering. 38(2020) 101556, https://
doi.org/10.1016/j.jwpe.2020.101556
Kuha J. (2004). AIC and BIC: comparisons of assumptions and performance. Sociological Methods & Research,
33(2), 188–229, https://doi.org/10.1177/0049124103262065
Kurek W. and Ostfeld A. (2013). Multi-objective optimization of water quality, pumps operation, and storage
sizing of water distribution systems. Journal of Environmental Management, 115, 189–197, https://doi.
org/10.1016/j.jenvman.2012.11.030
Le L. T., Nguyen H., Dou J. and Zhou J. (2019). A comparative study of PSO-ANN, GA-ANN, ICA-ANN, and ABC-
ANN in estimating the heating load of buildings’ energy efficiency for smart city planning. Applied Sciences,
9(13), 2630, https://doi.org/10.3390/app9132630
Levstek T. and Lakota M. (2010). The use of artificial neural networks for compounds prediction in biogas from
anaerobic digestion – a review. Agricultura, 7, 15–22.
Lindtner S., Schaar H. and Kroiss H. (2008). Benchmarking of large municipal wastewater treatment plants
treating over 100 000 PE in Austria. Water Science and Technology, 57(10), 1487–1493, https://doi.
org/10.2166/wst.2008.214
Ljung L. (2010). Perspectives on system identification. Annual Reviews in Control, 34(1), 1–12, https://doi.
org/10.1016/j.arcontrol.2009.12.001
Lundberg S. M. and Lee S.-I. (2017). A unified approach to interpreting model predictions. In I. Guyon, U. V.
Luxburg, S. Bengio, H. Wallach, R. Fergus, S. Vishwanathan and R. Garnett (eds), Advances in neural
information processing systems. Curran Associates, Inc., Vol. 30, pp. 4765–4774. Available at https://
proceedings.neurips.cc/paper/2017/file/8a20a8621978632d76c43dfd28b67767-Paper.pdf
Maier R. M., Pepper I. L. and Gerba C. P. (2009) Environmental microbiology. Academic press. Cambridge,
Massachusetts, vol. 397.
Maleki A., Nasseri S., Aminabad M. S. and Hadi M. (2018). Comparison of ARIMA and NNAR models for
forecasting water treatment plant’s influent characteristics. KSCE Journal of Civil Engineering, 22(9), 3233–
3245, https://doi.org/10.1007/s12205-018-1195-z
Manesis S., Sapidis D. and King R. (1998). Intelligent control of wastewater treatment plants. Artificial Intelligence
in Engineering, 12(3), 275–281, https://doi.org/10.1016/S0954-1810(97)10002-4
Metropolis N., Rosenbluth A. W., Rosenbluth M. N., Teller A. H. and Teller E. (1953). Equation of state
calculations by fast computing machines. The Journal of Chemical Physics, 21(6), 1087–1092, https://doi.
org/10.1063/1.1699114
Newhart K. B., Holloway R. W., Hering A. S. and Cath T. Y. (2019). Data-driven performance analyses of
wastewater treatment plants: a review. Water Research, 157, 498–513, https://doi.org/10.1016/j.watres.
2019.03.030
Newhart K. B., Marks C. A., Rauch-Williams T., Cath T. Y. and Hering A. S. (2020). Hybrid statistical-machine
learning ammonia forecasting in continuous activated sludge treatment for improved process control.
Journal of Water Process Engineering, 37, 101389, https://doi.org/10.1016/j.jwpe.2020.101389
Newhart K. B., Goldman-Torres J. E., Freedman D. E., Wisdom K. B., Hering A. S. and Cath T. Y. (2021). Prediction
of peracetic acid disinfection performance for secondary municipal wastewater treatment using artificial
neural networks. ACS ES&T Water, 1(2), 328–338, https://doi.org/10.1021/acsestwater.0c00095
Nielsen M. A. (2015). Neural Networks and Deep Learning. Determination Press, San Francisco, CA, USA.
Nybo P. J., Kallesøe C. S. and Lauridsen K. G. (2017). Method for Operating A Wastewater Pumping Station. US
Patent No. 9 719 241. US Patent and Trademark Office, Alexandria, VA, USA.
Odom G. J., Newhart K. B., Cath T. Y. and Hering A. S. (2018). Multistate multivariate statistical process control.
Applied Stochastic Models in Business and Industry, 34(6), 880–892, https://doi.org/10.1002/asmb.2333
Olsson G. (2006). Instrumentation, control and automation in the water industry – state-of-the-art and new
challenges. Water Science and Technology, 53(4–5), 1–16, https://doi.org/10.2166/wst.2006.097
Oppong G., Montague G. A., O’Brien M., McEwan M. and Martin E. B. (2013). Towards advanced control for
anaerobic digesters: volatile solids inferential sensor. Water Practice and Technology, 8(1) 7–17, https://doi.
org/10.2166/wpt.2013.002
Özcan F., Atiş C. D., Karahan O., Uncuoğlu E. and Tanyildizi H. (2009). Comparison of artificial neural network
and fuzzy logic models for prediction of long-term compressive strength of silica fume concrete. Advances in
Engineering Software, 40(9), 856–863, https://doi.org/10.1016/j.advengsoft.2009.01.005
Perendeci A., Arslan S., Tanyolaç A. and Çelebi S. S. (2009). Effects of phase vector and history extension on
prediction power of adaptive-network based fuzzy inference system (ANFIS) model for a real scale anaerobic
wastewater treatment plant operating under unsteady state. Bioresource Technology, 100(20), 4579–4587,
https://doi.org/10.1016/j.biortech.2009.04.049
Polit M., Estaben M. and Labat P. (2002). A fuzzy model for an anaerobic digester, comparison with experimental
results. Engineering Applications of Artificial Intelligence, 15(5), 385–390, https://doi.org/10.1016/S0952-
1976(02)00091-X
Ross T. J. (2010). Properties of Membership Functions, Fuzzification, and Defuzzification. In Fuzzy Logic with
Engineering Applications. John Wiley & Sons Ltd., pp. 89–116. https://doi.org/10.1002/9781119994374.ch4
Prakash A., Shukla N., Shankar R. and Tiwari M. K. (2008). Solving machine loading problem of FMS: an
artificial intelligence (AI) based random search optimization approach. In: Handbook of Computational
Intelligence in Manufacturing and Production Management, D. Laha and P. Mandal (eds), IGI Global,
Hershey, Pennsylvania, USA, pp. 19–43.
Puig S., Corominas L., Traore A., Colomer J., Balaguer M. D. and Colprim J. (2006). An on-line optimisation of
a SBR cycle for carbon and nitrogen removal based on on-line pH and OUR: the role of dissolved oxygen
control. Water Science and Technology, 53(4–5), 171–178, https://doi.org/10.2166/wst.2006.121
Reeves C. and Rowe J. E. (2002). Genetic Algorithms: Principles and Perspectives: A Guide to GA Theory. Springer
Science & Business Media. Boston, MA, USA.
Ren Z., Liner B., Ferguson C., Fisher A., Newhart K., Wang M. and Sharpless C. (2020). Report on NSF Mid-scale
Research Infrastructure Workshop for Intelligent Water Systems. NSF Award #CBET 2035032 (online).
Sadatiyan Abkenar S. M., Stanley S. D., Miller C. J., Chase D. V. and McElmurry S. P. (2015). Evaluation of genetic
algorithms using discrete and continuous methods for pump optimization of water distribution systems.
Sustainable Computing: Informatics and Systems, 8, 18–23, https://doi.org/10.1016/j.suscom.2014.09.003
Sadollah A. (2018). Which membership function is appropriate in fuzzy system? In: Fuzzy Logic Based in Optimization
Methods and Control Systems and its Applications, A. Sadollah (ed.), InTech, London, UK, pp. 3–6.
Schmidhuber J. (2015). Deep learning in neural networks: an overview. Neural Networks, 61, 85–117, https://doi.
org/10.1016/j.neunet.2014.09.003
Schwarz G. (1978). Estimating the dimension of a model. The Annals of Statistics, 6(2), 461–464, https://doi.
org/10.1214/aos/1176344136
Shankar V. K. A., Umashankar S., Paramasivam S. and Hanigovszki N. (2016). A comprehensive review on energy
efficiency enhancement initiatives in centrifugal pumping system. Applied Energy, 181, 495–513, https://doi.
org/10.1016/j.apenergy.2016.08.070
Shapley L. S. (1951). Notes on the N-Person Game–II: The Value of an N-Person Game. Rand Corporation, Santa
Monica, California, USA.
Sharma S., Sharma S. and Athaiya A. (2020). Activation functions in neural networks. International Journal
of Engineering Applied Sciences and Technology, 04(12), 310–316, https://doi.org/10.33564/IJEAST.2020.
v04i12.054
Shi C. Y. (2011). Mass Flow and Energy Efficiency of Municipal Wastewater Treatment Plants. IWA Publishing,
London, UK.
Steinberg D. and Colla P. (1995). CART: tree-structured non-parametric data analysis. San Diego, CA: Salford
Systems.
Sutton C. D. (2005). Classification and regression trees, bagging, and boosting. In: Handbook of Statistics. Data
Mining and Data Visualization, C. R. Rao, E. J. Wegman and J. L. Solka (eds), Elsevier, The Netherlands,
pp. 303–329.
Takagi T. and Sugeno M. (1983). Derivation of fuzzy control rules from human operator’s control actions. IFAC
Proceedings Volumes, 16(13), 55–60, https://doi.org/10.1016/S1474-6670(17)62005-6
Tchobanoglous G., Stensel H. D., Tsuchihashi R., Burton F. L. and Abu-Orf M. (2014). Fundamentals of chemical
coagulation. In: Wastewater Engineering: Treatment and Resource Recovery, G. Bowden, W. Pfrang and
Metcalf & Eddy (eds), McGraw-Hill Education, New York, NY, USA, pp. 460–473.
Tibshirani R. (1996). Regression Shrinkage and Selection via the Lasso. Journal of the Royal Statistical Society.
Series B (Methodological), 58(1), 267–288.
Tibshirani R., Saunders M., Rosset S., Zhu J. and Knight K. (2005). Sparsity and smoothness via the fused lasso.
