Wio 9781780405025
Wio 9781780405025
Wio 9781780405025
This is an Open Access book distributed under the terms of the Creative Commons
Attribution-NonCommercial-NoDerivatives 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.
Sewage
Treatment Plants
Economic Evaluation of Innovative
Technologies for Energy Efficiency
Edited by
Katerina Stamatelatou and
Konstantinos P. Tsagarakis
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.
Part I
Innovative technologies and economics in sewage
treatment plants – an overview . . . . . . . . . . . . . . . . . . . . . . . . 1
Chapter 1
Reducing the energy demands of wastewater treatment
through energy recovery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.1.1 Wastewater management . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.1.2 Energy demands for wastewater treatment . . . . . . . . . . . 4
1.2 Energy Recovery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.2.1 Use of efficient mechanical parts and sensors . . . . . . . . 7
1.2.2 Anaerobic digestion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.2.3 Fermentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.2.4 Microbial fuel cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.2.5 Energy recovery from sewage sludge . . . . . . . . . . . . . . 10
1.3 Concluding Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.4 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
Chapter 2
The principles of economic evaluation and cost-benefit
analysis implemented in sewage treatment plants . . . . . . . 15
María Molinos-Senante, Nick Hanley,
Francesc Hernández-Sancho and Ramón Sala-Garrido
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.2 Cost Benefit Analysis Methodology . . . . . . . . . . . . . . . . . . . . . . 16
2.2.1 Cost benefit analysis basis . . . . . . . . . . . . . . . . . . . . . . . 16
2.2.2 Internal benefit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.2.3 External benefit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.3 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.4 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
Chapter 3
Introduction to energy management in wastewater
treatment plants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
Catarina Silva, Helena Alegre and Maria João Rosa
3.1 Energy management of wastewater treatment plants
put into context . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.2 Energy management systems: highlights
of the ISO 50001 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
3.3 Energy management and infrastructure asset management . . . 40
3.4 A Framework of Energy Performance Indicators and Indices
for WWTPs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
3.4.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
3.4.2 Energy performance indicators . . . . . . . . . . . . . . . . . . . 43
3.4.3 Energy performance indices . . . . . . . . . . . . . . . . . . . . . 49
3.4.4 Methodology for PAS application . . . . . . . . . . . . . . . . . . 51
3.5 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
Chapter 4
Innovative energy efficient aerobic bioreactors
for sewage treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
Eoin Syron
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
4.2 Aeration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
4.2.1 Innovative process design and improvement . . . . . . . . . 58
4.3 Increasing Oxygen Transfer from a Bubble . . . . . . . . . . . . . . . . 59
4.3.1 Fine bubble diffusers and oxygen transferring
technologies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
4.3.2 Increasing contact time . . . . . . . . . . . . . . . . . . . . . . . . . 61
Chapter 5
Integration of energy efficient processes in carbon
and nutrient removal from sewage . . . . . . . . . . . . . . . . . . . . 71
Simos Malamis, Evina Katsou and Francesco Fatone
5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
5.2 Regulatory Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
5.3 Energy Considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
5.4 Conventional Biological Nutrient Removal Processes . . . . . . . . 74
5.4.1 Description of alternative conventional BNR
processes and configurations . . . . . . . . . . . . . . . . . . . . 74
5.4.2 BNR processes implemented in Europe and Northern
America . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
5.4.3 Energy requirements and cost of conventional
BNR processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
5.5 Innovative Bioprocesses in the Mainstream
and Sidestream . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
5.6 Nitrous Oxide Emissions in BNR . . . . . . . . . . . . . . . . . . . . . . . . 89
5.7 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
5.8 Acknowledgement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
5.9 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
Chapter 6
The aerobic granulation as an alternative to conventional
activated sludge process . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
Anuska Mosquera-Corral, Ángeles Val del Río,
Helena Moralejo-Gárate, Alberto Sánchez,
Ramón Méndez and José Luis Campos
6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
6.2 Basics of Aerobic Granulation . . . . . . . . . . . . . . . . . . . . . . . . . . 96
Chapter 7
Anaerobic digestion of sewage wastewater
and sludge . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
Katerina Stamatelatou
7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
7.2 The Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
7.3 The Technology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
7.4 Anaerobic Digestion of Sewage Sludge . . . . . . . . . . . . . . . . . . 119
7.4.1 Sonication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
7.4.2 Microwave . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
7.4.3 Thermal hydrolysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
7.4.4 Autohydrolyis – Enzymatic hydrolysis . . . . . . . . . . . . . 123
7.4.5 Other methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
7.4.6 Economic analysis of the pretreatment methods . . . . . 124
7.5 Anaerobic Digestion of Sewage . . . . . . . . . . . . . . . . . . . . . . . . 129
7.5.1 Pretreatment of sewage via anaerobic digestion . . . . . 130
7.5.2 Treatment of preconcentrated sewage via anaerobic
digestion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
7.6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
7.7 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
Chapter 8
Resource recovery from sewage sludge . . . . . . . . . . . . . . 139
M. G. Healy, R. Clarke, D. Peyton, E. Cummins,
E. L. Moynihan, A. Martins, P. Béraud and
O. Fenton
8.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
8.2 Defining Trends for Municipal Sludge Treatment . . . . . . . . . . . 140
Chapter 9
Odour abatement technologies in WWTPs: energy
and economic efficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
José M. Estrada, Raquel Lebrero, Guillermo Quijano,
N. J. R. Bart Kraakman and Raúl Muñoz
9.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
9.2 Odour Abatement Technologies . . . . . . . . . . . . . . . . . . . . . . . . 165
9.2.1 Design and economical parameters . . . . . . . . . . . . . . 168
9.3 Comparative Parametric Efficiency
Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
9.3.1 Energy consumption . . . . . . . . . . . . . . . . . . . . . . . . . . 173
9.3.2 Energy efficiency parameter . . . . . . . . . . . . . . . . . . . . 174
9.3.3 Sustainability efficiency parameter . . . . . . . . . . . . . . . 177
9.3.4 Robustness efficiency parameter . . . . . . . . . . . . . . . . . 179
9.3.5 Influence of the H2S concentration . . . . . . . . . . . . . . . . 181
9.3.6 Exploring alternatives to increase technology
efficiency: L/D ratio . . . . . . . . . . . . . . . . . . . . . . . . . . . 182
9.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184
9.5 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
Chapter 10
Instrumentation, monitoring and real-time control
strategies for efficient sewage treatment
plant operation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
Sergio Beltrán, Ion Irizar and Eduardo Ayesa
10.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
10.2 Instrumentation for Monitoring and Control Purposes . . . . . . . 190
10.3 Control of Aeration Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . 193
10.4 Control of Chemical Addition . . . . . . . . . . . . . . . . . . . . . . . . . . 197
10.5 Control of the Internal, External and Sludge Wastage
Flow-Rates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 198
10.5.1 Control of the nitrates internal flow-rate and the
carbon external addition . . . . . . . . . . . . . . . . . . . . . . . . 198
10.5.2 Control of the external flow-rate or sludge
recirculation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 200
10.5.3 Control of the sludge wastage flow-rate . . . . . . . . . . . . 200
10.6 Control of Anaerobic Processes . . . . . . . . . . . . . . . . . . . . . . . . 201
10.6.1 Technological barriers . . . . . . . . . . . . . . . . . . . . . . . . . 202
10.6.2 Applications of control in anaerobic digestion . . . . . . . 202
10.7 Plant-Wide Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205
10.8 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206
10.9 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207
Chapter 11
Microbial Fuel Cells for wastewater treatment . . . . . . . . . . 213
V. B. Oliveira, L. R. C. Marcon, J. Vilas Boas, L. A. Daniel
and A. M. F. R. Pinto
11.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
11.2 Operating Principle of a MFC . . . . . . . . . . . . . . . . . . . . . . . . . . 214
11.3 Fundamentals and Challenges . . . . . . . . . . . . . . . . . . . . . . . . . 215
11.4 Scale Up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
11.5 Operational Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
11.5.1 Effect of pH . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
11.5.2 Effect of temperature . . . . . . . . . . . . . . . . . . . . . . . . . . 223
11.5.3 Organic load . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224
11.5.4 Feed rate and shear stress . . . . . . . . . . . . . . . . . . . . . 225
11.6 Modelling Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 226
11.7 Economic Evaluation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 228
11.8 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230
11.9 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231
11.10 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231
Part II
Innovative technologies and economics in sewage
treatment plants – case studies . . . . . . . . . . . . . . . . . . . . . . 237
Chapter 12
Management optimisation and technologies application:
a right approach to balance energy saving needs and
process goals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239
P. Ragazzo, L. Falletti, N. Chiucchini and G. Serra
12.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239
12.2 Energy Saving with Maintenance and Control Operations . . . . 240
12.2.1 Initial situation of plants . . . . . . . . . . . . . . . . . . . . . . . . 240
12.2.2 Interventions on pumps and piping system . . . . . . . . . 242
12.2.3 Interventions on mixers and engines . . . . . . . . . . . . . . 243
12.2.4 Interventions on air compression and distribution . . . . 244
12.2.5 When energy and process efficiency do not agree . . . 246
12.3 Energy Saving Choosing the Right Technology . . . . . . . . . . . . 247
12.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 248
12.5 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
Chapter 13
Energy factory: the Dutch approach on wastewater
as a source of raw materials and energy . . . . . . . . . . . . . . 251
Ruud M. W. Schemen, Rutger Dijsselhof ,
Ferdinand D. G. Kiestra, Ad. W. A. de Man,
Coert P. Petri, Jan Evert van Veldhoven,
Erwin de Valk and Henry M. van Veldhuizen
Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 251
13.1 Energy Factory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 252
13.1.1 The concept . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 252
13.1.2 The history . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253
13.1.3 The present state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 254
13.1.4 Economic aspects . . . . . . . . . . . . . . . . . . . . . . . . . . . . 255
13.1.5 The future (Wastewater management roadmap
towards 2030) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 256
13.2 Cases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 257
13.2.1 LNG production at ‘s-hertogenbosch . . . . . . . . . . . . . . 257
13.2.2 Thermophilic digestion at STP echten . . . . . . . . . . . . . 262
13.2.3 Delivering biogas from STP olburgen to
potato industry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 263
Chapter 14
A new perspective on energy-efficiency and
cost-effectiveness of sewage treatment plants . . . . . . . . 269
Helmut Rechberger and Alicja Sobańtka
14.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 269
14.2 Methods and Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 271
14.2.1 Application of eSEA for the assessment of
the N-removal performance of STPs . . . . . . . . . . . . . . 271
14.2.2 Data of Austrian STPs . . . . . . . . . . . . . . . . . . . . . . . . . 273
14.3 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 274
14.3.1 Assessment of the N-removal performance of
STPs: eSEA vs N-removal rate . . . . . . . . . . . . . . . . . . 274
14.3.2 Determination of the best practice STP:
energy-efficiency and cost-effectiveness . . . . . . . . . . . 275
14.3.3 The influence of plant size . . . . . . . . . . . . . . . . . . . . . . 276
14.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 279
14.5 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 279
Chapter 15
Techno-economic assessment of sludge dewatering
devices: a practical tool . . . . . . . . . . . . . . . . . . . . . . . . . . . 283
Matteo Papa and Giorgio Bertanza
15.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 283
15.2 Description of the Methodology . . . . . . . . . . . . . . . . . . . . . . . . 284
15.2.1 Operating procedure for test execution . . . . . . . . . . . . 284
15.2.2 Data processing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 288
15.3 Application to a Real Case Study . . . . . . . . . . . . . . . . . . . . . . . 288
15.3.1 Technical issues . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 289
15.3.2 Economic issues . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 290
15.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 292
15.5 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 293
Chapter 16
Short-cut enhanced nutrient removal from anaerobic