Journal of the Royal Statistical Society: Series B (Statistical Methodology), 67(1), 91–108, https://doi.
org/10.1111/j.1467-9868.2005.00490.x
Torregrossa D., Hansen J., Hernández-Sancho F., Cornelissen A., Schutz G. and Leopold U. (2017). A data-driven
methodology to support pump performance analysis and energy efficiency optimization in waste water
treatment plants. Applied Energy, 208, 1430–1440, https://doi.org/10.1016/j.apenergy.2017.09.012
Traoré A., Grieu S., Puig S., Corominas L., Thiery F., Polit M. and Colprim J. (2005). Fuzzy control of dissolved
oxygen in a sequencing batch reactor pilot plant. Chemical Engineering Journal, 111(1), 13–19, https://doi.
org/10.1016/j.cej.2005.05.004
Trosset M. W. (2001). What is simulated annealing? Optimization and Engineering, 2(2), 201–213, https://doi.
org/10.1023/A:1013193211174
Turkdogan-Aydınol F. I. and Yetilmezsoy K. (2010). A fuzzy-logic-based model to predict biogas and methane
production rates in a pilot-scale mesophilic UASB reactor treating molasses wastewater. Journal of
Hazardous Materials, 182(1), 460–471, https://doi.org/10.1016/j.jhazmat.2010.06.054
Van Leeuwen J., Chow C. W. K., Bursill D. and Drikas M. (1999). Empirical mathematical models and artificial
neural networks for the determination of alum doses for treatment of southern Australian surface waters.
Aqua, 48(3), 115–127.
Vrieze S. I. (2012). Model selection and psychological theory: a discussion of the differences between the Akaike
information criterion (AIC) and the Bayesian information criterion (BIC). Psychological Methods, 17(2),
228–243, https://doi.org/10.1037/a0027127
Wallace J., Champagne P. and Hall G. (2016). Multivariate statistical analysis of water chemistry conditions in
three wastewater stabilization ponds with algae blooms and pH fluctuations. Water Research, 96, 155–165,
https://doi.org/10.1016/j.watres.2016.03.046
Wang X., Ratnaweera H., Holm J. A. and Olsbu V. (2017). Statistical monitoring and dynamic simulation of
a wastewater treatment plant: a combined approach to achieve model predictive control. Journal of
Environmental Management, 193, 1–7, https://doi.org/10.1016/j.jenvman.2017.01.079
Wise B. M. and Gallagher N. B. (1996). The process chemometrics approach to process monitoring and fault
detection. Journal of Process Control, 6(6), 329–348, https://doi.org/10.1016/0959-1524(96)00009-1
Wu G.-D. and Lo S.-L. (2008). Predicting real-time coagulant dosage in water treatment by artificial neural
networks and adaptive network-based fuzzy inference system. Engineering Applications of Artificial
Intelligence, 21(8), 1189–1195, https://doi.org/10.1016/j.engappai.2008.03.015
Wu G.-D. and Lo S.-L. (2010). Effects of data normalization and inherent-factor on decision of optimal coagulant
dosage in water treatment by artificial neural network. Expert Systems with Applications, 37(7), 4974–4983,
https://doi.org/10.1016/j.eswa.2009.12.016
Yuan M. and Lin Y. (2006). Model selection and estimation in regression with grouped variables. Journal of the
Royal Statistical Society: Series B (Statistical Methodology), 68(1), 49–67, https://doi.org/10.1111/j.1467-
9868.2005.00532.x
Zadeh L. A. (1973). Outline of a new approach to the analysis of complex systems and decision processes. IEEE
Transactions on Systems, Man and Cybernetics, 1100, 38–45.
Zangooei H., Delnavaz M. and Asadollahfardi G. (2016). Prediction of coagulation and flocculation processes
using ANN models and fuzzy regression. Water Science and Technology, 74(6), 1296–1311, https://doi.
org/10.2166/wst.2016.315
Zhang Z., Zeng Y. and Kusiak A. (2012). Minimizing pump energy in a wastewater processing plant. Energy, 47(1),
505–514, https://doi.org/10.1016/j.energy.2012.08.048
Zheng F. and Zhong S. (2011). Time series forecasting using a hybrid RBF neural network and AR model based on
binomial smoothing. World Academy of Science, Engineering and Technology, 75, 1471–1475.
Zou H. (2006). The adaptive lasso and its oracle properties. Journal of the American Statistical Association,
101(476), 1418–1429, https://doi.org/10.1198/016214506000000735
Chapter 17
Decarbonization policies and water
sector opportunities
Jason A. Turgeon1*, Steven A. Conrad2 and Peter A. Vanrolleghem3
1United States Environmental Protection Agency, Boston, MA, USA
2Colorado State University, Fort Collins, CO, USA
3Université Laval, Quebec, QC, Canada
*Correspondence: [email protected]
17.1 INTRODUCTION
The water sector sits in an unusual place in most economies. The sector, which encompasses drinking
water, wastewater treatment/water resource recovery, and stormwater management, is driven by a
wide variety of policies. Foremost among these are regulatory mandates to protect water quality and
human health. However, many other types of policies impact the sector. These include directives
to keep prices low (often below the true cost of water) for reasons of social equity or economic
development, inclusion in efforts to make local governments more sustainable, and efforts to make
essential services more resilient to natural disasters, pandemics, and security threats.
In this chapter, we will discuss some of the many policies and policy responses that impact efforts
to decarbonize the water sector and examine ways in which the sector can respond to sometimes
competing or conflicting policy demands. The scope and scale of policymaking varies greatly based on
the involvement of readers in the overall process, therefore we wrote this chapter with the following
readership in mind: Some readers may be involved in policy making but may not come from an
engineering background. Other readers may be engineers or scientists attempting to respond to policy
directives in the design and operation of water utilities. Whether the reader is approaching this chapter
from a policy development perspective or an operational perspective, we hope to convey a handful of
key concepts in this chapter. A unifying thread among these concepts is the need to seek out multiple
benefits whenever possible. Decarbonization is often a co-benefit of other policies discussed in this
chapter. The web of policy directives demands an integrated approach to resource management.
© 2022 The Editors. This is an Open Access eBook distributed under the terms of the Creative Commons Attribution Licence
(CC BY-NC-ND 4.0), which permits copying and redistribution for non-commercial purposes with no derivatives, provided the
original work is properly cited (https://creativecommons.org/licenses/by-nc-nd/4.0/). This does not affect the rights licensed or
assigned from any third party in this book. The chapter is from the book Pathways to Water Sector Decarbonization, Carbon Capture
and Utilization, Zhiyong Jason Ren and Krishna Pagilla (Eds.)
Istanbul, and elsewhere went to great lengths to create elaborate aqueduct systems to deliver fresh
water hundreds of miles by gravity alone, and to design wastewater collection systems that would
also remove used water and rainwater by gravity. Even in the earliest days of the industrial era,
steam driven pumping systems and rudimentary aeration systems for activated sludge processes were
recognized as enormous energy consumers.
In the United States, the passage of the Clean Water Act in October of 1972 was immediately
followed by a July 1973 report titled Electrical Power Consumption for Municipal Wastewater
Treatment, which found that electrical power consumption for municipal wastewater treatment was
about 1% of residential energy consumption, and that consumption was expected to double as more
treatment plants were built, with a further increase of 40–50% for tertiary facilities (Smith, 1973).
Forty years later, after the construction of many thousands of new publicly owned treatment facilities,
a large percentage of which are performing tertiary treatment, the Electric Power Research Institute
(EPRI) surveyed the sector and found that municipal wastewater treatment accounted for about
0.8% of total electric demand in the US (drinking water treatment and distribution accounted for an
additional 1%) (EPRI/WRF, 2013). Total energy use in 2012 across public water supply and treatment,
and municipal wastewater treatment in the US consumed 39.2 and 30.2 billion kilowatt hours (29.2
and 30.2 TWh) electricity respectively, approximately 1.85% of total electricity use in the US (EPRI/
WRF, 2013). While electric use in the sector indeed climbed over the decades between these two
reports, a combination of increased energy use in other sectors and energy efficiency efforts in the
water sector keep the overall percentage of electric consumption in the sector relatively low. At the
global level, the International Energy Agency (IEA) estimates that electricity consumption for water
and wastewater services is as high as 4% of world electric demand, in part driven by groundwater
pumping and desalination in areas without reliable freshwater supply (IEA, 2016).
While these numbers may be high, they are still quite small when placed in the context of overall
primary energy use. Primary energy use includes the fuel consumption needed for electrical energy
generation, including the wasted heat in fossil electric generation, transmission losses, and including
energy for industrial heat, space heat, water heat, and fuels for transportation of all types. Saul Griffith,
under contract to the US ARPA-E program, converted the EPRI estimates to primary energy as part of
his ‘Super Sankey’ diagram detailing energy flows across the US. At this scale, energy use is measured
in ‘quads’ or quadrillion BTUs, and the US uses about 100 quads per year for all economic activities.
In this metric, municipal water use accounts for 0.13 quads and wastewater for 0.1 quads (Otherlab,
2018). While the law of large numbers says that even one quarter of a quad is still an amazingly large
amount of energy, it may be difficult to convince policy makers to focus their attention on the water
sector given the scale of the overall energy use of the US.
This puts the water sector in a difficult position if the discussion of decarbonization stays focused on
electricity at the facility level. Drinking water and water resource recovery facilities use massive amounts
of electricity relative to many consumers, with a real and measurable impact on the environment.
In many cities, water and wastewater utilities are the single-largest electricity user (USEPA, 2021b).
However, in the context of state, national, and international energy and decarbonization policy
decisions, the amount of primary energy the sector uses are too small to warrant sector-specific policies.
Policy makers and utility managers need to consider the larger picture. If the goal is economy-
wide decarbonization and environmental protection, the optimal approach may not result in each
and every water resource recovery facility being a net-zero energy producer with its own complex
and capital intensive distributed electric system. We are not suggesting that water utility managers
and policy makers be content to simply sit and wait for the grid to decarbonize. At a minimum, both
policy and operations should continue to emphasize major efforts on energy efficiency, including
breakthrough technologies and approaches that drastically reduce energy use, regardless of what
happens to the grid. Already large urban water and wastewater systems may achieve enough economy
of scale to cost-effectively continue to pursue onsite renewable energy production. However, with
utility-scale renewable energy prices steeply declining to the point where it is now more cost-effective
in many areas to build new solar capacity than operate existing coal-fired power plants (IRENA,
2019), the grid in many countries may be able to deliver carbon-free electricity in sufficient quantities
at sufficiently low prices to decarbonize the sector’s electric use in the foreseeable future.
However, water resource recovery facilities offer other avenues to decarbonize beyond electricity:
organic carbon resources that can be used to displace fossil heating, vehicle fuel, or commodity
chemicals; recoverable nutrients that can displace energy-intensive nitrate fertilizers or limited sources
of phosphate fertilizers; vast sources of recoverable heat that can be used in district heating well
beyond the fenceline of a treatment plant; and recoverable water. These resources, especially when
complemented by similar ‘waste’ streams (also full of recoverable resources) from other industries
like food and beverage producers, agriculture, or even waste heat from data centers, may indeed have
enough value to society to warrant policies that encourage their recovery and reuse.
To fully decarbonize the wastewater sector, accounting for electric use alone is not enough. The
accounting must include N2O emissions and fugitive methane releases, as well as Scope 2 and 3
emissions such as chemical consumption and fuel used to transport sludge for offsite disposal. An
example of this type of accounting conducted by Water UK is in Figure 17.1 below. To address these other
sources of emissions, decision-makers need to seek out partnerships, think beyond onsite electricity
production, and deliver co-benefits. By expanding the scope of the effort, the water sector can play a
meaningful role in decarbonization in areas that go well beyond those impacted by electricity alone.
One example of non-electric decarbonization in the water sector is using the biogenic carbon resource
in biosolids. Current best practice is to use anaerobic digestion to convert about 50% of that carbon
2018/19 baseline
Total gross emissions
Overall
All
Admin ort
All
Process emissions
Drinking water
Grid electricity
Process emissions
Wastewater
Grid electricity
Figure 17.1 Reproduced from Water UK Net Zero 2030 Routemap, published by (Water UK, 2020) with kind permission.
to biogas, which must then go through multiple cleanup steps to be useful as a replacement for fossil
natural gas. Once cleaned, the gas can be used for heating fuel, as fuel for combined heat and power
generation, as compressed natural gas vehicle fuel, or be injected into the natural gas pipeline network
as ‘renewable natural gas,’ with the end use dictating the cleanup processes. The biosolids are often
co-digested with other wet feedstocks such as animal manures or food waste to assist the economics
through additional energy production and ‘tipping fees,’ which are received by digester operators
as payment for accepting outside wastes. Given the current low prices for natural gas, the emission
concerns related to fugitive methane emissions, the fact that only half the resource is converted to
energy, and the high capital and operation costs related to anaerobic digestion, gas cleanup, and onsite
energy generation, many utilities have found that this pathway is simply not cost effective, even with
co-digestion and tipping fee income, and even in regions that have strong policies supporting digestion
and relatively high energy prices, such as Massachusetts (USA).
In a future with ample renewable electricity on the electric grid and concerns about fugitive
emissions, some industry observers have suggested that a better path than biogas-specific policy
incentives would be to seek out technologies that convert the carbon to a more valuable end product.
A variety of technologies are currently under development, from arrested methanogenesis to produce
renewable commodity chemical building blocks around which biorefineries can be built (Bhatt et al.,
2020) to hydrothermal liquefaction (Chen, 2020). It is beyond the scope of this chapter to go into detail
on these technologies, but the most successful policies will be those that encourage the highest and
best use of carbon supplies, regardless of feedstock (biosolids, manure, or food waste) and regardless
of the technology used to recover the energy resource.
On the other hand, many experts feel that given the thousands of existing digesters in the US and the
deep technical expertise the industry has amassed, biogas facilities (to include wastewater digesters,
standalone food waste digesters, manure digesters, and facilities that co-digest some mix of those
feedstocks) deserve their own policy carve outs. Recent research from the Water Research Foundation
reveals an increasing focus on biogas policy and regulations, and, whereas for solar there are multiple
market entries, the wastewater sector represents a significant portion of the potential biogas energy
market (Kenway et al., 2019). Biogas resource recovery projects are more likely to progress given
familiarity with the technology. Further, the management of sector specific resources could not only
benefit energy production but could reduce carbon throughout the system. These are reasonable
policy differences, and it is possible that the correct answer is simply to accommodate both sides
with incentives for both biogas and developing new technologies such as arrested methanogenesis or
hydrothermal liquefaction. In the big picture, what really matters is that these wet carbon feedstocks
are captured and used to offset fossil fuels.
17.2.2 Concept 2: There is no overarching policy that mandates decarbonization in the water
sector globally
Policymakers and water industry professionals have often struggled to develop and respond to policies
that encourage decarbonization or its proxy, reduced fossil energy use. This is not for a lack of effort.
In 2019, the Water Research Foundation investigated the ‘Opportunities and Barriers for Renewable
and Distributed Energy Resource Development at Drinking Water and Wastewater Utilities.’ As the
intricate chart in Figure 17.2 demonstrates, in the US alone there are dozens of policies that seek
to impact the water sector’s energy use in some fashion. However, far from spurring widespread
industry adoption of low-cost on-site renewable energy generation at water and wastewater facilities,
the resulting policy matrix is bewilderingly complex and indeed illegible at a normal printed scale.
In Figure 17.3, a zoomed in section of the policy web (highlighted in the red box above) further
illustrates the complex policy landscape.
Faced with such overwhelming complexity in the energy policy landscape, which is layered on top
of the regulatory mandates to protect public health and the environment, and local policies that aim
to keep water/sewer rates low for a variety of reasons, it is unsurprising that the pace of adoption of
Political Heat Recovery Algae Farming in PSC - Public Service Company of Colorado (Xcel Energy)
Motivators from Sewer lines Wastewater and PUCT - Public Utilities Commission of Texas
Geo- Extract Oil in PUCN - Public Utilities Commission of Nevada
or Treatment
thermal RE - Renewable Energy
Anci l
Ancillary
A Sludge In-line turbines in PV Solar Stabilization Facility
Demand Energy ave Energy onsite/near Co-digestion Solar Panels - Solar Thermal
Services Combustion - pressure-controlled Ponds; Solar REG - Renewable Energy Goal
Response Converters Digestion onsite/near facility
facility includes CHP water supply system REP - Renewable Energy Plan
RES - Renewable Energy Standard
RPS - Renewables Portfolio Standard
RRP - Renewable Reward Program
Algae Biofuel RRCEO - Renewable, Recycled and Conserved Energy Objective
SBC - San Bernardino Country
SEDS - Solar Energy Development Standards
Energy Storage Tidal/Wave Wind Heat Recovery SGIP - Self-Generating Incentive Program
SGP - Smart Grid Project
Hydropower
Biogas Solar
Wind Solar Energy
Production Stabilization
Tidal/Wave Ponds; Solar Storage
Heat
Recovery
Hydropower Solar
Wind Biogas
Tidal/Wave Production Heat Recovery
Energy Storage
In-line turbines in
Sludge Windmills - Solar Panels -
pressure- Ancillary Energy Demand
Heat Recovery Wave Energy Co-digestion Combustion - Anaerobic onsite/near onsite/near Solar Floating
controlled water Services Response
from Sewer lines Solar Stabilization Geo- Converters Digestion facility facility Thermal Solar PV
In-line turbines in Co- Sludge supply system
Windmills - or Treatment Anaerobic Solar Panels - Ponds; Solar thermal
Energy pressure-controlled digestion Combustion - Solar Floating
Demand onsite/near Facility onsite/near
Wave Ancillary Geo- Solar PV
Response facility
Energy Services Heat Recovery from
Converters Sewer lines or
Treatment Facility
Algae Farming
in Wastewater
Algae Farming in
and Extract Oil
Wastewater and
Extract Oil in
Objectives National
Local/Regional
Decarbonization policies and water sector opportunities
Figure 17.2 Visual Overview of Regulatory Drivers for Distributed Energy Resource Recovery and Renewable Energy applicable to water and
wastewater utilities, http://dx.doi.org/10.13140/RG.2.2.24472.01286. Reprinted with permission (Kenway et al., 2019).
307
by guest
308
onsite distributed renewable energy in the water sector remains stubbornly low despite a decades-long
focus on energy use from policy makers and water sector NGOs.
This overlapping array of national, state, and local policies, each of which was created with good
intentions, makes it difficult for any single technology or technique to win widespread adoption. This
system requires each utility to go through the laborious and expensive process of tailored research
and technology analyses and trying to create tailor-made systems that fit their unique mix of policies
and other drivers. As a result, only the most motivated utilities have been able to achieve net zero
energy, and almost none have become the net energy producers that engineers and scientists agree
they could be. Even those that have become net zero, like the Gresham, Oregon, USA, facility, still
rely on additional inputs of energy such as solar and imported organic waste (Modern Power Systems,
2019).
Policies can be technology forcing, like regulations mandating numerical pollutant limits. For
example, in the UK, the Climate Change Act 2008 set legally binding limits on the total amount
of greenhouse gas emissions the nation can emit for a given five-year period (UK Department for
Business, Energy & Industrial Strategy, 2021). They can create market forces that drive efficiency,
such as the global carbon tax proposed by Canadian Prime Minister Justin Trudeau at COP26 in
Glasgow (Tasker, 2021). They can be funding driven, such as the Water Security Grand Challenge
from the US Department of Energy (US DOE, 2021), or voluntary and largely unfunded, like the
US Environmental Protection Agency’s Water Reuse Action Plan (US EPA, 2021a). Or they can
combine regulation and funding, like the Massachusetts food waste ban and accompanying funding
for development of anaerobic digesters (Massachusetts DEP, 2021).
There are several important takeaway messages from this complex policy matrix. One is simply
that policies, whether they be national, regional, state/provincial, or local, are not uniform and can
sometimes be contradictory. Another is that there are a variety of forms that policies can take. These
include policies that energize creativity by creating opportunities, such as cap-and-trade methods
(and even within cap-and-trade, there are different types of policies), and policies like carbon taxes
that may achieve similar goals through different mechanisms. Each is shown to drive different but
overlapping reasons for renewable energy investments in the sector (Strazzabosco et al., 2020).
As a second example, there is an international focus on addressing systemic racism in the wake of
recent incidents in the US and around the world. While this may seem far removed from decarbonizing
the water sector, it may have very real impacts on water projects. Traditionally, our industry’s facilities,
despite their contributions to public health, have been considered a burden to their host communities
in terms of odors and localized air emissions, truck traffic, and unsightly facilities barricaded from
the public with high concrete walls and chain link fences. Projects that are sited in or near population
centers of racial minorities and/or economically distressed people are coming under increased scrutiny
for their impacts on those populations. One approach to mitigate these negative impacts is the redesign
of water resource recovery facilities to provide public amenities. The Wusong Wastewater Treatment
Plant upgrade in Shanghai, China (Figure 17.4), won an ‘award of merit’ in 2019 from Engineering
News-Record for its incorporation of an indoor botanical garden that does double-duty as both a publicly
accessible park and integral piece of the treatment train that boosts energy efficiency, controls odors,
and reduces the overall physical space needs of the facility (Engineering News Record, 2019). It does not
require a great leap of imagination to think that future treatment plants will be required offer similar
co-benefits to their host populations, especially when those populations are historically oppressed.
Perhaps one of the greatest areas for potential co-benefits regarding the decarbonization of
wastewater comes from capturing the low-grade heat contained in the water. Given the vast quantities
of wastewater running under the streets in most urban centers, recovering even a small portion of
the thermal energy contained in the wastewater using heat exchangers offers the potential for highly
efficient district heating that can offset other, more carbon-intensive, heating sources. Researchers in
the UK modeled four treatment facilities and found that recovering thermal energy for district heating
offered the potential to reduce carbon emissions by 30–110 kg CO2e/yr.pop. (Hawley & Fenner, 2012).