supernatants: pilot scale results and full scale
development of the S.C.E.N.A. process . . . . . . . . . . . . . . . 295
Daniele Renzi, Stefano Longo, Nicola Frison, Simos Malamis,
Evina Katsou and Francesco Fatone
16.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 295
16.1.1 Removal or recovery? . . . . . . . . . . . . . . . . . . . . . . . . . 296
16.2 Short-Cut Nitrogen Removal and
Via-Nitrite Enhanced Phosphorus Bioaccumulation:
Fundamentals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 297
16.3 Capital and Operating Cost of Anaerobic Sidestream
Treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 298
16.3.1 Energy consumptions and costs of short-cut nitrogen
removal from anaerobic sidestream . . . . . . . . . . . . . . . 298
16.4 S.C.E.N.A. System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 301
16.4.1 Pilot-scale results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 301
16.4.2 S.C.E.N.A. system integrated in conventional
treatment of sewage sludge . . . . . . . . . . . . . . . . . . . . . 302
16.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 306
16.6 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 307
Chapter 17
Investigation of the potential energy saving in a pilot-
scale sequencing batch reactor . . . . . . . . . . . . . . . . . . . . . 311
Luca Luccarini, Dalila Pulcini, Davide Sottara and
Alessandro Spagni
17.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 311
17.1.1 Sequencing batch reactors . . . . . . . . . . . . . . . . . . . . . 312
17.1.2 Automation of sequencing batch reactors . . . . . . . . . . 313
17.2 Description of the Case Study . . . . . . . . . . . . . . . . . . . . . . . . . 314
17.2.1 Pilot plant . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 314
17.2.2 Process monitoring . . . . . . . . . . . . . . . . . . . . . . . . . . . 315
17.2.3 EDSS architecture . . . . . . . . . . . . . . . . . . . . . . . . . . . . 317
17.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 320
17.3.1 Nitrification time . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 321
17.3.2 Dissolved oxygen consumption . . . . . . . . . . . . . . . . . . 322
17.3.3 Cost analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 323
17.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 323
17.5 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 324
17.6 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 324
Chapter 18
Economic impact of upgrading biogas from anaerobic
digester of sewage sludge to biomethane for public
transportation: Case study of Bekkelaget wastewater
treatment plant in Oslo, Norway . . . . . . . . . . . . . . . . . . . . . 327
Rashid Abdi Elmi and G. Venkatesh
18.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 327
18.2 Wastewater Treatment and Sludge Handling at
Bekkelaget WWTP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 330
18.3 Biogas Handling at Bekkelaget WWTP . . . . . . . . . . . . . . . . . . 331
18.4 The Economics of the Upgrading Facility . . . . . . . . . . . . . . . . . 333
18.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 338
18.6 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 339
Chapter 19
A wind PV hybrid system for power supply of a sewage
treatment plant in a small town in southern Brazil . . . . . . 341
Giuliano Daronco and Alexandre Beluco
19.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 341
19.2 The Sewage Treatment Plant Considered in this Study . . . . . . 342
19.3 Components of the Energy System . . . . . . . . . . . . . . . . . . . . . 345
19.4 Simulations with Homer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 348
19.5 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 349
19.6 Final Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 353
19.7 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 354
Sewage treatment plants (STPs) have been evolved over time to adapt to the growth
of cities, the environmental changes (including climate change), the economic
conditions and, finally, the requirements of society under the influence of both
environment and economy. Initially, the goal of STPs was to simply release
the water of the drains from the pollutants before discharging it back to the
environment. As a result, the STPs were designed on the principle of the activated
sludge process, which is energy consuming and does not take into account the
potential of energy and nutrient recovery. The technological achievements in the
fields of monitoring and control, the design of stable and efficient processes (both
physicochemical and biological), the development of suitable benchmarking and
economic tools have begun to change the philosophy of STP from treatment to
valorisation facilities. This means that, sewage treatment should be incorporated
into a more holistic management scheme, which aims at reducing the pollutants
as well as enhancing nutrient, water and energy recycling in order to maintain the
environment’s integrity in an economic feasible but also efficient way.
In this respect, “Sewage Treatment Plants: Economic Evaluation of Innovative
Technologies for Energy Efficiency” focuses on the novel, energy and/or economic
efficient technologies or modification of the conventional, energy demanding treatment
facilities towards the concept of energy streamlining and their economic impact. The
book brings together knowledge from Engineering, Economics, Utility Management
and Practice and helps to provide a better understanding of the real economic value
with methodologies and practices about innovative energy technologies and policies
in STP. It consists of two parts; the first part is dedicated to critical discussion of
technologies aiming at enhancing the energy efficiency of STP including economic
aspects as well, while the second part includes case studies demonstrating the
economic impact of applying the energy efficient technologies at full scale.
The first two chapters are introductory. The first one briefly overviews novel, but
well established technologies in a STP as well. The second one explains how the
cost benefit analysis methodology can be used to assess the economic feasibility
of a technology or change in the operation of a STP. Chapter three focuses on how
strategic management, when regarding a STP as a whole, may lead to a better
performance at a lower cost (from a total asset life cycle point of view). Chapter four
presents the save in energy in the case of aerobic bioprocesses alternatives to the
conventional activated sludge process and when advanced technologies of oxygen
transfer are applied. More particularly, the nutrient removal technologies in energy
efficient integrated systems are discussed in chapter five, while the promising
aerobic granulation is the subject of chapter six. The application of anaerobic
digestion and recent developments in the field of both sewage and sewage sludge
treatment is presented in chapter seven. Focusing on the sewage sludge not only for
energy but also for nutrient recovery is the subject of chapter eight. Besides liquid
and solid effluents, STP produces gases that affect the atmospheric environment.
In chapter 9, an energetic and economic efficiency analysis of common odour
abatement technologies in STPs is performed. The advances in monitoring and
control boosted the performance and improved the economics of the STP. This is
examined in chapter 10, which also addresses the plant wide control. Although the
Microbial Fuel Cell technology is still technically far from its full scale application,
it deserves attention due to the rapid evolvements in this field (chapter 11).
Chapter 12 is the first case study presented in the second part of the book,
based on the experience of two companies managing integrated water service
in northeastern Italy and focusing on the energy savings in municipal STPs.
Next, the concept of the energy factory for STP is introduced and case studies
of implementing this approach in the Netherlands are presented (chapter 13). In
the case studies of Austrian STPs, the energy consumption and costs are related
to nitrogen removal efficiency and plant size (chapter 14). A methodology for
evaluating sludge dewatering devices is presented in chapter 15 and a case study
example of the implementation of this methodology is given. Chapter 16 proposes
an enhanced nutrient removal process, which is necessary if anaerobic digestion
becomes the core technology in STP, so that the nutrient rich anaerobic effluents
are adequately treated. The subject of chapter 17 is the Sequencing Batch Reactor
(SBR) technology and the potential for energy savings though aeration schemes,
as has been demonstrated in pilot scale studies. Next, the cost impact of changing
the end use of biogas and transform a STP in Norway to an energy supplier of the
public transportation sector is presented. Chapter 19 finalizes the second and last
part of the book with a study that shows how the alternative energy sources can be
integrated into STP to contribute into cost reduction of the plant.
On the completion of this collected volume, we would like to thank the
contributing authors for sharing their experience and perspective of future STPs.
Katerina Stamatelatou
Konstantinos P. Tsagarakis
1.1 Introduction
Wastewater treatment is a significant aspect of water industry that safeguards
public health, natural environment and allow for a high quality of life and
economic development. The rapid population growth in highly urbanized and
industrialized societies has resulted to the production of large volumes of
wastewater, which require energy and cost-intensive treatment to be sanitized
and safely discharge into receiving water bodies. In order to meet discharge
limits, existing wastewater treatment facilities utilize energy-intensive
treatment techniques, although current scientific knowledge can provide the
know-how to achieve energy saving and recovery in treatment plants. This
chapter gives a brief overview of well-established as well as novel technologies
that have the potential to reduce energy demands of existing, typical wastewater
treatment facilities, either by energy recovery or saving during treatment, in
order to reduce the environmental footprint and attain energy efficient treatment
facilities.
1.1.1 Wastewater management
Water and wastewater management are highly important and interdependent tasks
that can strongly affect human well-being and quality of life. If left untreated,
wastewater can pollute surface and ground water reservoirs, thus posing serious
threats onto public health and the environment. Hence, the role of water and
wastewater industry is to provide reliable protection and safely discharge wastewater
into the aquatic environment. However, rapid and localized population growth has
led to large volumes of clean water being consumed daily and respectively large
volumes of wastewater being produced, which stresses even more the existing
wastewater facilities. On top of this a rapid deterioration of the quality of water
reservoirs, mainly due to the increased urbanization, industrialization and farming
activities, is observed. This is evident by the excess of organic pollutants and
nutrients (N and P) loads in aquatic bodies. All the above indicate that more
intensive water and wastewater treatment technologies, which are associated with
high energy demands and costs, need to be adopted to safeguard public health and
the natural environment.
Although estimations vary, on average the daily municipal water use per
capita reaches 400 L in USA (USGS, 2014), while the mean municipal water
consumption in Europe is half, about 200 L (EC, 2012), with substantial
differentiations among EU countries. In developing countries the municipal water
use per capita is substantially lower, reaching an order of magnitude less than
the developed ones (UNDP, 2006). Used water is collected in sewage systems
and then is led to treatment plants, as to be sanitized and safely discharged to
environment and/or recycled for agriculture and other uses. In the UK about
625 × 103 km of sewers are used daily to collect over 11 × 106 m3 of municipal
and industrial wastewater (DEFRA, 2012). These vast quantities should be
treated before ending up to receiving water bodies, but that is not always the case.
For example, in Abbey Mills Pumping Stations, London around 16 × 106 t of raw
wastewater is annually discharged to the River Lee, ending up to river Thames.
In USA, in 2008 60.41 × 109 m3 of municipal wastewater were produced, of
which 47.2 × 109 m3 were collected and finally only 40.89 × 109 m3 were treated
(FAO, 2014).
Wastewater treatment comprises various physical, chemical and biological
processes, as well as their combination, in order to produce an effluent that can
be safely disposed to environment without causing any short or long term adverse
effects to humans or other living beings. Nonetheless, in order to meet wastewater
discharge permits, high energy demands are required, leading to high operational
costs and making wastewater management unsustainable. Therefore, more efficient
and energy friendly treatment systems, that require lower to zero external amounts
of energy to operate and hence lower operational costs, should be introduced in
large scale.
However, there is no doubt that as our demand for clean water increases, so
does the total amount of energy needed to safely discharge wastewater into the
environment.
For example, over 1010 L of sewage are produced every day in England and
Wales and it takes approximately 6.34 GWh of energy to treat this volume
of sewage, which is almost 1% of the average daily electricity consumption of
England and Wales (POST, 2007). Moreover, Shoener et al. (2014) reported that
current energy-intensive approaches to wastewater treatment, which consume
roughly 0.3–0.6 kWh m−3 (i.e., 3% of U.S. electricity demand), further contribute
to climate change through greenhouse gas emissions from electricity production
(Shoener et al., 2014).
Nevertheless, to accurately estimate the actual energy demands of a wastewater
treatment facility, treatment stages and utilized technologies should be taken into
account. In addition, energy demands are strongly related to the physicochemical
characteristics of sewage (i.e., organic load, total solids, etc.) and the desirable use
of the final effluent (i.e., aquifer recharge, agriculture use, etc.), since these affect
the degree of treatment intensity. Typically, a sewage treatment plant consists of
five main stages, as described below (POST, 2007):
• Pre-treatment: includes bar screens to remove large objects, a flow
equalization tank and a grit removal channel.
• Primary treatment: consists of a primary sedimentation tank where solids
are physically settled out by gravity.
• Secondary treatment: typically is based on an activated sludge system,
where bacteria are used to convert organic pollutants to carbon rich
sludge.
• Tertiary treatment: might include UV irradiation, activated carbon filters
or other advanced techniques to further remove non-biodegradable organic
matter and/or disinfect the water.
• Sludge treatment: usually incineration, or sludge thickening and disposal is
applied.
Table 1.1 presents a typical energy demands’ breakdown for a common
wastewater treatment facility. It is evident that the highest amount of energy,
that is, 55.6%, is consumed in the activated sludge aeration process. The primary
clarifier and sludge pumps is the second largest energy demanding stage, it
consumes 10.3%, followed by heating for digesters (7.1%) and solids dewatering
(7%). All the above stand for about 80% of the total energy demands of a
common treatment facility. Since, as described above, conventional wastewater
treatment processes are energy-intensive and hence not environmentally friendly,
future strategies should focus on reducing energy demands and enabling zero
to negative energy treatment requirements, as to create economic incentives
and enable access to sustainable sanitation in both developed and developing
communities (Shoener et al., 2014).
1.2 Energy Recovery
To take a step towards wastewater treatment facilities that have zero to negative net
energy demands (i.e., energy produced during treatment is greater than the energy
required for their operation), all potential energy saving and energy production
steps in a typical treatment facility should be identified. Figure 1.1 illustrates how
and where within the train system of wastewater treatment, the greatest potential
for energy saving and recovery can be achieved.
Figure 1.1 Processes that have the potential to save and/or recover energy within
a wastewater treatment facility.
pressure sensors, allowed air flowrates optimization, thus avoiding excess of air
sparging consequently reducing its energy demands. The aeration capital cost
for this solution was similar to conventional disc aerators, while energy and air
requirements were reduced by 20% and 33%, respectively.
1.2.2 Anaerobic digestion
Anaerobic digestion (AD) is a well-known natural process where biodegradable
materials are broken down by the action of microorganisms, in the absence of
oxygen, which result to decreased organic loads and simultaneously to the
production of biogas. Biogas is a mixture of gases that mainly consist of methane
(typically 60–65%) and carbon dioxide. The process takes place in sealed anaerobic
digesters under appropriate temperatures of about 30 to 38°C (mesophillic
digestion) or about 49–57°C (thermophillic digestion), with the first being a more
stable process (Reith et al., 2003). AD can be divided into three main steps, (a)
hydrolysis, where microorganisms split the organic matter to simpler forms in the
presence of water, (b) volatile acid fermentation, which include acidogenesis – and
acetogenesis, with end products being acetic acid, carbon dioxide, and hydrogen,
and (c) methane formation, where products from the previous step are converted to
methane and carbon dioxide. AD can take place either in a single stage, digestion is
performed in a single tank at constant temperature, or in multiple stages, different
tanks or different temperatures or both, are used. The latter finds favourable use
in wastewater management since it allows AD facilities to optimize both organic
removal and biogas production.
Wastewaters, as well as the sludge that is generated in the aeration stage, are
rich in organic matter and therefore can be used to produce energy (biogas) and
simultaneously reduce their organic load through AD. Depending on its quality
and quantity, biogas can be used for heating purposes, electricity production or can
be fed into a combined heat and power (CHP) system to provide heating to AD and
power the high energy intensive processes within the treatment plant, such as the
aeration blowers (Cao & Pawlowski, 2012). In general, anaerobic digesters are able
to create enough biogas to maintain their own heating temperature and provide
heat and/or electricity to other stages of the plant and to the building facilities on
site (Caldwell, 2009).
AD is a well-established solution for energy recovery and organic load
reduction, presenting both environmental and economic benefits, while its
application is steadily increasing in wastewater treatment facilities throughout
the world. For example, the volume of biogas captured and utilized in two
Norwegian wastewater treatment facilities rose from 8.1 × 106 m3 in 2000 to
14.6 × 106 m3 in 2007 (Venkatesh & Elmi, 2013). Furthermore, in 2005–2006
the UK water industry generated 493 GWh from AD, while currently, with 110
AD facilities installed, it annually generates approximately 800 GWh through
AD of sewage sludge treatment (Mills et al., 2014; POST, 2007). To add, most
1.2.3 Fermentation
During fermentation specific microorganisms, in the absence of oxygen, follow
a certain metabolic pathway and convert monomers (sugars) to acids. Anaerobic
wastewater digestion to generate methane as a final product is a well-established
technique. Nonetheless, if the growth of methanogenic bacteria is inhibited, thus
preventing methane formation, and only hydrogen producing microorganisms
are left to flourish, then acetogenesis will be the last step of AD, thus generating
hydrogen (H2), acetic acid and CO2 (Reith et al., 2003).
H2 is a high energy density (122 KJ/g) fuel that produces zero CO2 emissions
when burned. Nonetheless the most common H2 generation processes are steam
reforming of natural gas and water electrolysis, which are extremely energy
and cost-intensive (Su et al., 2010; Argun & Kargi, 2011). Therefore, increasing
research interest has been directed towards more sustainable and energy-efficient
techniques for its production. Among them, anaerobic wastewater fermentation has
proven to be a promising process that operates under mild conditions and requires
low energy demands, since it achieves both waste reduction and clean energy
production, namely H2 (Chen et al., 2008).
H2 production through wastewater fermentation can be achieved (a) under the
presence of light (photo-fermentation), where light provides metabolic energy,
(b) under the absence of light (dark-fermentations), where organic compounds
provide metabolic energy, or (c) by a combination of both techniques (combined-
fermentation) (Su et al., 2010; Argun & Kargi, 2011). The latter has been reported
to provide higher H2 yields and can also achieve higher reduction of the effluents’
organic load (Chen et al., 2008).
In photo-fermentation anaerobic photosynthetic bacteria, such as Rhodobacter
and Rhodopseudomonas, catalyze organic acids, such as acetic and butyric acids
and more simple ones, as glucose, fructose and sucrose, while in dark fermentation
anaerobic bacteria, such as Clostridium and Enterobacter, can catalyze glucose,
sucrose, starch and cellulosic materials to produce H2 (Su et al., 2010). Restriction
factors of applying the process at large scale include low hydrogen yields (i.e.,
typically less than 15% of the maximum theoretical potential), high cost and
the need for carbohydrate-rich wastewaters, thus this technology has yet to be
effectively introduced at industrial scale.
content, may contain hazardous substances, such as heavy metals and persistent micro-
pollutants. Sewage sludge contains from 0.25–12% solids by weight, depending on
the wastewater treatment technique that was adopted (Tchobanoglous et al., 2003).
Therefore, sludge management tradition handling routes, such as agricultural use,
can be unsafe, while sludge incineration is associated with high energy demands
and costs and landfill disposal faces various legislation restrictions, for example,
Directive 2000/76/EEC and 2003/33/EEC (Manara & Zabaniotou, 2012).
Alternative management processes include the thermochemical treatment of
sludge in the absence of oxygen or in oxygen-starved environments, as to prevent
combustion. Under carefully controlled conditions and extreme temperatures
(350–1000°C), sludge may undergo chemical reactions to produce fuels that can
be used for heat and/or energy production and simultaneously achieve organic load
removal. Processes include gasification, which produces syngas, and pyrolysis,
which produces bio-oil. These are potential alternatives to sludge incineration, but
similarly operational costs are still high, especially when using high temperatures.
Also, special consideration should be given to the monitoring of operating
conditions to avoid any formation of harmful by-products, such as hydrogen
cyanide (Samolada & Zabaniotou, 2014).
1.2.5.1 Pyrolysis
During pyrolysis sewage sludge is thermally decomposed in an oxygen-free
environment to gases (biogas), liquids (bio-oil) and solids (biochar). The major
product obtained from this process is the bio-oil, which can be used as a fuel, the
same stands for biogas, as well as a source of valuable chemical products. Biochar,
a carbon-rich solid, can be used in various applications ranging from agriculture
to adsorbent material for contaminants in soils, depending on its quality (Agrafioti
et al., 2013).
Pyrolysis, a rather endothermic process (100 kJ kg−1), operates at temperatures
ranging from 350°C to 1000°C. Pyrolysis by-product formation is affected by
the process operating conditions, such as temperature and pressure as well as the
initial sludge characteristics. Therefore, when bio-oil is the target, fast pyrolysis is
employed, during which high heating rates, moderate temperatures (500°C) and
short gas residence times (<2 s) are applied, whilst when biochar is the desired
product, slow pyrolysis, characterized by mild temperatures (350–600°C) and
heating rates, is applied (Leszczynski, 2006).
Pyrolysis is a ‘greener’ technology when compared to incineration, since the
lower operating temperatures applied and the absence of oxygen result to toxic-free
by-products. In contrast, oxygen utilization and high temperatures applied during
incineration process can result to the formation of toxic substances, such as furans
and dioxins. Although research has been focused on pyrolysis of sewage sludge
for bio-oil production, thus recovering energy within the wastewater treatment
facility, large scale applications of the technology is limited. This is due to the
need for relatively complex and expensive equipment and the need for using drying
feedstock (Samolada & Zabaniotou, 2014).
1.2.5.2 Gasification
The gasification process uses heat, pressure and steam to convert carbonaceous
materials, in the presence of oxygen and/or steam, into a synthesis gas called
syngas, which is a mixture of CO, H2 as well as N2 and traces of CO2, CH4 and
other hydrocarbons and slag. Gasification mainly transforms organic materials to
combustible gas or syngas, using between 20% and 40% of the oxygen required
for total combustion, whereas pyrolysis is a thermochemical reaction carried
out at elevated temperatures (500–1000°C) and theoretically in an oxygen-free
environment.
Gasification has the advantage of reducing the volume of sewage sludge and
toxic organic compounds; while simultaneously it generates syngas that can be
used for heat (e.g., syngas from sewage sludge has a heating capacity of about 4 MJ
m−3) or electricity (i.e., in fuel cells) production (Dogru et al., 2002; Judex et al.,
2012). In addition, problems commonly faced in incineration process, like the need
for supplementary fuel and emissions of toxic by-products, such as SOx and NOx,
heavy metals and fly ash, can be avoided by the gasification process.
Limitations of the technology include feedstock characteristics, such as moisture
(>90% dry solids) content, and the complexity of the reactors design, such as design
of the feeding system, mixing and separation of the feedstock. Also, the generated
syngas must be cleaned and purified before its further use and the high cost of the
initial set-up still prevents the wide application of this technology at large-scale.
1.3 Concluding Remarks
Reducing energy demands and increasing energy recovery in wastewater treatment
facilities can be a feasible venture by means of current technological advances. If
existing treatment facilities are upgraded as to achieve lower energy demands and
simultaneously take advantage of energy harvesting techniques from wastewater
and sewage sludge, then positive net energy facilities could exist, that will further
benefit local or national communities by providing the excess heat and energy.
In this chapter the energy demands of a typical sewage treatment plant as well
as options to reduce them were demonstrated. Furthermore, techniques that can
achieve substantial energy recovery, within the various treatment stages, were
presented. It is clear that scientific knowledge and the know-how to create energy,
and thus save cost exist and can lead to the establishment of highly sustainable
sewage treatment plants. These technologies are described and discussed in detail
in this book, while emphasis is given to economic aspects of wastewater treatment
facilities. Finally, successful case studies of energy recovery during wastewater
treatment are demonstrated in Part II of this book.
1.4 References
Agrafioti E., Bouras G., Kalderis D. and Diamadopoulos E. (2013). Biochar production by
sewage sludge pyrolysis. Journal of Analytical and Applied Pyrolysis, 101, 72–78.
Ahn Y., Hatzell M. C., Zhang F. and Logan B. E. (2014). Different electrode configurations
to optimize performance of multi-electrode microbial fuel cells for generating power
or treating domestic wastewater. Journal of Power Sources, 249, 440–445.