One implementation of this technology is in Vancouver, Canada, at the False Creek Neighborhood
Energy Utility, where the self-funded project ‘eliminates more than 60% of the greenhouse gas pollution
associated with heating buildings.’ (City of Vancouver, n.d.) However, when developers sought to
recreate this approach just a few miles away in Seattle, Washington, USA, they ran into regulatory
hurdles that prevented them from tapping into this resource. In response, the local authorities
developed a standardized approach to permitting wastewater heat recovery projects, with the goal
of providing multiple benefits that include lowering individual building’s carbon emissions, giving
developers another tool to meet stringent energy codes, and attracting a broader range of tenants,
buyers, and investors (Landers, 2021).
Another example of a wastewater heat recovery project with co-benefits beyond decarbonization
comes from the town of Avon, Colorado, USA. Here, local regulators were concerned that effluent
temperatures from the local wastewater treatment facility were raising the temperature of the receiving
water and impairing cold-water fish species. In response to this policy driver, the town incorporated a
small district heat system (Figure 17.5). The town’s largest municipal energy user is the town recreation
center, which includes multiple heated pools and about 3700 square meters of heated space. By using
waste heat from the wastewater to heat the pools, provide building heat, and to provide salt-free snow
melting on town sidewalks, the town reduces effluent temperatures while simultaneously providing
a low-cost heat source to offset fossil fuels used in the town facilities (Strehler et al., 2010). The town
purchases wind power to offset the electric demand of the heat pump, ensuring the system is zero
carbon. (Avon, Colorado, 2021).
As a practical approach to creating projects that achieve multiple policy goals and co-benefits at the
same time, some local governments have embraced the approach of Integrated Resource Management
(IRM), sometimes called Integrated Resource Recovery (IRR). This is an interdisciplinary, cooperative
project management approach that relies heavily on early stakeholder involvement and an iterative
process to continually refine a project, ensuring broad support before it comes up to a vote or similar
public approval process (Thurm, 2016). The government of Singapore has used a variation of this
approach to integrate multiple objectives, resulting in such showcase facilities as the Marina Barrage,
which simultaneously provides 10% of the country’s water needs, alleviates flooding, and provides
public access with over 15 million visits per year for tourism and recreation (Chye, 2018).
Combining several of these themes, we challenge the reader to envision water resource recovery
facilities as a cohesive part of a larger integrated whole. Despite their tremendous public health
benefits, our facilities are perceived as being burdensome on local communities and, indeed, are being
fought off as tools of institutional oppression. In the US state of New Jersey, legislation passed in
Figure 17.5 Avon, CO, wastewater heat recovery schematic. Image courtesy of Jennifer Strehler/CDM Smith.
2020 requires the state to evaluate the environmental and public health impacts of facilities including
sewage treatment plants, sludge incineration facilities, resource recovery facilities, and cogeneration
facilities – all of which are part of our current concept of water resource recovery facilities – on
overburdened communities (State of New Jersey, 2020).
Imagine, instead of a world where our industry’s facilities are considered to be part of a systemic
burden placed on disadvantaged communities, a world in which water resource recovery facilities are
also viewed as community assets, places that are attractive to live next to and which bring up, not
down, their communities by offering beautiful, clean public spaces full of gardens, as in Wusan.
Today, our facilities, despite receiving 5–10 times more chemical and thermal energy in the influent
than is required to clean that influent (Water Research Foundation, n.d.), are vast energy consumers.
Imagine, instead, a world in which these facilities, through a combination of efficiency and energy recovery,
provide low-cost carbon-free power and district heat to their host communities, as in Vancouver, Canada.
Today, the vast majority of facilities simply pass on water cleaned to the bare minimum of regulatory
standards to a receiving water, with no attempt made to recover that water to offset the need for new
potable water upstream of the treatment plant. Imagine, instead, a world in which water resource
recovery facilities treat that water as a resource, as in Singapore.
17.2.4 Concept 4: Water reuse and energy recovery may be more beneficial in a distributed
setting where the water and energy are recovered at/near the point of generation
Increasing energy demand for water systems, increasing energy prices, advancement towards green
energy and greenhouse gas emissions mitigation goals, and the impacts of climate change, are all
driving water and wastewater utilities to seek investment in self-generated energy. Self-generation
of energy may reduce costs, improve system resilience and reliability, and reduce GHG emissions.
Renewable forms of distributed energy available to utilities can include organic matter in wastewater,
hydropower, thermal heat in wastewater, waste heat from converting gas to electricity, solar, and
wind. Drawing on all these forms of energy, wastewater treatment plants can potentially generate
far more energy than their sites require (Kenway et al., 2019). Collectively, the thermal and organic
energy potential contained within wastewater is estimated to be approximately eight times the amount
of treatment energy required by the wastewater sector in 2012 and a large portion of this potential
remains untapped (Kenway et al., 2019). Further, as utilities can store energy (e.g. as gas or water
at elevation), and can often shift their own energy demands through time, they are an important
potential element of a more renewables-based future grid.
Distributed resources and renewable energy technologies have evolved significantly over the past
ten years with overall distributed energy resource (DER) capacity expected to continue to grow due
to innovation lowering costs, along with continuation of federal and state subsidies. This trend is seen
particularly with solar and wind renewable technologies, which are becoming more competitive with
conventional technologies. For municipal and private water and wastewater utilities there is a unique
opportunity to invest in localized resource recovery.
Capacity and generation for DER technologies have been growing steadily over the past ten years.
For example, total renewable energy generation has grown from 360 TWh in 2000 to 844 TWh in
2020 in the US (USEIA, 2021). Capacity of renewables increased over 300% to 265 GW between 2000
and 2020. Detailed statistics on the distributed and behind the meter fraction of these total numbers
are generally not readily available. However, distributed and small-scale systems generally contribute
less than 1% of generation or capacity. During peak periods, distributed systems contribute more. For
example, in 2017, five classes of ‘behind the meter’ DER’s contributed 44 GW, approximately 6% of the
total US summer peak demand of 769 GW (St. John, 2018). Distributed solar and small-scale combined
heat and power (under 50 MW) contributed nearly 80% of the influence. Smart thermostats, electric
vehicles and distributed energy storage contributed the balance of DER influence.
Energy generation in the water sector globally is dominated by biogas-based technologies. In the
UK the water sector provided 8.5% from renewable sources, with 80% from biogas from anaerobic
digestion (Howe, 2009). In the US there are over 14,500 publicly operated wastewater treatment
plants (WWTP) treating an average flow of approximately 32,345 million gallons per day (MGD)
(Shen et al., 2015) and, of those, 1027 have a capacity above 5 MGD and treat 80% of the wastewater
generated. According to Tarallo et al. (2015), about 851 trillion British thermal units (BTU) of energy
is contained within the wastewater of these 1027 WWTP annually.
Despite a considerable amount of thermal energy carried in the wastewater (2800 megajoules of
waste heat per person annually released to the sewer (Larsen et al., 2016), very little is currently
utilized. State-level analysis by the University of Queensland (Hivert, 2019) confirmed a total potential
of approximately 200,000 GWh/y similar to that estimated by Tarallo (2014). Chemical energy in the
form of biogas has been successfully recovered for many years. Biogas is mostly recovered from the
process of anaerobic digestion of the sewage sludge, even if there are examples of biogas recovery
directly from the wastewater stream (Degarie et al., 2000). Biogas generation from sewage sludge can
be enhanced by combining the sludge with an external organic feedstock, this is called co-digestion.
There are now a range of international examples of water and wastewater utilities successfully
implementing DER projects using on-site digester gas (biogas), solar-PV, hydro, wind turbines and
other renewable sources. Water and wastewater utilities are often good candidates for DER as they
can own large amounts of contiguous land, have high (and movable) energy demand, and can provide
other types of ancillary grid services. Utilities are also participating in demand response programs
by using emergency generators and other energy sources to offset peak grid electricity demands.
However, identifying key options, and navigating regulatory requirements, tariff structures, dynamic
policy positions, and workforce capacity have been great challenges to invest with certainty.
Lacking any unifying policy for decarbonization and renewable energy adoption in the water sector,
adoption has been driven by a multitude of factors from financial opportunities to broad climate
mitigation goals. Strazzabosco et al. (2020) investigated drivers in renewable energy adoption in the
Australian water industry and found reducing energy costs as the most significant factors influencing
renewable energy projects. However, the study notes that compulsory greenhouse gas emissions
reduction requirements as the most influential policy supporting renewable energy projects. Similar to
findings in the US (Kenway et al., 2019), the Australian study questioned the relevance of government
financial policy or renewable energy markets as being influential for the water industry, suggesting a
broader role for the water industry in decarbonization looking at district level opportunities.
One example of this role is in how Metro Vancouver in Vancouver, Canada is directly injecting
cleaned, excess, biomethane from wastewater treatment plants into existing natural gas distribution
systems in response to British Columbia’s climate action. Discussions about a biogas upgrade project
was surrounded by the opportunities to use all elements of liquid waste as resources to reduce Metro
Vancouver’s corporate greenhouse gas (GHG) emissions, and the region’s GHG emissions (Kenway
et al., 2019). Metro Vancouver had signed the provincial government’s Climate Action Charter in
2007, along with almost all other local governments in British Columbia, which committed Metro
Vancouver to be carbon neutral by the 2012 reporting year. The Lulu Island Wastewater Treatment
plant (LIWWTP) Green Biomethane Project in Vancouver, British Columbia is a collaboration
between Metro Vancouver and natural gas provider FortisBC. The objective of this collaboration is to
clean unused biogas to pipeline quality, allowing it to be sold to FortisBC as renewable biomethane.
Renewable natural gas (or biomethane) is produced from biogas as a byproduct of anaerobic digestion
from the LIWWTP. Routinely, biomethane gas is used to provide heat to the plant’s buildings and
digesters and excess gas is safely flared into the atmosphere. From 2014 to 2016 the total energy
demand and biomethane used by the LIWWTP was about 216 220 and 84 640 GJ, respectively, for all
three years and the percentage of heat demand met by biomethane was about 40% for all three years.
An interconnection agreement was established between FortisBC and Metro Vancouver to sell excess
biomethane from the LIWWTP.
Another example is the management of organics. By routing urban organic waste to regional
wastewater treatment plants, the water sector can reduce methane emissions at solid waste facilities
and also improve co-digestion and electricity production. For example, the New York City Department
of Environmental Protection (NYCDEP) and Waste Management New York have a collaboration
to source separate organic waste and preprocessed food scraps in order to improve co-digestion
processes at Brooklyn’s Newtown Creek Wastewater Treatment Plant. The program is a result of a
Mayor Bloomberg’s PlaNYC initiative to make New York City to be the most sustainable city in the
world. The PlaNYC set goals for an 80% reduction in GHG emissions by 2050, energy-neutral in-city
wastewater treatment operations by 2050, maximizing beneficial use while minimizing fugitive
emissions of biogas by 2050, and reaching zero waste to landfill by 2030 (City of New York, 2021).
In the collaboration, NYCDEP reuses the biogas produced from anaerobic digestion in 13 of its
14 treatment plants. The gas is most often utilized in on-site boilers for heating or for powering
equipment. At the Newtown Creek facility, 250 tons of source separated food waste (collected
and processed by Waste Management) is injected into the wastewater treatment plant digesters to
augment biogas production.
While there are many challenges with DER development, three stand out: (i) integration into
the grid; (ii) a need for location-specific knowledge; and (iii) integrated planning. Metering and
grid interconnection procedures are shaping the future of the electric grid (IREC/VSI, 2014). Lack
of coordination in planning and deployment of DER, as well as lack of adequate management
systems, will increase the cost of infrastructure upgrades and reduce the full value of DER as
experienced in Germany (EPRI, 2014). An integrated grid and focusing on co-benefits can enable
higher penetration of DER, reduce voltage loss and environmental impact, defer capacity upgrade,
engage in demand management programs, and improve power system resiliency (EPRI, 2014).
However, water and wastewater utilities find barriers emerge as traditional energy utilities can be
challenged by DER as a new option to traditional energy utility business models (IREC/VSI, 2014;
Willis et al., 2012, 2015).
Key policy and regulatory changes to support DER include minimizing the regulations that small/
medium size WWTPs must meet or look for ways to promote cooperation amongst wastewater
treatment districts. For example, not all plants would need a digester if there were more interaction
around plants. Larger treatment facilities could more actively source organics from smaller plants
or other sources for co-digestion. Wastewater utilities could also establish a closer and mutually
beneficial relationship with municipalities and other food producers or look into developing a campus
environment.
Campus environments would make it easier to manage resource recovery opportunities. For
example, WWTP, city hall, landfills, and council properties can be considered all part of the same
campus. This would allow for surplus energy produced to be shared on a ‘campus grid or micro-grid.’
Increasingly, water and wastewater infrastructure are considered critical and therefore given license
to operate micro-grids as part of a resiliency policy. Such an approach could be further enhanced
by taking a ‘whole-community planning’ view and work with federal, regional, and local planning
agencies, who are more often dependent on local utility services.
Overall, there should be greater organizational support for ‘testing-out’ DER options before
energy and cost crisis drives DER development. Pilot projects could lead to incorporating energy
consumption and DER opportunity into planning of plant designs, help establish connections with
planning divisions to explore energy reduction opportunities as part of infrastructure and asset
management planning (i.e. upgrade and rehab of pump stations, system configurations, incorporate
distributed energy such as micro-turbine installations or energy saving techniques, etc.). Investment in
personnel, able to work with renewable energy within water or wastewater systems is needed.
Looking forward, water utilities entering the DER market need to consider the market forces,
the changing technologies, the changing subsidies, and the changing regulation, but they must also
consider others’ interests to direct and modify market change. The financial returns on a DER project
installed today are likely to change over its operating life. Some jurisdictions will continue subsidies
and regulatory policies that support DER. Others will limit the economic advantages of DER.
17.2.5 Concept 5: Change is constant. We need water policy and technological platforms that
can more easily adapt to change
Many of the other chapters in this book discuss innovation and efficiency. They look at a wide variety
of technologies that could help us get from where the industry is now to some idealized world where
we have decarbonized water in the future. The authors of this book, with many hundreds of combined
years of experience in academia, government, engineering, and facility management, have come
together to try to envision how we can possibly change our industry’s entire way of doing business.
We do this because, even though the wastewater sector has provided the greatest public health service
in human history, we have done so without regard to our part in a growing environmental catastrophe
that threatens to negate the work we have done in the last two centuries. And now, we are pivoting to
become a part of the solution to this new public health threat.
However, we are constrained by a system that never envisioned this kind of change. Our rules
and regulations, our engineering approaches and technologies, our funding mechanisms, our
methods of communicating to the public we serve, indeed every facet of our system, were all
designed under a set of assumptions that was based in an industrial mindset. Clean water would
be made dirty. Dirty water would be collected. And our industry would then make the water clean
again, using as much energy and as many chemicals and as much manpower as needed, with little
regard to the cost.
To reduce waste and inefficiency, we pursued economies of scale, investing huge amounts in
industrialized water treatment facilities designed to last for decades, with the assumption that not
much about the way we made the water dirty would change, that populations would either stay stable
or grow, that we only needed to make the water clean enough to swim in or fish in and that we knew,
by and large, what level of cleanliness that was. These industrial facilities were sited in places that
were convenient for designers, and it was tacitly assumed that they would, as all industrial facilities
were assumed to do, make life a little unpleasant for the neighbors.
Most of this set of assumptions came from the environmental movement of the late 1960s and
early 1970s, which resulted in the creation of public agencies charged with environmental protection
in nations around the world. Those agencies then created water quality standards and guidelines
for treatment plants. And around the world, these assumptions – that our facilities should be large,
expensive, and built to last; that they would have access to unlimited energy and chemical resources;
that they should discharge treated water without further reuse; that they need not necessarily be good
neighbors – have proven surprisingly durable.
In fact, these assumptions were outdated almost as soon as they were implemented. Before the
industrialized world had even finished designing and building its first fleet of wastewater treatment
plants, we were already faced with challenges. Our initial policies assumed that pollution came from
a pipe, overlooking the contribution that stormwater plays. Our policies assumed that we need only
treat for a handful of pollutants, overlooking first the nutrients nitrogen and phosphorus, then failing
to anticipate the dangers of personal care products and pharmaceuticals, and most recently being
caught off guard by the presence of microplastics and per- and polyfluoroalkyl substances (PFAS) that
are problematic in even the most minute quantities. And our founding policies made the disastrous
assumption that we could simply tap into a limitless supply of fossil fuels for energy and chemicals
forever, with no adverse effects.
So, just as the world faces a series of cascading crises – attacks on science and democracy, the
existential threat of climate change, systemic oppression, widening inequality, and a global pandemic
that has killed millions of people and disrupted economies around the world, to name but a few –
the wastewater treatment industry also faces its own set of cascading crises. We are being asked to
simultaneously:maintain aging infrastructure;
• adapt to a changing climate that threatens to disrupt our industry with a biblical series of floods,
droughts, and sea level rise;
• prepare for the eventuality of water reuse;
• rapidly build new facilities in areas where population is growing and maintain older, now
oversized facilities in areas where populations have shrunk;
• decarbonize our sector, one of the most energy-intensive of all public services;
• and to do all of the above in an equitable way that does not add to the historic burden placed
on communities of color, indigenous peoples, and poor people in countries around the world.
This is no small set of changes from what was originally envisioned in the 1960s, and yet it is not
anywhere near a comprehensive list. It does not account for the massive migration we can expect if
even mid-range climate predictions come true (Lustgarden, 2020). It does not account for the desperate
need for our industry to finally develop some version of water and sanitation systems that work well
for the developing world with the resources available to them. It does not account for the impact
future pandemics, which may be waterborne, could have on our systems, and it does not account for
the unknowns that none of us have yet contemplated.
The intention here is not to depress the reader with an insurmountable list of things that are
going wrong. Human beings are endlessly inventive and innovative. In barely a decade, we have gone
from having essentially no electric cars on the market to one in which over 50% of the new car sales
in Norway are now electric (The Guardian, 2021). In that same period, solar PV has gone from an
expensive indulgence for the rich to ‘the cheapest electricity in history’ (Evans & Gabbatiss, 2020). If
the automotive and electrical industries, both of which dwarf the wastewater industry, can pivot and
adapt that quickly, there is no reason to think that we cannot.
However, to do so will require us to adopt policies that anticipate constant change. We need policies
that allow for flexibility and innovation, in an industry that is famously reluctant to innovate. In 2013,
a group of researchers proclaimed that ‘there is an innovation deficit in urban water management.’
After examining the reasons for this deficit, they arrived at the inescapable conclusion that ‘(t)o solve
current urban water infrastructure challenges, technology-focused researchers need to recognize
the intertwined nature of technologies and institutions and the social systems that control change’
(Kiparsky et al., 2013).
In other words, technology alone will not get us out of this mess. We need policies and institutions
that are willing to support us as we innovate and iterate, and that means challenging our current set
of assumptions. Our policies need to be supportive of a broad range of technologies, not pick chosen
winners and losers. Facilities may not need to last 40 or more years. They may not need to be massively
scaled to be efficient. They must not be bad neighbors. They must be able to be upgraded or replaced
easily and cost-effectively as we find new threats to our environment and public health in our used
water.
electric grid. Further, to fully decarbonize the wastewater sector, accounting for electric use alone
is not enough. The accounting must include N2O emissions and fugitive methane releases, as well as
Scope 2 and 3 emissions such as chemical consumption and fuel used to transport sludge for offsite
disposal. To address these other sources of emissions, decision-makers need to seek out partnerships
and deliver co-benefits. These include using the carbon and thermal energy resources contained in
wastewater to offset fossil energy use in other sectors. We encourage policy makers to seek out policies
that are technology neutral and that foster breakthrough innovation.
The second concept discusses the lack of an overarching policy mandating water sector
decarbonization. The water sector has broad policy mandates to protect human health and the
environment through the provision of safe drinking water and sewage treatment. However, these
policies did not anticipate for the interplay of the energy used to provide these crucial services with
the impacts of the energy used to do so. In practice, many efforts to improve public and environmental
health in the water sector come at the expense of climate through increased energy and chemical use
or by increasing emissions like N2O. In their current form, policies around water sector energy use
and decarbonization are a bewildering and complex web of sometimes contradictory local, state/
regional, national, and even international policies. This is a natural follow-on to the first concept, as
many of these policies were not designed with the water sector in mind. Although we do not offer any
specific policy suggestions in this section, we are optimistic that efforts discussed in the other sections
will eventually allow for a clear path to decarbonizing this sector.
The third concept area covers the idea of co-benefits with other policy areas. Put simply, projects
that can meet multiple goals will be more likely to move forward than projects that do not. Here, we
challenge the reader to envision water resource recovery facilities as a cohesive part of a larger integrated
whole. We offer examples of co-benefits including water efficiency projects that also deliver large energy
savings, water resource recovery facilities designed to serve as public botanical gardens and recreation
facilities that are true community assets, and facilities that provide district heat for their neighbors. This
concept area lends itself more to responses to existing policies than to the development of new ones.