Argun H. and Kargi F. (2011). Bio-hydrogen production by different operational modes
of dark and photo-fermentation: an overview. International Journal of Hydrogen
Energy, 36, 7443–7459.
Barua P. K. and Deka D. (2010). Electricity generation from biowaste based microbial fuel
cells. International Journal of Energy, Information and Communications, 1(1), 77–92.
Caldwell P. (2009). Energy efficient sewage treatment can energy positive sewage treatment
works become the standard design? In: Proceedings of the 3rd European Water and
Wastewater Management Conference, 22–23 September.
Cao Y. and Pawłowski A. (2012). Sewage sludge-to-energy approaches based on anaerobic
digestion and pyrolysis: brief overview and energy efficiency assessment. Renewable
and Sustainable Energy Reviews, 16, 1657–1665.
Chen C.-Y., Yang M.-H., Yeh K.-L., Liu C.-H. and Chang J.-S. (2008). Biohydrogen
production using sequential two-stage dark and photo fermentation processes.
International Journal of Hydrogen Energy, 33, 4755–4762.
DEFRA (2012). Waste water treatment in the United Kingdom – 2012. Department for
Environment, Food and Rural Affairs, Nobel House 17, Smith Square London SW1P
3JR.
DEFRA (2013). Anaerobic digestion strategy and action plan. Department for Environment,
Food and Rural Affairs, Nobel House 17, Smith Square London SW1P 3JR.
Dogru M., Midilli A. and Howarth C. R. (2002). Gasification of sewage sludge using a
throated downdraft gasifier and uncertainty analysis. Fuel Processing Technology, 75,
55–82.
Du Z., Li H. and Gu T. (2007). A state of the art review on microbial fuel cells:
a promising technology for wastewater treatment and bioenergy. Biotechnology
Advances, 25, 464–482.
EC (2012). Would you drink your wastewater? European Commission, Directorate-General
for the Environment, Luxembourg: Publications Office of the European Union, ISBN:
978-92-79-22529-1.
FAO (2014). AQUASTAT database, food and agriculture organization of the United Nations
(FAO). Website accessed 12 September 2014.
Judex J. W., Gaiffi M. and Burgbacher H. C. (2012). Gasification of dried sewage sludge:
status of the demonstration and the pilot plant. Waste Management, 32, 719–723.
Leszczynski S. (2006). Pyrolysis of sewage sludge and municipal organic waste. Acta
Metallurgica Slovaca, 12, 257–261.
Manara P. and Zabaniotou A. (2012). Towards sewage sludge based biofuels via
thermochemical conversion – a review. Renewable and Sustainable Energy Reviews,
16, 2566– 2582.
Mills N., Pearce P., Farrow J., Thorpe R. B. and Kirkby N. F. (2014). Environmental
& economic life cycle assessment of current & future sewage sludge to energy
technologies. Waste Management, 34, 185–195.
2.1 Introduction
Traditional wastewater treatment technologies, most of them based on activated
sludge, have been widely implemented in the last decades over the world (Gavasci
et al. 2010). However, growing public concern over environmental protection and
increasing energy costs have led to the development of innovative technologies for
energy saving. Improving energy efficiency is a challenge that should be taken into
account in the construction of new wastewater treatment plants (WWTPs), in the
renovation of the plants and in the operation of all facilities.
The development and implementation of innovative technologies for energy
efficiency involves costs and benefits that should be assessed. Economic feasibility
studies are an essential tool in the decision making process for the implementation
of new technologies alternatives in the field of wastewater treatment (Molinos-
Senante et al. 2012).
One of the most popular tools to assess the economic feasibility of any project is
cost-benefit analysis (CBA) since it ensures the economic rationality of investments
testing whether the benefits of action outweigh the costs. The approach followed
in the performance of CBA in the evaluation of projects has been modified taken
into account the objectives of the development policies. There are three stages
(Molinos-Senante et al. 2010):
(1) Traditional approach: it is a financial analysis based on the comparison of
incomes and costs generated during the life of the project, that is, what are
known as internal or private impacts. It follows a clear economic approach
aimed to increase the level of welfare in monetary terms, typically defined
as profits.
(2) Socio-economic approach: this arises when the concept of social equity
is incorporated. The aim is to achieve equitable income distribution, or at
least to include some kind of income-related weights into the calculation of
benefits and costs to different groups.
(3) CBA involving environmental externalities valuation: It results from the
incorporation of environmental criteria in the decision-making process.
This type of CBA originated in the 1980s and become more widespread
in the 1990s (Pearce & Nash, 1981; Sudgen & Williams, 1988; Hanley &
Spash, 1993; among others).
Wastewater treatment in general and innovative wastewater treatment
technologies for energy efficiency in particular have important associated
environmental benefits which are defined in economic terms as positive
externalities. Hence, the assessment of the economic feasibility of wastewater
treatment processes must be carried out through CBA instead of financial analysis.
Otherwise, the environmental benefits of cutting pollution and to reduce energy
consumption and consequently greenhouse gas (GHG) emissions would be
underestimated since they are not accounted by the market (unless governments
or the market offers payments for reductions in carbon emissions, for example
through carbon trading).
Other reasons for selecting CBA as the preferred method to assess the economic
feasibility are that: (i) it allows planners and decision-makers to take a long-term
view of the project lifetime; (ii) it provides a project ranking, which, for all practical
purposes, proves to be quite scientific and satisfactory (Molinos-Senante et al.
2013a) and; (iii) it clearly sets the impacts of a project in terms of who is affected,
by how much, and when (Hanley & Barbier, 2009).
Since CBA starts from the premise that a project should only be commissioned if
all benefits exceed the aggregate costs, the benefits of each proposal are compared
with their costs by using a common analytical methodology (Eq. (2.1)).
NP = B1 + BE (2.1)
where NP is the net profit (total income – total costs); B1 is the total internal benefit
(internal income – internal costs); and BE is total the external benefit (positive
externalities – negative externalities). A project is economically feasible if, and
only if, NP > 0. If the result of the calculation is NP < 0, then the project is not
economically feasible. The best option offers the highest net profit (Benedetti et al.
2006; Chen & Wang, 2009). Moreover, total income can be divided by total costs
to get a ratio which can be used to rank policies/project that are competing for
scarce funds, with the option of having the highest benefit to cost ratio being the
most preferred (De Anguita et al. 2011).
The implementation of an innovative wastewater treatment technology is a
project whose life period is more than one year, and as a result, the internal
and external benefits must be adjusted for the time they will occur. For this
reason the NP must be expressed in present value terms. By means of a properly
chosen discount rate the investor becomes indifferent regarding cash amounts
receiving at different points of time. The net present value is calculated as shown
in Eq. (2.2):
T
NPt
NPV = ∑ (1 + r )
t =1
t (2.2)
NPV is the net present value, NPt is the net profit at time t; r is the discount rate
and T is the project lifespan.
NPV results will determine the project’s feasibility. As well as NP, a positive
NPV means that the investment will be profitable and the project can be accepted.
If NPV is negative, the investment is not economically feasible. Therefore, the
decision rule is to select the option that will induce NPV optimisation. This NPV
rule can be linked to overall social welfare by the Kaldor-Hicks principle, namely
that a positive NPV implies that the gainers could compensate the losers and still
be better off (Hanley & Barbier, 2009).
It should be highlighted that the selection of the lifespan of the technologies is
always a controversial choice since it is well known that it depends on many factors
including the maintenance and management of the facilities.
Regarding the discount rate, higher discount rates favours solutions that are
weighted toward future spending, that is, those with relatively high operating costs
and lower investment cost (Woods et al. 2013). There is much debate over which
discount rate governments should use in public sector policy and project appraisal.
Step 6 Recommendations
(1) Specify the set of alternatives to the project. The CBA compares the NPV of
investing resources in different projects or alternatives. In the majority of the
situations, the alternative to be compared with the one proposed is the status
quo situation, that is, the situation in which the project is not carried out.
(2) Identify the incomes, costs and positive and negative externalities. Once the
alternatives to the project have been specified, the next step is to establish
the internal and negative impacts of each one. Wastewater treatment creates
a number of externalities, including negatives such as GHG emissions
and biological and chemical risks if the treated water is reused and
positive externalities as health benefits, education services, and especially
environmental benefits.
(3) Quantify the incomes, costs and positive and negative externalities. Almost
certainly, this is the most complicated phase of a CBA. On the one hand,
internal impacts are those that have a price determined by the market and
therefore, can be quantified directly. On the other hand, the quantification
of the externalities is much more complex since they have not a price
determined by the market. However, it does not mean that they do not
have value since they contribute to improve people welfare. To quantify
externalities, specific economic valuation methods are needed. Hence, it is
possible to standardize all the units involved in the CBA.
(4) Calculate the net present value. As it is shown in Eq. (2.1), the NPV is
defined as the addition of internal benefits and external benefits. Hence
all the parameters involved to calculate the NPV must be expressed in the
same units (monetary units).
(5) Carry out a sensitivity analysis. So far none of the steps described to
apply a CBA has been taken into account the existence of uncertainty.
Accounting for uncertainty is important in the development of any CBA
since uncertainty could influence the ranking of selecting projects (Flores-
Alsina et al. 2012). Xu and Tung (2008) reviewed a large number of
methodologies applied to deal with uncertainty in the water and wastewater
management including Monte Carlo simulations (Prat et al. 2012), fuzzy
logic models (Kafetzis et al. 2010), Bayesian network models (Barton et al.
2005), statistical tolerances (Bonilla et al. 2004), among others.
(6) Make a recommendation based on the NPV and sensitivity analysis. As it
has been pointed out, the alternative that generates the highest NPV will
be chosen, assuming that some other alternatives have a positive NPV.
Moreover, the sensitivity analysis could show that the project with the
highest NPV is not the best option when uncertainty is considered.
2.2.2 Internal benefit
Internal benefit is the difference between internal costs and internal incomes (that
is, private benefits minus private costs). The internal impacts are those directly
2.2.2.1 Internal cost
In a wastewater treatment project, internal costs are composed by investment costs
(IC) and operation and maintenance costs (OMC) of the facility. If we focus on
cost assessment, both IC and OMC should be adjusted for the time they will occur.
The cost estimation on an annual basis, that is, the total annualised equivalent cost
(TAEC) can be calculated (Eq. (2.3)):
r (1 + r )T
TAEC = I + OMC (2.3)
(1 + r )T − 1
where TAEC is the total annualised equivalent cost in €/year; IC are the investments
costs in €; OMC are the operation and maintenance costs in €/year; r is the discount
rate; and T is the useful life-span of the project.
In the planning of a new investment, cost functions are a useful tool to quantify
IC and OMC as they show the relationship between the dependent variable (cost)
and independent variables (a set of representative variables of the process). Costs
functions are also useful for comparing different treatment technologies from
an economic point of view (Hernández-Sancho et al. 2011). Therefore, cost
functions are widely used to predict IC and OMC of wastewater treatment projects
(Panagiotakopoulos, 2004; Tsagarakis et al. 2003; Nogueira et al. 2007).
In the framework of ‘water and wastewater economics’, there are three main
methodologies to develop costs functions (Molinos-Senante et al. 2013b):
(1) The facility is viewed as a system consisting of components or subsystems,
each of which is simulated in detail (Panagiotakopoulos, 2004). Following
an engineering approach, the design parameters are allowed to assume
values within a wide but realistic range, thus simulating many alternative
facility forms, each with its own estimation.
(2) In the so-called ‘factor method’, major cost drivers related to specific major
cost parameters are known and they are directly estimated (Le Bozec,
2004). Though the use of conversion coefficients for the cost drivers,
estimates from one region or country can be transferred to another.
(3) Statistical and mathematical methods are often used when cost figures
(actual or estimates) are available. These figures might relate to set-up cost
and/or operating cost to the main variables of the facilities.