The fourth concept area examines the potential to deploy distributed energy resources that take
advantage of the unique attributes of water resource recovery facilities. These include receiving carbon
and thermal energy resources in the wastewater influent, energy intensive processes that could receive
renewable energy ‘behind the meter’ from onsite generation, and the possibility to integrate with other
public sector services in a campus-like setting to share these resources. Policy suggestions include
removing barriers to implementing these existing technologies.
The final concept area covers the constantly changing demands placed on the water resource
recovery sector and the need for policies that are flexible enough to accommodate these changes.
Here, we point to research showing that there is an innovation deficit in the water sector. We call for a
rethinking of policies that lock the sector into decades-long investments in large, complex, industrial
facilities and instead allow for a more rapid, iterative approach to solving water challenges that will
continue to evolve even after we have achieved decarbonization.
Resource recovery from Water: from concept to standard practice. Editors: Ilje Pikaar, Xia Huang,
Francesco Fatone, Jeremy S. Guest. https://www.sciencedirect.com/journal/water-research/
special-issue/104CRLSTGFT
Mobilizing for a zero carbon America: Jobs, jobs, jobs, and more jobs, A Jobs and Employment Study
Report. Saul Griffith, Sam Calisch, Alex Laskey. Rewiring America. July 29, 2020. https://www.
ourenergypolicy.org/resources/mobilizing-for-a-zero-carbon-america-jobs-jobs-jobs-and-more-jobs/
Heat Pumps Using Waste Water in Goethenburg, Sweden. Case study on www.celsiuscity.eu, Jan
16, 2020. https://celsiuscity.eu/heat-pumps-using-waste-water-in-gothenburg-sweden/
Water UK Net Zero 2030 Routemap: Unlocking a net zero carbon future (online resource provided
by Water UK): https://www.water.org.uk/routemap2030/
Emerging solutions to the water challenges of an urbanizing world. Tove A. Larsen, Sabine
Hoffmann,,Christoph Lüthi, Bernhard Truffer, Max Maurer. https://www.science.org/doi/10.1126/
science.aad8641
Municipal wastewater sludge as a renewable, cost-effective feedstock for transportation biofuels
using hydrothermal liquefaction. Timothy E. Seiple, Richard L. Skaggs, Lauren Fillmore, André
M. Coleman. https://doi.org/10.1016/j.jenvman.2020.110852
REFERENCES
Avon, Colorado (2021). Avon’s Heat Recovery System. Retrieved from Avon, Colorado town website: https://
www.avon.org/926/Avons-Heat-Recovery-System#:∼:text=The%20first%2Dof%2Dits%2D,buildings%20
in%20the%20town%20core
Bhatt A. H., Ren Z. and Tao L. (2020). Value proposition of untapped wet wastes: carboxylic acid production
through anaerobic digestion. iScience, 23(6), 101221, https://doi.org/10.1016/j.isci.2020.101221
Biden J. (2021). President Biden’s Full Inauguration Speech. Retrieved from NY Times: https://www.nytimes.
com/2021/01/20/us/politics/biden-inauguration-speech-transcript.html
Chen W. T. (2020). A perspective on hydrothermal processing of sewage sludge. Current Opinion in Environmental
Science and Health, 14, 63–73. https://doi.org/10.1016/j.coesh.2020.02.008
Chye K. T. (2018). From Resource to Asset: Building a Water-resilient Singapore. Retrieved from Global
Infrastructure Initiative: https://www.globalinfrastructureinitiative.com/article/resource-asset-building-water-
resilient-singapore
City of New York (2021). OneNYC. Retrieved from City of New York: OneNYC: http://onenyc.cityofnewyork.us/
City of Vancouver (n.d.). False Creek Neighbourhood Energy Utility. Retrieved from https://vancouver.ca/home-
property-development/southeast-false-creek-neighbourhood-energy-utility.aspx
Degarie C., Crapper T., Howe B., Burke B. and Mccarthy P. (2000). Floating geomembrane covers for odour
control and biogas collection and utilization in municipal lagoons. Water Science and Technology, 42, 291–
298, https://doi.org/10.2166/wst.2000.0664
Engineering News Record (2019). Award of Merit, Water/Wastewater: Wusong Wastewater Plant Upgrade –
Botanical Garden. Retrieved from Engineering News-Record: https://www.enr.com/articles/47613-award-of-
merit-waterwastewater-wusong-wastewater-plant-upgrade_botanical-garden
EPRI (2014). The Integrated Grid: Realizing the Full Value of Central and Distributed Energy Resources. Electric
Power Research Institute, Palo Alto.
EPRI/WRF (2013). Electricity Use and Management in the Municipal Water Supply and Wastewater Industries.
Electric Power Research Institute, Palo Alto.
Evans S. and Gabbatiss J. (2020). Solar is now ‘cheapest electricity in history’, confirms IEA. Retrieved from Carbon Brief:
https://www.carbonbrief.org/solar-is-now-cheapest-electricity-in-history-confirms-iea#:∼:text=Multiple%20
Authors,-Simon%20EvansJosh&text = The%20world’s%20best%20solar%20power,Agency’s%20World%20
Energy%20Outlook%202020
Hawley C. and Fenner R. (2012). The potential for thermal energy recovery from wastewater treatment works in
southern England. Journal of Water and Climate Change, 3(4), 287–299, https://doi.org/10.2166/wcc.2012.013
Hivert G. (2019). State Analysis of Biogas Opportunities in WWTP. Internal Project Report, University of
Queensland, Brisbane.
Howe A. (2009). Evidence: Renewable Energy Potential for the Water Industry. UK Environment Agency, Bristol.
IEA (2016). Water Energy Nexus, Excerpt From the World Energy Outlook. International Energy Agency, Paris.
IREC/VSI (2014). Freeing the Grid 2006–2014. Interstate Reenwable Energy Council and Vote Solar Initiative,
Oakland.
IRENA (2019). Renewable Power Generation Costs in 2019. International Renewable Energy Agency, Abu Dhabi.
Retrieved from https://www.irena.org/-/media/Files/IRENA/Agency/Publication/2020/Jun/IRENA_
Power_Generation_Costs_2019.pdf
Kenway S., Conrad S., Jawad M., Gledhill J., Bravo R., McCalley J. and Howe C. (2019). Opportunities and Barriers
for Renewable and Distributed Energy Resource Development at Drinking Water and Wastewater Utilities
(No. 4625). Water Research Foundation, Denver.
Kiparsky M., Sedlak D., Thompson J. B. and Truffer B. (2013). The innovation deficit in urban water: The need
for an integrated perspective on institutions, organizations, and technology. Environmental Engineering
Science, 30(8), 395–408, https://doi.org/10.1089/ees.2012.0427
Landers J. (2021). King County, Washington, promotes sewer heat recovery as renewable energy source.
Retrieved from ASCE Source: https://www.asce.org/publications-and-news/civil-engineering-source/
civil-engineering-magazine/article/2021/01/king-county-washington-promotes-sewer-heat-recovery-as-
renewable-energy-source (Accessed 15 February 2022).
Larsen T. A., Hoffmann S., Luthi C., Truffer B. and Maurer M. (2016). Emerging solutions to the water challenges of
an urbanizing world. Science (New York, N.Y.), 352(6288), 928–933, https://doi.org/10.1126/science.aad8641
Lustgarden A. (2020). The Great Climate Migration. Retrieved from The New York Times: https://www.nytimes.
com/interactive/2020/07/23/magazine/climate-migration.html (Accessed 15 February 2022).
Massachusetts DEP (2021). Anaerobic Digestion & Organics Diversion. Retrieved from Massachusetts DEP:
https://www.mass.gov/lists/anaerobic-digestion-organics-diversion (Accessed 15 February 2022).
Modern Power Systems (2019). Achieving Energy Net Zero at the Gresham Wastewater Treatment Plant. Retrieved
from Modern Power Systems: https://www.modernpowersystems.com/features/featureachieving-energy-
net-zero-at-the-gresham-wastewater-treatment-plant-7254009/ (Accessed 15 February 2022).
Otherlab (2018). US Energy Flow Super Sankey. Retrieved from Otherlab: https://www.otherlab.com/blog-posts/
us-energy-flow-super-sankey with additional data specific to the water and wastewater sectors retrieved
from Otherlab’s interactive tool at http://www.departmentof.energy/ (Accessed 15 February 2022).
Shen Y., Linville J. L., Urgun-Demirtas M., Mintz M. M. and Snyder S. W. (2015). An overview of biogas production
and utilization at full-scale wastewater treatment plants. Renewable and Sustainable Energy Reviews, 50,
346–362. https://doi.org/10.1016/j.rser.2015.04.129
Smith R. (1973). Electrical Power Consumption for the Wastewater Treatment Industry. Office of Research and
Monitoring, National Environmental Research Center. US Environmental Protection Agency, Cincinnati.
Spang E., Holguin A. and Loge F. (2017). The estimated impact of california’s urban water conservation mandate
on electricity consumption and greenhouse gas emissions. Environmental Research Letters, 13(1), 1–10.
https://doi.org/10.1088/1748-9326/aa9b89
State of New Jersey (2020). Governor Murphy Signs Historic Environmental Justice Legislation. Retrieved from
Official Site of the State of New Jersey: https://www.nj.gov/governor/news/news/562020/20200918a.shtml
(Accessed 15 February 2022).
St. John J. (2018). Distributed Energy Poised for Explosive Growth on the U.S. Grid. Retrieved from Greentech
Media: https://www.greentechmedia.com/articles/read/distributed-energy-poised-for-explosive-growth-on-
the-us-grid (Accessed 15 February 2022).
Strazzabosco A., Conrad S., Lant P. and Kenway S. (2020). Expert opinion on influential factors driving
renewable energy adoption in the water industry. Renewable Energy, 162, 754–765, https://doi.org/10.1016/j.
renene.2020.08.054
Strehler J., Vanderburgh S., Parry D. and Rynders T. (2010). Colorado community benefits from installing waste
heat recovery system; ASME Paper 90479. Proceedings of ASME 2010 4th International Conference on
Energy Sustainability, ASME, Phoenix, AZ.
Tarallo S. (2014). Utilities of the Future: Energy Findings. Water Environment Research Foundation, Alexandria.
Tarallo S., Shaw A., Kohl P. and Eschborn R. (2015). A Guide to Net-Zero Energy Solutions for Water Resource
Recovery Facilities. Water Environment Research Foundation, Alexandria.
Tasker J. P. (2021). Trudeau calls for global carbon tax at COP26 summit. Retrieved from CBC News: https://www.
cbc.ca/news/politics/trudeau-carbon-tax-global-1.6233936 (Accessed 15 February 2022).