Previous studies (Sipala et al. 2005; Gonzalez-Serrano et al. 2006) illustrated
that the statistical method is the most common approach for developing cost
functions. Steps from the collection of the raw data to the generation of the costs
functions are shown in Figure 2.2.
(1) Sort through the data basis of technology. Sorting means distinguish
between the various options for saving energy or achieving other objectives
previously defined.
(2) Choose a reference year for economic valuation. Due to the difficulty to
obtain economic data in the framework of wastewater treatment, sometimes
the reference year of all available information is not homogeneous. In this
case, it is necessary to choose a reference year, which generally is the year
of analysis. The costs for other years must be updated.
(3) Decide on the cost components that will be included in the cost functions.
Usually, the treatment capacity of the plant is considered the most important
factor to determine IC and OMC. In this sense, it is very important choose
the size measure of the facilities. In WWTPs, two functional units can be
used, namely population equivalent and volume of wastewater treated.
(4) Choose the functional form of the cost function. The formulation of IC
and OMC functions is based on the assessment of the relationship between
the dependent variable C (cost) and the independent variables X (volume
treated or population equivalent). For this purpose, different models can be
used, such as:
b
Inverse: C =a+
X
Power: C = aX b
Logarithmic: C = a + blnX
Quardratic: C = a + bX + CX 2
(5) Adjust all available data to comply with the choices in Step 3 regarding
cost components. In case that a cost component is missing from the report
cost figure, it must be estimated on the basis of information from other
sources.
(6) Having the sets of the adjusted figures and using appropriate statistical
methods, ‘best-fit’ cost functions are generated. A common method
to get model parameters is ordinary least squares regression analysis.
Subsequently, the significance of the independent variables should be
tested. In doing so, a statistical hypothesis test should be carried out.
(7) Evaluate the quality of the adjustment. The most common indicator to
evaluate the quality of the adjustment is the coefficient of determination (R)
which measures the proportion of total variability of the dependent variable
relative to its average according to the regression model. Its value is ranged
within [0, 1]. If the determination coefficient value is 1, the adjustment
between actual and estimated data is perfect. A value of 0 indicates that
there is no relationship between the variables.
2.2.2.2 Internal income
In a general study of the economic feasibility of a WWTP, internal income
includes revenues obtaining from the sale of the by-products that can be recovered
during the wastewater treatment process. In areas under water stress, the sale
of the recycled water may play a vital role to ensure the economic feasibility of
some water reuse projects. It should be taken into account that if the reclaimed
water is used in agriculture, the nitrogen and phosphorus content in the water
entails a saving in the fertiliser costs (Nogueira et al. 2013). Other incomes may
be obtained from the sale of nutrients (mainly phosphorus) recovered during
wastewater treatment and from the sale of stabilised sewage sludge to be used
after composting.
Focusing on the implementation of technologies for improving energy efficiency,
additional incomes must be quantified and incorporated in the economic feasibility
study. Some technologies involve a reduction in the consumption of energy;
therefore, there is an economic saving that should be taken into account. Other
processes allow recovering energy from wastewater or from sewage sludge that
can be used in the WWTP itself or sold, which supposes an additional income that
cannot be overlooked in the economic feasibility study.
Taking into account internal cost and internal income, the internal benefit for
one year is expressed as follows (Eq. 2.4). To estimate the NPV for the life-span of
the project, the internal benefit must be updated using Eq. (2.2).
BI = ∑ AUM ⋅ SPM + ∑ AE
i
i i
j
j ⋅ SPE j − ( IC + OMC ) (2.4)
where BI is the internal benefit (€/year); AUMi is the annual unit of the material i
recovered such as reclaimed water (m3), phosphorus (kg), nitrogen (kg), composted
sludge (kg), and so on; SPMi is the selling price of recovered material, i (€/m3
or €/kg); AEi is the annual energy recovered through anaerobic digestion, sludge
incineration, and so on, in the form, j such as heat, electricity, and so on (kWh);
SPEj is the selling price of the recovered energy, j, (€/kWh); IC are the investment
costs (€/year); and OMC are the operational and maintenance costs (€/year).
2.2.3 External benefit
An externality is an effect of a purchase or use decision by one party (or group
of parties) on another party who did not have a choice and whose interests were
not taken into account (Hussen, 2004). In other words, an externality is generated
when an economic operation between agents A and B, produces effects on a third
agent C, without any monetary transaction between A and C, or between B and C.
However, the absence of market does not imply the absence of value.
While any internal impact can be calculated directly in monetary units, the
quantification of external impacts requires the use of economic valuation methods
due to the absence of market prices. This requirement is a major difference in
applying CBA rather than of financial analysis (Molinos-Senante et al. 2013a).
Following the same approach as for the internal benefit, the external benefit for
one year is expressed as follows (Eq. (2.5)). As well as internal benefit, it should be
updated for the life-span of the project.
BE = PE − N E (2.5)
Where BE is the external benefit (€/year); PE are the positive externalities such
are health and environmental benefits (€/year); NE are the negative externalities
such as GHG emissions (€/year). External benefits should include the value of
avoided damage costs due to the operation of the plant for example, the value of
avoided damages to recreation.
2.2.3.1 External cost
The benefits of wastewater treatment are obvious, however treatment processes
also result in environmental impacts (Friedrich et al. 2009), such as eutrophication,
and contributions to climate change (Lassaux et al. 2007).
Due to social and political concerns about climate change, there is growing
interest in minimising the consumption of energy in WWTP. Energy consumption
is twofold from the perspective of assessing the economic feasibility of the
wastewater treatment process. On the one hand, as it has been pointed previously, it
is an internal cost. On the other hand, energy consumption is a negative externality,
which should not be overlooked since WWTPs consume a significant amount of
electricity which involves the indirect emission of GHG.
Although IPCC Guidelines (2007) state that CO2 emissions have an impact
factor of 0 kg CO2eq when CO2 has biogenic origins (Doorn et al. 2006) nowadays,
there is an increasing interesting in estimating not just indirect GHG emissions
from energy consumption, but also direct GHG emissions. This is because it has
been verified that IPCC guidelines underestimate the values of GHG emissions
regardless of its origin (biogenic or not) (Foley et al. 2010).
Subsequently, a methodology is described that estimates the economic value
of the GHG emissions, that is, to estimate the value of the negative externalities
associated to wastewater treatment (Molinos-Senante et al. 2013a).
Indirect GHG emissions should be estimated based on WWTP energy demands.
At first, taken into account the national electrical production mix (national scheme
of electrical production), each GHG emission can be estimated. Subsequently, both
direct and indirect emissions should be converted to equivalent CO2 emissions
using 100-year global warming potential coefficients (IPCC, 2007).
Once total GHG emissions have been quantified in physical terms, the next
step is to express them in monetary units. For this purpose, it should be noted
that in the context of the Kyoto Protocol, a well-organised emissions trading has
been developed. For example, in Europe the European Union’s Emissions Trading
System (EU ETS) was implemented in 2005, which integrates more than 11,000
power stations and industrial plants accounting for the 40% of total GHG emissions
in the European Union. The price of CO2 emissions depends on supply and demand,
as well as other macroeconomic factors (Molinos-Senante et al. 2013a).
The average price paid through the EU ETS (or other CO2 market) during a time
period may be used as a proxy to the price of CO2eq emissions. As a reference and
based on SENDECO database, the average market price of CO2 from 2009 to 2012
was 11.9 €/t (SENDECO, 2013). However, there is some concern that the European
carbon market currently set lower prices for CO2 emissions.
2.2.3.2 External benefits
In the context of wastewater treatment, the US Environmental Protection Agency
(EPA) identified that wastewater regeneration and reuse provides the following
environmental benefits (EPA, 1998): (i) decreased diversion of freshwater from
sensitive ecosystems; (ii) decreased discharge to sensitive water bodies; (iii) recycled
water may be used to create or enhance wetlands and river banks; and (iv) recycled
water can reduce and prevent pollution.
Different methodologies for the quantification and internalisation of
environmental externalities arising from investment projects have been developed
from economic theory. Conventional valuation methods can be classified as follows
(Molinos-Senante et al. 2012):
– Methods not based on demand curves such as the replacement cost method,
opportunity cost method, dose-response method, among others. They use
production or cost functions and provide a ‘value to cost’ type approach.
From a methodology point of view, these methods are not complex but they
require considerable experimental information.
– Methods based on demand curves. They belong to the ‘value to value’
approach and they are used to determine the total economic value of goods
and services that have no market (Hanley & Barbier, 2009). They are
classified as:
Indirect methods such as travel cost method and hedonic price method. They
{{
rely on the use of data from actual transactions by individuals. The value of
the environmental good is deducted from the complementary relationship
between it and other goods with market price (Pearce & Turner, 1990).
Direct methods. They are known as stated preference methods since they
{{
are based on the demand approach (Hanley et al. 2006). This approach
responds to the neoclassical view that economic value arises from the
interaction between an individual and an environmental asset as an
expression of individual preference, assuming that these preferences
are a reflection of the maximum utility. The primary categories of
stated preference methods are the contingent valuation method and
choice modelling techniques (see Figure 2.3). The contingent valuation
method is based on the creation of a hypothetical market through a
surveying process where individuals declare their willingness to pay
(WTP) (or be compensated) for an improvement (or degradation) of the
quality of the environmental good being analysed (Genius et al. 2005).
There are several ways to ask WTP questions in contingent valuations
surveys, which are known as elicitation methods. As it is shown in
Figure 2.3 there are four types of elicitation methods. In the open-
ended format, repondents are asked to state their maximum WTP for
the amenity to be valued while in dichotomous choice, respondents
are asked if they are WTP single randomly assigned amount on all-or
nothing basis. The iterative bidding is a series of dichotomous choices
questions starting with an initial low bid that nearly all respondents who
have a WTP > 0, would be willinging to pay. Finally, in the payment
card format, respondents might announce their WTP to the values
listed on the card. Alternatively, the choice modelling techniques are
based on ranking or rating a series of “product profits” that characterise
products with specific attribute levels (Pearce & Özdemiroglu, 2002).
The idea of the contingent ranking method is to give a set of alternatives
which consists of a given amount or a given level of a specific good and
a corresponding realistic price. The alternatives specified in advance
are then ranked (ranking contingent), scored (rating contingent) or
selected (discrete choice experiments) (Slothuus et al. 2002). In most
of the applications related to water resources, the quantification of
these externalities has been made using the stated preference methods
(Guimaraes et al. 2011).
u
D0 ( x, u) = Min θ : ∈ P( x ) (2.6)
θ
where P(x) is a vector of outputs that are technically viable and use the vector of
inputs x, (u/θ) is the outputs ratio in production frontier, while θ is a ratio between
zero and one, that is, D0 ( x, u) ∈ [0,1].
The relationship of duality between the distance function and the revenue
function (Shephard, 1970) is the basis to estimate shadow prices since it creates
the link between relative and absolute price. The relationship between the two
functions can be expressed as in (Eq. (2.7)):
R( x, u) = Max u{ru: D0 ( x, u) ≤ 1}
(2.7)
D0 ( x, u) = Max r{ru: R( x, u) ≤ 1}
where R(x, u) is the revenue function and r represents output prices. Under the
assumption that distance and revenue functions are differentiable, the Lagrange
multiplier method and Shephard’s dual lemma enable us to calculate shadow prices.