The Guardian (2021). Electric cars rise to record 54% market share in Norway. The Guardian, p. n.a. Retrieved from
https://www.theguardian.com/environment/2021/jan/05/electric-cars-record-market-share-norway#:∼:
text = Batter y%20electric%20vehicles%20(BEVs)%20made,Road%20Federation%20(OF V)%20said
(Accessed 15 February 2022).
Thurm B. (2016). Toward an Integrated Resource Management strategy? Retrieved from Innovative Governance
of Large Urban Systems: https://iglus.org/toward-an-integrated-resource-management-strategy/ (Accessed
15 February 2022).
UK Department for Business, Energy & Industrial Strategy (2021). 2019 UK Greenhouse Gas Emissions, Final
Figures. UK Department for Business, Energy & Industrial Strategy, London.
US DOE (2021). About the Water Security Grand Challenge. Retrieved from Energy.gov: https://www.energy.gov/
water-security-grand-challenge/water-security-grand-challenge (Accessed 15 February 2022).
USEIA (2021). Annual Energy Outlook 2021 with Projections to 2050. US Energy Information Administration,
Washington, DC.
US EPA (2021a). Water Reuse Action Plan. Retrieved from US EPA: https://www.epa.gov/waterreuse/water-
reuse-action-plan (Accessed 15 February 2022).
USEPA (2021b). Energy Efficiency for Water Utilities. Retrieved from United States Environmental Protection
Agency: Sustainable Water Infrastructure: https://www.epa.gov/sustainable-water-infrastructure/energy-
efficiency-water-utilities (Accessed 15 February 2022).
Water Research Foundation (n.d.). Water Research Foundation. Retrieved from Energy Optimization: https://
www.waterrf.org/sites/default/files/file/2020-07/4949-EnergyOptimization.pdf
Water UK (2020). Net Zero 2030 Routemap. Available at: https://www.water.org.uk/routemap2030/ (accessed 12
December 2021)
Willis J., Stone L., Durden K., Beecher N., Hemenway C. and Greenwood R. (2012). Barriers to Biogas for
Renewable Energy. IWA Publishing, London.
Willis J., Andrews N., Stone L., Cantwell J. and Greenwood R. (2015). Identification of Barriers to Energy
Efficiency and Solutions to Promote These Practices. IWA Publishing, London.
Chapter 18
Outlook for the carbon-negative
circular water economy
Glen T. Daigger*
University of Michigan, Ann Arbor, MI, USA
*Correspondence: [email protected]
18.1 INTRODUCTION
The previous chapters of this book have articulated many of the opportunities currently available
to the water sector to decarbonize. Numerous innovators and early adopters are investigating
options, conducting trials, and implementing options to increase energy efficiency, reduce carbon
footprint, and recover resources from the water cycle. They are doing this in the general absence of
regulatory drivers but are doing so based on a combination of practical concerns and organizational
and community values. Reduced reliance on outside supplies for critical resources, such as energy,
provides utilities with practical advantages, such as increased ability to manage and control operating
costs under variable economic conditions. Recovery of products that are valued by customers not
only provides revenue that at least partially off-set costs but also provides greater assurance that
management options for these residuals will continue to be available. If viewed as wastes, rather than
valuable products, opposition to associated management options (such as landfilling) can arise and
threaten the ability to manage these residuals. Utilities and communities also reduce carbon emissions
and recover resources to reduce their environmental footprint, in conformance with their broader
commitment to environmental protection. The knowledge and experienced gained by these innovators
and early adopters is essential to better understand what is possible and which of the available options
may best fit various situations. The success being achieved by these innovators and early adopters also
provides examples and evidence needed by others to subsequently adopt some of the new technologies
and approaches being investigated, consistent with the social processes underlying the typical S-curve
of adoption of innovations and new technologies (Rogers, 2003).
The momentum for change within the water sector is certainly accelerating, but there is much
more to do. Long term, the water sector needs to transform to function effectively as we transition
from the current linear economy to a circular one – in fact the water sector can provide leadership for
other sectors providing essential public services. Two questions that may be asked are: (1) will a broad
range of water sector actors adopt new resource recovery technologies and practices, or only a modest
proportion and (2) how can the current transition be accelerated.
© 2022 The Editors. This is an Open Access eBook distributed under the terms of the Creative Commons Attribution Licence
(CC BY-NC-ND 4.0), which permits copying and redistribution for non-commercial purposes with no derivatives, provided the
original work is properly cited (https://creativecommons.org/licenses/by-nc-nd/4.0/). This does not affect the rights licensed or
assigned from any third party in this book. The chapter is from the book Pathways to Water Sector Decarbonization, Carbon Capture
and Utilization, Zhiyong Jason Ren and Krishna Pagilla (Eds.)
Figure 18.1 Progressive development of water-sensitive cities. From Brown et al. (2009).
sector. This is not to discount the role of regulations, when consistently enforced, to promote reliable
service provision. This progression of functions is continuing today, as innovative and early adopter
water sector utilities (water, wastewater, flood control) have transitioned from one stage to the next in
response to a combination of need and opportunity. Broader regulations subsequently followed. This
progression from innovator and early adopters to the establishment of regulations will be discussed
further in a following section.
The water sector has also acted on broader environmental concerns over the years. Water supply
systems have historically been designed and operated to be as energy efficient as possible, for example
incorporating hydropower whenever possible. Energy-efficiency has also been a principal concern
for wastewater treatment historically, and land application of biosolids which beneficially uses the
nutrients and organic matter content of biosolids is a historic and continuing practice in many locations.
Water reuse, both indirect and direct, is a historical practice, with rapidly expanding applications in
many locations. These resource efficiency and recovery actions have been accomplished, of course,
in the context of meeting regulatory requirements. The numerous examples presented in the previous
chapters of this book illustrate current progress addressing global energy and environmental issues by
the water sector. It is imperative at the present time, however, that efforts such as these be dramatically
increased and expanded in scope by the water sector.
Phosphorus, Biochar,
Volume
Electricity, Energy
Heat,
Biogas
Water Water
Figure 18.2 Resource recovery pyramid aligned with the recovery of resources from the water cycle. Adapted from
a diagram presented by Jes LaCour Jansen based on similar diagrams in van der Hoek et al. (2016) and as can be
seen at http://www.betaprocess.eu/the-value-pyramid.php.
in addition to water, include biogas, heat, electricity (produced, e.g., through a combined heat and
power, CHP, system using biogas as fuel), biosolids, and phosphorus (e.g., as struvite). The range of
potential products illustrates, however, the greater diversity of products that can be recovered, along
with their increasing value both monetarily and to society. Missing from this pyramid are inorganic
substances other than the nutrients nitrogen and phosphorus, such as minerals and metals, which
have been demonstrated to offer significant potential value (Westerhoff et al., 2015). Resources can
also be extracted from other segments of the urban water cycle, for example the recovery and reuse of
coagulants used for water treatment.
An aside. One of the reasons that I like the resource recovery pyramid in Figure 18.2 is that it
reminds me of Maslow’s Hierarchy of human needs that is a fundamental principal of the science of
psychology. Maslow’s Hierarchy progresses from basic human needs for survival to those that are
more aspirational like personal esteem and self-actualization. I also see that same progression in the
progressive development of water sensitive cities presented in Figure 18.1, the only difference being that
the vertical progression of Figure 18.2 and the classic presentation of Maslow’s Hierarchy is presented
horizontally in Figure 18.1. One important message that I take from this is that psychology teaches
us that people can generally envision progression from one level of the hierarchy to the next as being
possible, but not jumps of two or more levels. Thus, if one is referring to Maslow’s Hierarchy when
seeking to motivate people to action, one needs to understand where the target audience is at on the
Hierarchy and how the actions one wants them to take can elevate them to the next level. Asking them
to take actions that will elevate them two levels is often not successful. People can view moving up ‘one
step’ to be possible, but not two or more steps. Does this same psychology apply in the water sector?
Do we need to understand where an individual utility is at on the pyramid in Figure 18.2 and focus
on encouraging them to ‘just take the next step’? I suggest that these are good questions to consider.
As shown in previous chapters, many of the present on-going efforts in the water sector focus on
increasing the recovery of carbon in used water through the capture of carbon in the liquid stream and
conversion to biogas through anaerobic digestion. These efforts include mainstream processes, such as
use of anaerobic membrane bioreactors (AnMBR), or through the capture of carbon and stabilization
in anaerobic digesters. As suggested by its lower location on the Figure 18.2 pyramid, this may represent
a high quantity but relatively low-value product compared to others. Moreover, as the electrical grid
transitions to renewable sources such as solar and wind, the production and use of biogas through
a CHP system may represent a decreasing contribution to reducing global environmental impacts.
Economic value may also decline over time as the cost to produce electricity decreases, as is occurring
in numerous locations as the cost for solar and wind energy systems decline. In short, the water sector
will need to ‘climb’ the pyramid illustrated in Figure 18.2 to continue to add value to society.
Factors other than the inherent value of a subject product can significantly influence the valorization
of recovered resources. Consider the case of water. Drinking water produced from ‘natural’ water
resources, such as surface water and groundwater, is often viewed by consumers as the ‘gold standard’
that other ‘water products’ are compared to. Non-potable water products may be accepted by users
for their desired use, but often only at a reduced price compared to potable water, even if potable
water produces no added value. Consideration of the value of water for human consumption is further
complicated by the fact that, because water is a human right, its price rarely reflects the true cost of
producing and distributing it. While alternate methods to secure the right for those for which the
true cost make it unaffordable are available, such as subsidies, they are rarely used. Thus, water may
be one of the lowest priced and least appreciated basic requirements. Then there is the issue of the
acceptance of potable water reuse, whether indirect or direct, even though this water may be of higher
quality than traditional drinking water. Fortunately, this latter situation is changing (although slowly)
as potable reuse is becoming increasingly acceptable to the public. One may wonder to what extent the
water sector may have contributed to the current situation through the historic practice of ‘extolling
the virtue’ of drinking water and not better educating the public about the entire water cycle. The key
point, however, is that influencing perceptions may be as important as the actual practices employed.
being produced from food processing residuals, including single cell protein from wastewater. Efforts
are also underway to use algae to produce a variety of products, not only single cell protein but also
other, higher-value products. These systems may use traditional open ponds, but the development of
photobioreactors is also being pursued. Willy Verstraete and colleagues (Pikaar et al., 2018) have also
proposed approaches that can fully use the nutrients in used water to produce high-quality single
cell protein. Approaches such as these deserve serious consideration and further development when
viewed in the context of the resource recovery pyramid (Figure 18.2).
into the plan as it evolves. Integrated planning and implementation approaches, such as these, need
to be further developed and become the norm. Important learning occurs, not only within individual
utilities and communities, but also by the profession as a whole. Promoting learning across the entire
profession is an essential function of our professional associations.