This deduction of shadow prices for undesirable outputs means assuming that the
shadow price of an absolute desirable output coincides with the market price. If m is
a desirable output (treated wastewater or reclaimed water in our case) whose market
price is rm equal to its shadow price (rm0 ), and if m′ is undesirable output (a pollutant
removed from wastewater) and rm′ is the shadow price of each undesirable output,
for all m′ ≠ m, the absolute shadow prices are given by (Färe et al. 1993) (Eq. (2.8)):
∂D0 ( x, u) /∂um
rm′ = rm0 (2.8)
∂D0 ( x, u) /∂um′
N M
ln D0 ( x k , u k ) = α 0 + ∑β
n =1
n ln xnk + ∑α
m =1
m ln umk
(2.9)
N N
1
+
2 ∑∑β
n =1 n ′=1
nn ′ (ln xnk )(ln xnk′ )
M M N M
+ ∑∑
m =1 m ′=1
α mm ′ (ln umk )(ln umk ′ ) + ∑∑γ
n =1 m =1
nm (ln xnk )(ln umk )
benefits based on the estimation of shadow prices of pollutants are much lower
than in the case of the traditional methodologies (demand approach) since it is not
required any surveying process.
Nevertheless, the quantification of environmental benefits using the shadow
price methodology also has some limitations in relation to the stated preference
methods since they may be more appropriate than the shadow price method when
the aim is to estimate the total economic value. The cost production approach
methodology may be useful to quantify environmental impacts derived from
production processes while demand approach methodologies can be applied in a
wider context.
2.3 Conclusions
Economic analysis provides tools, information, and instruments for streamlining
the decision-making process. Hence, in the field of wastewater treatment, economic
feasibility studies are a useful tool for selecting the most appropriate option from
among a range of technological alternatives.
Within methodologies to evaluate the economic feasibility of any project, cost
benefit analysis (CBA) provides a comprehensive assessment since, unlike financial
analysis, CBA integrates not just the costs and income with the market value but
also with the positive and negative externalities.
The current chapter presents a framework to assess the economic feasibility of
any innovative technology taking into account both internal and external impacts.
Regarding internal costs it has been illustrated that using cost functions is a common
methodology to estimate both investment, as well as operation and maintenance
costs. Since externalities are not considered by the market, their quantification
requires economic valuation methods. There are two main approaches to estimate
the positive externalities associated to wastewater treatment namely the demand
approach and cost production approach.
As a general conclusion, we emphasise that when the economic feasibility of
a wastewater treatment technology is assessed, water companies and/or water
management authorities should consider impacts with and without market values.
Otherwise, the quality and relevance of the results will be seriously biased and an
mis-estimation of benefits will occur. Moreover, uncertainty could influence the
economic feasibility of wastewater treatment technologies. To narrow uncertainty,
it is essential to perform a sensitivity analysis based on statistical methodologies or
follow the ‘ceteris paribus’ approach.
2.4 References
Almansa C. and Martínez-Paz J. M. (2011). What weight should be assigned to future
environmental impacts? A probabilistic cost benefit analysis using recent advances on
discounting. Science of the Total Environment, 409, 1305–1314.
Barton D. N., Saloranta T., Bakken T. H., Solheim A. L., Moe J., Selvik J. R. and Vagstad
N. (2005). Using Bayesian network models to incorporate uncertainty in the economic
analysis of pollution abatement measures under the water framework directive. Water
Science and Technology: Water Supply, 5(6), 95–104.
Benedetti L., Bixio D. and Vanrolleghem P. A. (2006). Benchmarking of WWTP design by
assessing costs, effluent quality and process variability. Water Science and Technology,
54(10), 95–102.
Bonilla M., Casasús T., Medal A. and Sala R. (2004). An efficiency analysis with tolerance
of the Spanish port system. International Journal of Transport Economics, XXXI(3),
379–400.
Chen R. and Wang C. (2009). Cost-benefit evaluation of a decentralized water system for
wastewater reuse and environmental protection. Water Science & Technology, 59(8),
1515–1522.
De Anguita P. M., Rivera S., Beneitez J. M., Cruz F. and Espinal F. M. (2011). A GIS
Cost-Benefit analysis-based methodology to establish a payment for environmental
services system in watersheds: application to the Calan River in Honduras. Journal of
sustainable Forestry, 30(1), 79–110.
Doorn M. R. J., Towprayoon S., Manso-Vieria S. M., Irving W., Palmer C., Pipatti R.,
and Wang C. (2006). Wastewater treatment and discharge. In: Waste. 2006 IPCC
Guidelines for National Greenhouse Gas Inventories, S. Eggleston, L. Buendia,
K. Miwa and K. Tanabe (eds), IGES, Japan, pp. 6.1–6.28.
EPA (US Environmental Protection Agency) (1998). Water recycling and reuse: the
environmental benefits. Ed. Water Division Region IX. http://www.epa.gov/region09/
water/recycling/brochure.pdf (accessed 28 October 2013).
Färe R., Grosskopf S., Lovell C. A. and Yaisawarng S. (1993). Derivation of shadow prices
for undesirable outputs: a distance function approach. Review of Economics and
Statistics, 75(2), 374–380.
Färe R., Grosskopf S. and Weber W. L. (2001). Shadow prices of missouri public
conservation land. Public Finance Review, 29(6), 444–460.
Flores-Alsina X., Corominas L., Neumann M. B. and Vanrolleghem P. A. (2012). Assessing
the use of activated sludge process design guidelines in wastewater treatment plant
projects: a methodology based on global sensitivity analysis, Environmental Modelling
and Software, 38, 50–58.
Foley J., de Haas D., Yuan Z. and Lant P. (2010). Nitrous oxide generation in full-scale
biological nutrient removal wastewater treatment plants. Water Research, 44(3),
831–844.
Friedrich E., Pillay S. and Buckley C. A. (2009). Environmental life cycle assessments for
water treatment processes – A South African case study of an urban water cycle. Water
SA, 35(1), 73–84.
Gavasci R., Chiavola A. and Spizzirri M. (2010). Technical-economical analysis of selected
decentralized technologies for municipal wastewater treatment in the city of Rome,
Water Science and Technology, 62(6), 1371–1378.
Genius M., Manioudaki M., Mokas E., Pantagakis E., Tampakakis D. and Tsagarakis K. P.
(2005). Estimation of willingness to pay for wastewater treatment. Water Science and
Technology: Water Supply, 5(6), 105–113.
Gonzalez-Serrano E., Rodriguez-Mirasol J., Cordero T., Koussis A. D. and Rodriguez J. J.
(2006). Cost of reclaimed municipal wastewater for applications in seasonally stressed
semi-arid regions. Journal of Water Supply: Research and Technology- AQUA, 54(6),
355–369.
Guimaraes M. H., Sousa C., Garcia T., Dentinho T. and Boski T. (2011). The value of
improved water quality in guadiana estuary-a transborder application of contingent
valuation methodology. Letters in Spatial and Resource Sciences, 4(1), 31–48.
Hanley N. and Spash C. (1993). Cost – Benefit Analysis and the Environment. Edward
Elgar, Cheltenham.
Hanley N., Colombo S., Tinch D., Black A. and Aftab A. (2006). Estimating the benefits
of water quality improvements under the water framework directive: are benefits
transferable? European Review of Agricultural Economics, 33(3), 391–413.
Hanley N. and Barbier E. B. (2009). Pricing Nature: Cost-Benefit Analysis and
Environmental Policy. Edward Elgar, Cheltenham.
Hernández-Sancho F., Molinos-Senante M. and Sala-Garrido R. (2011). Cost modelling for
wastewater treatment processes. Desalination, 268, 1–5.
HM, Treasury. (2003). The Green Book Appraisal and Evaluation in Central Government.
TSO, London.
Hussen A. M. (2004). Principles of Environmental Economics, Routledge, New York.
IPCC (Intergovernmental Panel on Climate Change) (2007). Summary for Policy Makers.
In: Climate Change 2007: The Physical Science Basis. Contribution of Working Group
I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change,
S. Solomon, D. Qin, M. Manning, Z. Chen, M. Marquis, K. B. Averyt, M. Tignor and
H. L. Miller (eds), Cambridge University Press, Cambridge and New York.
Kafetzis A., McRoberts N. and Mouratiadou I. (2010). Using fuzzy cognitive maps to
support the analysis of stakeholders’ views of water resource use and water quality
policy. Studies in Fuzziness and Soft Computing, 247, 383–402.
Lassaux S., Renzoni R. and Germain A. (2007). Life cycle assessment of water from the
pumping station to the wastewater treatment plant. International Journal of Life Cycle
Assessment, 12, 118–126.
Le Bozec A. (2004). Methodology for the Determination of Production Costs and Full
Costs of Municipal Solid Waste Treatments. Deliverable Report of the Project: Aid
in the Management and European Comparison of Municipal Solid Waste Treatment
methods for a Global and Sustainable Approach. Cemagert, France.
MacLeod S. P. and Filion Y. R. (2012). Issues and implications of carbon-abatement
discounting and pricing for drinking water system design in Canada. Water Resources
Management, 26(1), 43–61.
Molinos-Senante M. (2011). Techno-economical efficiency, cost modelling and economic
valuation of environmental benefits from wastewater treatment: methodology and
applications. PhD dissertation, University of Valencia.
Molinos-Senante M., Hernández-Sancho F. and Sala-Garrido R. (2010). Economic
feasibility study for wastewater treatment: a cost-benefit analysis. Science of the Total
Environment, 408, 4396–4402.
Molinos-Senante M., Hernández-Sancho F. and Sala-Garrido R. (2012). Economic
feasibility study for new technological alternatives in wastewater treatment processes:
a review. Water Science and Technology, 65(5), 898–906.
Molinos-Senante M., Hernández-Sancho F. and Sala-Garrido R. (2013b). Cost modeling
for sludge and waste management from wastewater treatment plants: an empirical
approach for Spain. Desalination and water treatment, 51(28–30), 5414–5420.
Woods G. J., Kang D., Quintanar D. R., Curley E. F., Davis S. E., Lansey K. E. and Arnold
R. G. (2013). Centralized versus decentralized wastewater reclamation in the houghton
area of Tucson, Arizona. Journal of Water Resources Planning and Management,
139(3), 313–324.
Xu Y.-P. and Tung Y.-K. (2008). Decision-making in water management under uncertainty.
Water Resources Management, 22(5), 535–550.
1www.epa.gov/statelocalclimate/local/topics/water.html.
the supporting services, such as in vehicles and office and workshop buildings,
which is not addressed in this chapter.
Wastewater treatment plants (WWTPs) are pieces of a more general complex
puzzle – the wastewater system as a whole. While recognizing the importance
of improving the efficiency of use of energy in economic and environmental
terms, most utilities and research teams tend to explore and implement sectorial
approaches – equipment or process-based, thus lacking a global analysis of the whole
system functioning. Instead, a strategic approach may lead to much better service
at lower total asset life-cycle costs. Selection of plant location, targets of treatment
process effectiveness and reliability, and control of rain water inflow are examples
of aspects that should be addressed in an integrated way (i.e., understanding the
behaviour of the wastewater system as a whole), and incorporating a long term
analysis planning horizon of WWTPs.
Let us take a simple example to illustrate the idea. An energy manager needs
to compare the behaviour and potential for improvement of two similar systems
in terms of energy, in order to prioritize intervention efforts (Figure 3.1). Both
systems start at a pumping station at level 0.00 m, that pumps the wastewater from
a gravity network to a wastewater treatment plant. The diameter, material and
length of the pressurized sewer are identical in both cases. The level of the WWTP
inflow point is 65.00 m in System 1 and 45.00 m in System 2. A volume of 800
m3/day is pumped daily during 8 h per day,in both cases. There is a flow control
valve upstream the WWTP in the case of System 2.
The question this energy manager needs to answer is ‘which system is less
efficient and, therefore, has more potential for improvement?’. The table in Figure
3.1 shows that System 1 has a lower pump efficiency, and a higher specific and
total energy consumption. These are typical indicators to assess energy efficiency.
When comparing the two systems using these indicators, it seems safe to reply
‘System 1’ to the question. However, a system analysis may provide a different
view. Table 3.1 shows the calculations of the global energy efficiency, assessed
as the ratio between the energy supplied to each system and the energy actually
supplied to the users, assuming a no-losses situation. If there were losses, the
downstream flow should be the authorized consumption.
System 1 System 2
Daily energy input supplied to the system 212.5 184.5
(Einput = γ ⋅ Qt ⋅ Ht /η × 8) (kWh)
Daily energy supplied to the consumers 141.6 98.0
(Esupplied = γ ⋅ Qt ⋅ Hs × 8) (kWh)
Overall energy efficiency 1.50 1.88
(Einput /Esupplied) (−)
Table 3.1 shows that the conclusion is opposite: System 2 is globally less efficient
than System 1. Figure 3.2 shows the energy line for both systems.
The head-loss caused by the flow control valve is much higher than the pump
inefficiency. This example demonstrates that managing energy on an ‘asset by
asset’ basis may easily fail to pinpoint the critical energy inefficiencies. The most
cost-efficient intervention is likely to be changing the pump of System 2, in order
to eliminate the need for the flow control valve.
Although not exclusively, typical sources of inefficiencies are:
• Control valves: as illustrated in the example, are a very typical asset where
energy is wasted, particularly in the network part, in general less relevant
inside the WWTP; as a result, special attention shall be paid to the valves
located downstream pumps.
• System design: very rarely systems are solicited as planned by the time
of their original planning, design and construction; as illustrated in the
Figure 3.4 A xis of energy performance (adapted from ISO 50001: 2011).
Figure 3.5 Aims and benefits of applying ISO 50001 in a wastewater utility.
ISO 50001 Energy Management System is suitable for all businesses regardless
of their size, geography or industry. It is particularly effective in energy intensive
industries, as it is in the general situation of wastewater utilities. It is also effective
if the utility faces greenhouse gas (GHG) emissions regulation or legislation, or for
Figure 3.6 Management system model of ISO 50001 (ISO 50001: 2011).
Each box of Figure 3.6 corresponds to a set of requirements in ISO 50001: 2011.
The most relevant for the establishment of a framework of energy performance
assessment system for WWTPs, object of the following section, relates to the
planning and checking stages of this model and of its overarching Plan-Do-Check-
Act framework. For instance, under the item ‘Energy objectives, targets and
programmes’ of the ‘Plan’ section, this standard states that the organization shall:
• Establish and document measurable energy objectives and targets at the
relevant function and levels within the organization.
• Set specific targets for those controllable parameters that have a significant
impact on energy efficiency.
The IAM analyses shall therefore be based on the long term behavior of
functional units of assets. Inevitably, assets in different stages of their life cycle
coexist. A time window of some decades needs to be used in the analysis, in order
to embrace the life cycle of the key assets (Alegre & Coelho, 2012).
The IAM AWARE-P methodology, developed in Portugal, incorporate industry’s
best practices and is currently being promoted and adopted internationally (e.g., by
the international Water Association), associating the ISO 24500 (ISO 24510/24511:
2007) and the ISO 55000 (ISO 55000/ISO 55001/ISO 55002: 2014) principles with
the characteristics of the urban water systems. In the core of the methodology
(Figure 3.7) are the establishment of clear corporate objectives and an educated
choice of assessment criteria, supported by adequate metrics and quantifiable
targets (Alegre & Covas, 2010; Almeida & Cardoso, 2010; Alegre & Coelho, 2012;
Alegre et al. 2013).
Figure 3.7 IAM planning process at each decision levels (Alegre et al. 2010).
The process is applied at the strategic, tactical and operational decisional levels
in the utility, striving for alignment of objectives, metrics and targets between
levels, as well as consistent feedback across levels (Figure 3.8).
One of the main advantages of such a structure approach is the establishment
of a decision process that is transparent, defendable and yet simple, allowing to
compare and prioritize intervention alternatives potentially of a very different
nature.
Figure 3.8 Alignment and feedback between decision levels (Alegre et al. 2010).
PI system one should select a set of PIs according to the assessment objectives, in
this case the WWTP energy performance in a given year.
The energy performance is transversal to four assessment groups described above,
ER, RU, BP and Fi. Hence, the PI set for energy includes (Silva et al. 2013): (i) four
1st level PIs (Table 3.2), one from each of those groups, and (ii) 14 complementary
PIs (Table 3.3), related to WWTP reliability (e.g., adequacy of plant treatment and
pumping capacities, recycling and aeration control, and inspection of key equipment)
and renewable (wind and solar photovoltaic) energy production.
The reference assessment period is the calendar year, though PIs expressed in
temporal terms are formulated to accommodate other reference assessment periods.
In order to ensure unit coherence and allow for comparisons, for these PIs (e.g.,
wtER35.1, Table 3.3), values calculated for other reference assessment periods are
converted into annual values, multiplying by ‘(365 days/year)/Assessment period
(day)’ (Silva et al. 2012).
The PIs are identified by a code that includes six or eight fields (the last two
are optional). These fields identify the system (t for WTP and wt for WWTP),
the assessment group (e.g., ER for the plant Efficiency and Reliability assessment
group) and the number (e.g., 01) of the PI. Two additional fields (number and/
or letter) can be used to identify an alternative processing rule (designated by
a different number) or a speciation (using a different letter for each species).
Examples of alternative processing rules are wtER35.1 and wtWER35.2 for Pump
inspection or Inspected pumps, respectively (Table 3.3) and wtRU03.1, wtRU03.2
and wtRU03.3 for energy consumption per m3 of treated wastewater, per kg BOD
removed or kg COD removed, respectively (Table 3.2).
PIs are defined by default per m3 of treated wastewater (e.g., wtRU03.1,
Table 3.2) but they may be expressed per kg BOD or kg COD mass removed
by dividing the obtained value by the explanatory factor wtEF03 for BOD or
wtEF04 for COD (Silva & Rosa, 2014). The explanatory factors (EFs, Table 3.4)
evaluate complementary aspects of the plant performance that assist the correct
interpretation of some PIs.
Tables 3.2 and 3.3 also present, whenever applicable, the PI reference
values defining three performance levels: “good” (), “acceptable” () and
“unsatisfactory” ().
The reference values are based on the PI results of 17 Portuguese WWTPs in a
5-year period (2006–2010) and on the relations derived from literature data. The
reference values for the unit energy consumption reflect the inverse relations with
the volume treated and are specific for activated sludge systems (conventional, with
coagulation/filtration (C/F) and with nitrification and C/F) and trickling filters.
The reference values for energy production were derived based on the methane
generation potential and literature data; those of net use of energy were considered
the difference between the references for energy consumption and energy
production (Silva & Rosa, 2014).
Examples of application of the energy PIs are given in Silva and Rosa (2014).
PI code, units, processing rule and reference values for “good” (), “acceptable” () and “unsatisfactory” () performance
wtRU03.1 – Energy consumption [kWh/m3]
Energy consumption (kWh)/Treated wastewater (m3)
Trickling filters ≤0.185 + 1127/TW ]0.185 + 1127/TW; 0.231 + 1409/TW[ ≥0.231 + 1409/TW
Activated sludge (AS) ≤0.280 + 1192/TW ]0.280 + 1192/TW; 0.350 + 1490/TW[ ≥0.350 + 1490/TW
AS + Coagulation/Filtration ≤0.325 + 1384/TW ]0.325 + 1384/TW; 0.406 + 1730/TW[ ≥0.406 + 1730/TW
AS with nitrification + ≤0.424 + 1362/TW ]0.424 + 1362/TW; 0.530 + 1703/TW[ ≥0.530 + 1703/TW
Coagulation/Filtration
wtRU03.2 – Energy consumption [kWh/kg BOD removed]
Energy consumption (kWh)/BOD mass removed (kg)
by marinhoeng@hotmail.com
Table 3.2 First level PIs for energy performance (adapted from Silva & Rosa, 2014) (Continued).
PI code, units, processing rule and reference values for “good” (), “acceptable” () and “unsatisfactory” ()
performance
wtBP18.2 – Production of energy from biogas [%]
Energy produced from biogas (kWh)/Energy consumption (kWh)*100
wtER08 – Net use of energy from external sources [kWh/m3]
(Energy acquired to external sources – Energy sold to external users (kWh))/Treated wastewater (m3)
Trickling filters ≤0.185 + 1127/TW ]0.185 + 1127/TW − 0.0009 BOD5; ≥0.231 + 1409/TW
− 0.0009 BOD5 0.231 + 1409/TW − 0.0007 BOD5[ − 0.0007 BOD5
Activated sludge (AS) ≤0.280 + 1192/TW ]0.280 + 1192/TW − 0.0009 BOD5; ≥0.350 + 1490/TW
AS with nitrification + Coagulation/ ≤0.424 + 1362/TW ]0.424 + 1362/TW − 0.0009 BOD5; ≥0.530 + 1703/TW
Filtration − 0.0009 BOD5 0.530 + 1703/TW − 0.0007 BOD5[ − 0.0007 BOD5
3
wtFi05 – Electrical energy costs [€/m ]
Electrical energy costs (€)/Treated wastewater (m3)
TW = treated wastewater (m3/d); BOD5 = influent BOD5 (mg/L); PI code, units, processing rule and reference values
Table 3.3 Complementary PIs for assessing WWTP energy performance (adapted from Silva et al. 2013).
by marinhoeng@hotmail.com
PI code, units and processing rule Ref. Values
wtER13 – Adequacy of plant capacity [%]
n n [80; 100]
Qtd × Jd + Qtd × K d [60; 80[
∑ d =1 ∑ d =1
1 − n × 100 [0; 60[
Qtd
∑ d =1
(Continued)
Introduction to energy management in wastewater treatment plants 47
Table 3.3. Complementary PIs for assessing WWTP energy performance (adapted from Silva et al. 2013) (Continued ). 48
by marinhoeng@hotmail.com
PI code, units and processing rule Ref. Values
wtER20 – Pumping utilisation upstream of the WWTP [%] = Maximum daily volume of raw [80; 90]
wastewater (RW) pumped (m3)/RW maximum pumping capacity (m3)*100
wtER21 – Pumping utilisation downstream of the WWTP [%] = Maximum daily volume of treated [70; 80[ or
wastewater (TW) pumped (m3)/TW maximum pumping capacity (m3)*100 ]90; 95]
[0; 70[ or
>95
wtER28 – Recycling control [%]
Recycling pumps with automatic control (no.)/Recycling pumps (no.)*100
wtER29 – Aeration or mixing control [%]
Aerators or mixers with automatic commands (no.)/Aerators or mixers (including reserves) (no.)*100
wtER35.1 – Pump inspection [no./(pump.year)]
(Pump inspections (no.) × 365 (day/year)/Assessment period (day))/Pumps (no.)
(Dewatering and sludge transportation equipment inspections (no.) × 365 (day/year)/Assessment period (day))/
Dewatering and sludge transportation equipment (no.)
wtER35.2 – Inspected pumps [%/year] 100
Inspected pumps (no./year)/Pumps (no.) × 100 ]90; 100[
[0; 90]
wtER37.2 – Inspected aerators or mixers [%/year]
Inspected aerators or mixers (no./year)/Aerators or mixers (no.) × 100
wtBP19 – Production of wind and solar energy [%]
Energy produced by wind power and solar photovoltaic (kWh)/Energy consumption (kWh)*100
“good” (), “acceptable” () and “unsatisfactory” () performance
Introduction to energy management in wastewater treatment plants 49
The PXs range between zero and 300, in which: 300 corresponds to a situation
where the performance is ‘excellent’; values between 300 and 200 reflect ‘good’
performances; values between 200 and 100 are ‘acceptable’; 100 corresponds to
the ‘minimum acceptable’ performance; values below 100 reflect ‘unsatisfactory’
performance (Silva et al. 2013, 2014a, b).
To define the performance function, reference values are defined for each
performance level based on the ranges recommended in the literature for each unit
operation/process variant.