Time is of the essence, not only because of the need to reduce the environmental footprint of
the water sector but also the nature of the change process. O’Callaghan et al. (2018) have studied
the adoption of new technologies and practices by the water sector, demonstrating that the classic
S-curve of adoption applies (mentioned above, but see Rogers (2003) if background in this model
is needed). They also confirm the long timeframe for adoption by the water sector. Examining
differences in adoption rates of new technologies and innovations, they divide these into needs driven
and value driven categories (O’Callaghan et al., 2019). They find that needs driven technologies and
innovations are generally adopted faster than value driven ones. Unfortunately, resource recovery
and decarbonization changes are often viewed and evaluated as value driven, suggesting the need to
change this paradigm to recognize the urgency with which we need to be reducing our environmental
footprint. The long timeframe historically experienced within the water sector may also not be inherent
but a consequence of past practices and attitudes, as discussed above. Incorporation of evolving
planning and implementation approaches, such as described above, coupled with the implementation
of more flexible and adaptable infrastructure, may relieve some of the factors historically constraining
the rapid adoption of new technologies, approaches, and practices by the water sector.
We also need to more fully understand the processes by which change occurs in the water sector
and incorporate this expanded understanding into our efforts to accelerate change. In this regard, I
refer to work by Rebekah Brown and colleagues who have studied the evolution of water management
practices and the associated technologies, most particularly in Melbourne, Australia (Barron, et al.,
2017; Brodnik and Brown, 2018; Brown, et al., 2013; and Brown, 2005). Figure 18.3 illustrates the
process as a progressive dialogue between those advocating for the new approach (advocating narrative)
and those opposing it (contrasting narrative). This brings to the fore another model of change that I was
exposed to many years ago that was presented to me as the ‘rule of thirds’. This model suggests that
people in general fall into one of three groups in their response to new ideas and concepts. A portion (in
this model, represented by one of the ‘thirds’) like new ideas, another portion is resistant to new ideas
(another ‘third’) and the remainder are uncertain. The model presented in Figure 18.3 is consistent
with the ‘rule of thirds’ and frames it as a dialogue between those who welcome new ideas and concepts
and those who tend to resist new ideas and concepts. The important take-away from the ‘rule of thirds’
is that these two groups, those who welcome and those who resist new ideas and concepts, are the
actors in the resulting dialogue, but the audience is those who are uncertain. Thus, to create change it is
not necessary to convince those who are naturally resistant to new ideas or concepts, but rather those
who are uncertain and who tend to not engage in the dialogue, at least not initially. Of course, this is
not to suggest that membership in one of these three cohorts is good or bad. Rather, it simply reflects
human nature when a sufficient number of people are involved.
Figure 18.3 also illustrates the nature of the dialogue, progressing from the emergence of issue(s), to
identification of solutions and their initial implementation, and finally to the development of policies
and regulations and embedding the results into emerging practice. Table 18.1 further identifies the
principal actors engaged in these dialogues at each stage, highlights the importance of bridging
agents and/or institutions to facilitate the dialogue illustrated in Figure 18.3, and the progression
of knowledge that underpins progress to refine and define the new ideas and concepts. Pilot and
demonstration projects are essential to develop the knowledge needed for the dialogue to progress.
Tools to consolidate the knowledge gained at each stage are also listed. An important outcome of
reviewing this process is that evidence must first be produced to support policymaking, regulation,
and translation into standard practice. This model makes it clear that pilot studies, demonstration
projects, and actual applications are essential to provide the evidence needed for policymaking and
regulations to subsequently occur. There are some who envision the transition process as beginning
with policies and regulations. This is not generally the case, as good policies and regulations
need to be based on evidence, and evidence is often needed to develop the consensus required for
the development and acceptance of policies and regulations. This emphasizes the importance of the
activities already occurring throughout the water profession, as they are essential components of the
change process. We must capitalize on these nascent actions though, to consolidate the progress being
achieved and accelerate its translation into policies, regulations, and standard practice. Again, this
can be an important role for our professional associations.
sanitation (Bos, 2016; Hirano & Latorre, 2020a, 2020b). Society has committed to achieve this right
universally through adoption of water and sanitation as a human right and through expression of the
sustainable development goals.
Table 18.2 Excerpts from the table of contents of The leadership challenge (Kouzes & Posner, 2017).
Good policies and regulations can certainly assist with the necessary on-going transformations. As
discussed in Chapter 17 and illustrated in Figures 18.3 and Table 18.1, however, it must be recognized
that policies and regulations lag behind emerging practice. Evidence and experience are needed to
create the consensus needed for the adoption of policies and regulations, and to form the basis for the
development of constructive policies and regulations. Thus, we must always be ‘pushing the envelope’
to both learn and provide the basis for change. This is a matter of leadership.
My favorite book on leadership is The Leadership Challenge by Kouzes and Posner (2017). I was
first introduced to it (the third edition actually) about 25 years ago and have found its content to be
very useful, both for myself and when I needed to work with others to increase leadership skills. Table
18.2 summarizes the five core practices Kouzes and Posner have found to be the foundation for good
leadership. Kouzes and Posner also emphasize that leadership is not an inherent but rather a learned
skill. Leadership by the water profession is truly the core element of the path forward. As indicated in
Table 18.2, effective leadership boils down to five core practices. The fact that we have so many leaders
who are already practicing this skill makes it clear that the water sector is up to the task. We just need
to keep moving forward and use every opportunity to accelerate the process. Together, we can make
the water sector more sustainable, resilient, and equitable.
REFERENCES
Barron N. J., Kuller M., Yasmin T., Castonguay A. C., Copa V., Duncan-Horner E., Gimelli F. M., Jamali B., Nielsen
J. S., Ng K., Novalia W., Shen P. F., Conn R. J., Brown R. R. and Deletic A. (2017). Towards water sensitive
cities in Asia: an interdisciplinary journey. Water Science and Technology, 76(5), 1150–1157. https://doi.
org/10.2166/wst.2017.287
Bos R. (2016). Manual of the Human Rights to Safe Drinking Water and Sanitation for Practitioners. IWA
Publishing, London.
Brodnik C. and Brown R. (2018). Strategies for developing transformative capacity in urban water management
sectors: the case of Melbourne, Australia. Technological Forecasting & Social Change, 137, 147–159. https://
doi.org/10.1016/j.techfore.2018.07.037
Brown R. R. (2005). Impediments to integrated urban stormwater management: the need for institutional reform.
Environmental Management, 36(3), 455–468. https://doi.org/10.1007/s00267-004-0217-4
Brown R. R., Keath N. and Wong T. H. F. (2009). Urban water management in cities: historical, current and future
regimes. Water Science and Technology, 59(5), 847–855. https://doi.org/10.2166/wst.2009.029
Brown R. R., Farrely M. A. and Loorback D. A. (2013). Actors working the institutions in sustainability transitions:
the case of Melbourne’s stormwater management. Global Environmental Change, 20(2), 287–297.
Daigger G. T. (2011). A practitioner’s perspective on the uses and future developments for wastewater treatment
modelling. Water Science and Technology, 63(3), 516–526. https://doi.org/10.2166/wst.2011.252
Daigger G. T. (2017). Flexibility and adaptability: essential elements of the WRRF of the future. Water Practice &
Technology, 12(1), 156–165. https://doi.org/10.2166/wpt.2017.019
Hilton S. P., Keoleian G. A., Daigger G. T., Zhou B. and Love N. G. (2020). Life cycle assessment of urine diversion
and conversion to fertilizer products at the city scale. Environmental Science & Technology, 55(1), 593–603.
https://doi.org/10.1021/acs.est.0c04195
Hirano M. and Latorre C. (2020a). Guidelines for Public Participation in the Regulation of Urban Water Services.
IWA Publishing, London.
Hirano M. and Latorre C. (2020b). Tools for Public Participation in the Regulation of Urban Water Services. IWA
Publishing, London.
Kouzes J. M. and Posner B. Z. (2017). The Leadership Challenge: How to Make Extraordinary Things Happen in
Organizations, 6th edn. John Wiley & Sons, Hoboken, New Jersey.
Malekpour S., de Haan F. J. and Brown R. R. (2016). A methodology to enable exploratory thinking in
strategic planning. Technological Forecasting & Social Change, 105, 192–202. https://doi.org/10.1016/j.
techfore.2016.01.012
O’Callaghan P., Daigger G., Adapa L. and Buisman C. (2018). Development and application of a model to study
water technology adoption. Water Environment Research, 90, 563–574. https://doi.org/10.2175/1061430
17X15054988926479
O’Callaghan P., Adapa L. and Buisman C. (2019). Analysis of adoption rates of needs driven versus value
driven innovation water technologies. Water Environment Research, 91, 144–156. https://doi.org/10.1002/
wer.1013
Pikaar I., de Vrieze J., Rabaey K., Herrero M., Smith P. and Verstraete W. (2018). Carbon emission avoidance
and capture by producing in-reactor microbial biomass based food, feed and slow release fertilizer:
potentials and limitations. Science of the Total Environment, 644, 1525–1530. https://doi.org/10.1016/j.
scitotenv.2018.07.089
Rogers E. M. (2003). Diffusion of Innovations, 5th edn. Free Press, NY, NY.
Sedlak D. (2014). Water 4.0: The Past, Present, and Future of the World’s Mot Vital Resource. Yale University
Press, New Haven, Connecticut.
van der Hoek J. P., de Fooij H. and Struker A. (2016). Wastewater as a resource: strategies to recover resources from
Amsterdam’s wastewater. Resources, Conservation and Recycling, 113, 53–64. https://doi.org/10.1016/j.
resconrec.2016.05.012
Westerhoff P., Lee S., Yang Y., Gordon G. W., Hristovski K., Halden R. U. and Herckes P. (2015). Characterization,
recovery opportunities, and valuation of metals in municipal sludges from U.S. Wastewater treatment plants
nationwide. Environmental Science & Technology, 49, 9479–9488. https://doi.org/10.1021/es505329q
Page numbers in “f” refer to Figures. Page numbers in “t” refer to Tables.
This book aims to fill an important gap for different stakeholders to gain
knowledge and skills in this area and equip the water community to further
decarbonize the industry and build a carbon-free society and economy. The
book goes beyond technology overviews, rather it aims to provide a system
level blueprint for decarbonization. It can be a reference book and textbook
for graduate students, researchers, practitioners, consultants and policy
makers, and it will provide practical guidance for stakeholders to analyse and
implement decarbonization measures in their professions.
@IWAPublishing