A portfolio of state-variables of energy PXs is proposed in Table 3.5. The
reference values for energy PXs are under development. Whenever necessary,
equations are being developed to produce the reference values as a function of
other key-parameter(s).
Table 3.5 State-variables relevant for assessing the WWTP energy performance
(adapted from Silva et al. 2013).
(Continued)
Table 3.5 State-variables relevant for assessing the WWTP energy performance
(adapted from Silva et al. 2013) (Continued).
2Applicable to conventional (with or without sludge return) and enhanced primary sedimentation.
3Variables applicable to several AS variants: complete mix; plug flow (conventional, extended
aeration); oxidation ditch (C removal, C + N removal); A/O; MLE; Bardenpho (4-stage, 5-stage); A2/O;
UCT; VIP; SBR (C removal), SBR (C + N removal) and SBR (C + N + P removal). The performance
functions are AS-type specific.
4Except for SBR.
5Variables applicable to different filter packing (mono and dual media) and height; performance
and rotary drum thickening. The performance functions are thickener-type specific.
8Applicable only to anaerobic digestion.
9Applicable to centrifuge, belt filter press, recessed-plate filter press and sludge drying beds. The
assessment period (usually, a calendar year) and the PXs assess ‘where’ (unit
operations/processes and equipment) and ‘when’ the performance did satisfy or fail
the pre-established objectives and the distance that remains to achieve these targets
(Silva et al. 2014a), enabling the identification of improvement actions.
As introduced earlier, prior to WWTP energy efficiency one should verify
the plant effectiveness and reliability, that is, the compliance over time with the
quality requirements of the treated water (Silva et al. 2014a). One may then select
the parameters and unit operations or processes for which one intends to assess and
optimise the removal efficiency to enhance the treatment reliability (Silva et al. 2014b).
The continuous improvement of WWTP performance requires the verification
and, eventually, the (re)definition of objectives and (re)selection of the corresponding
PIs and PXs, which restarts the PDCA cycle of PAS application (Figure 3.10).
Figure 3.10 PDCA methodology for the continuous improvement of WWTP energy
performance.
3.5 References
Alegre H., Baptista J. M., Cabrera Jr. E., Cubillo F., Duarte P., Hirner W., Merkel W. and
Parena R. (2006). Performance Indicators for Water Supply Services. 2nd edn, Manual
of Best Practices Series, IWA Publishing, London, ISBN: 1843390515.
Lingsten A. and Lundkvist M. (2008). Description of the Current Energy Use in Water
and Wastewater Systems in Sweden (in Swedish). The Swedish Water & Wastewater
Association, SWWA.
Matos R., Cardoso A., Ashley R., Duarte P., Molinari A. and Schulz A. (2003). Performance
Indicators for Wastewater Services. Manual of Best Practices Series, IWA Publishing,
London, ISBN: 19002229006.
Mizuta K. and Shimada M. (2010). Benchmarking energy consumption in municipal
wastewater treatment plants in Japan. Water Science and Technology, 65(10), 2256–2262.
NREL (2012). Energy Efficiency Strategies for Municipal Wastewater Treatment Facilities.
National Renewable Energy Laboratory, Technical Report NREL/TP-7A30-53341.
Colorado, January.
PG&E (2003). Municipal Wastewater Treatment Plant Energy Baseline Study. Pacific Gas
and Electric Company (PG&E). San Francisco, CA, June.
Quadros S. (2010). Desenvolvimento de um sistema de avaliação de desempenho de
estações de tratamento de Águas residuais urbanas. PhDThesis, Azores University,
Angra do Heroísmo, p. 286 (in Portuguese).
Quadros S., Rosa M. J., Alegre H. and Silva C. (2010). A performance indicators system for
urban wastewater treatment plants. Water Science and Technology, 62(10), 2398–2407.
Rodriguez-Garcia G., Molinos-Senante M., Hospido A., Hernández-Sancho F., Moreira
M. T. and Feijoo G. (2011). Environmental and economic profile of six typologies of
wastewater treatment plants. Water Research, 45, 5997–6010.
Rosa M. J., Ramalho P., Silva C., Vieira P., Quadros S. and Alegre H. (2010). PASt21 –
Iniciativa Nacional de Avaliação de Desempenho de ETA e ETAR Urbanas (‘PASt21’ –
the Portuguese initiative for performance assessment of water and wastewater treatment
plants). In: Proc. 10.° Congresso de Água. Alvor (Algarve). March, p. 14 (in Portuguese).
Silva C. (2008). Aplicação de medidas de avaliação de desempenho a estações de tratamento
de Água da Águas do Algarve (Application of performance assessment measures to
Águas do Algarve water treatment plants). MSc Thesis. Algarve University, Faro,
p. 286 (in Portuguese).
Silva C. and Rosa M. J. (2014). Energy Performance Indicators of Wastewater Treatment –
A Field Study with 17 Portuguese Plants. IWA World Water Congress, Lisbon,
September.
Silva C., Ramalho P., Quadros S., Alegre H. and Rosa M. J. (2012). Results of ‘PASt21’ – the
portuguese initiative for performance assessment of water and wastewater treatment
plants. Water Science and Technology – Water Supply, 12(3), 372–386.
Silva C., Ramalho P. and Rosa M. J. (2013). Curative actions in wastewater treatment
systems. Chapter 2 of PART I of report D43.1 – Results from Overall Exploration of
Potential for Sustainability Improvements. TRUST Project.
Silva C., Quadros S., Ramalho P. and Rosa M. J. (2014a). A tool for assessing treated
wastewater quality in urban WWTPs. Journal of Environmental Management, 146,
400–406.
Silva C., Quadros S., Ramalho P., Alegre H. and Rosa M. J. (2014b). Translating removal
efficiencies into operational performance indices of wastewater treatment plants.
Water Research, 57, 202–214.
UNESCO (2014). The United Nations World Water Development Report 2014: Vol. 1 –
Water and energy (WWDR 2014). United Nations Educational, Scientific and Cultural
Organization, UNESCO 2014. ISBN 978-92-3-104259-1.
4.1 Introduction
Sewage treatment accounts for over 1% of the total electrical requirements in the
developed world and with increasingly tighter effluent discharge regulations the
energy requirement for acceptable sewage treatment will increase even further.
With this in mind there is a growing demand for energy efficient technologies
(Caffoor, 2008). These technologies must achieve equivalent or superior levels of
effluent treatment using fewer resources and therefore having less total impact on the
environment. Biological treatment technologies which have been the cornerstone
of Sewage Treatment Plants (STP) for the past 100 years have traditionally fallen
into two categories (1) Extensive and (2) Intensive. While extensive, low or zero
energy systems can be used in rural or remote areas, for urban settlements intensive
treatments systems are required to reduce the space required for the STP and hence
a high energy demand is associated with these technologies. In recent decades
there has been significant development and improvement in intensive biological
treatment processes with the development of processes such as the Membrane
BioReactor (MBR) and the Moving Bed Biofilm Reactor (MBBR). While these
technologies have reduced the volumetric capacity required for treatment they also
have a high energy consumption rate. Anaerobic treatment processes on the other
hand require little or no energy input above pumping and mixing requirements.
There have been major advances in anaerobic treatment with the direct anaerobic
treatment of sewage possible in warm climates, (Sghezzo et al. 1988) and currently
there is a strong focus on the development of the anaerobic MBR for the treatment
of sewage. Despite these advances aerobic treatment is still required for the removal
of many different pollutants which are not broken down anaerobically for example,
ammonia, and also to treat the wastewater to a sufficient quality so that it can be
discharged back into the surface water. Recently there has been a move toward
innovation in aerobic biological reactors, with reduced energy requirement being of
key importance. Not only does this save on operating cost and carbon footprint but
there is also the added incentive of achieving an energy neutral STP by maximizing
the energy being produced and minimizing energy consumed. With this in mind
this chapter will focus on innovative aerobic biological processes and technologies
which are still in development or are at the early stages of development. The main
aspect of these technologies will be their potential for significant overall energy
reduction. While some of these bioreactors may or may not be utilised for the
direct treatment of the sewage stream they still have the potential to be used in a
STP for the treatment of one of the streams and offset the total energy demand and
operational cost.
4.2 Aeration
In STPs the single largest user of energy is the aeration of the biological treatment
processes. The current standard technology of using fine bubbles created via blowers
and diffusers, has been studied for many years and despite many improvements being
made throughout this time, the process remains highly inefficient, with a Standard
Oxygen Transfer Efficiency (SOTE) of 5% per m water depth, resulting in only 30%
of the oxygen supplied being transferred to the water in a 6 m deep tank. Activated
Sludge aeration requires 55.6% of the total energy consumption in a treatment plant
having an average energy requirement of 0.634 kWh/m3 sewage treated (Caffoor,
2008). Therefore the direct aeration energy requirements are 0.348 kWh/m3 of
sewage treated. It is estimated that thru optimisation of operations, retrofitting of
more energy efficient blowers and diffusers and innovation in the process design
used in existing wastewater treatment systems a reduction of the energy consumption
by 20% is quite feasible. (UKWIR Report 2010) While it is important that this is
carried out on existing systems to give a ‘quick win’ the potential for further energy
savings still exist through innovative aerobic treatment processes.
Tank, but in most cases the influence of the parameter requires additional energy
and the modification may have an energy neutral or even energy negative impact.
e.g., By increasing the tank water depth from 3 m to 6 m a rising bubble has
approximately twice the residence time in the water with which to transfer oxygen.
But for air to exit a diffuser plate underneath 6 m of water it must have pressure
greater than 6 m hydrostatic head. This results in an increase of 80% in the
theoretical energy required to achieve a two fold increase in OTE.
4.3.1.1 Smaller bubbles
Decreasing the bubble size which results in an increased gas-liquid surface area
available for mass transfer (Table 4.1) has so far achieved a relatively good success
at increasing the overall energy efficiency. This has resulted in fine bubble aeration
becoming the standard aeration system recommended a head of coarse and
medium bubble aeration, despite fine bubble aeration having a higher maintenance
requirement.
Table 4.1 Increase in bubble surface area available for mass transfer due to
decreasing bubble size.
Smaller than fine bubbles are ultrafine bubbles or micro bubbles, but concerns
still exist regarding these technologies for the aeration of wastewater. Microbubbles
can produce a much higher gas liquid interfacial surface area, but they also have
a higher energy cost than typical gas distributors, the microbubble generators can
allow downsizing of the aeration tank and shorten the overall residence time of
the wastewater so that a reduction in the overall cost can be achieved (Terasaka
et al. 2011). Despite having close to 100% Oxygen Transfer Efficiency it has been
observed that microbubble generation is influenced by fouling which can occur on
the porous membranes (Liu et al. 2013) and the aeration process can have adverse
effects on the overall process, causing floatation and/or the break-up of activated
sludge flocs. (Liu et al. 2012). Disadvantages of micro bubble aeration include
higher capital cost, a higher head loss across the diffuser, increased air filtration
requirements, and a tendency for the microbubble generation membrane to tear
when over-pressurized.
4.3.2.1 Diffusaire
With a view to achieving a higher Oxygen Transfer Efficiency with respect to the
energy input Duiffusaire, an Israeli company have developed a system to increase
the retention time of the bubble in an activated sludge tank. Air is added to the
wastewater in a vertical tube placed in the Activated Sludge tank. The wastewater
to be aerated is then pumped in such a manner as to create a downward flow in
the tube and thereby increase the residence time of the bubble in the liquid, and
allow more time for oxygen transfer. (Yousfan et al. 2011) The aerated liquid along
with some entrapped air then escapes from the bottom of the tube back into the
activated sludge tank. It is claimed that diffusaire proprietary technology can
reduce the energy requirements for aeration by up to 50% (Diffusaire, Advanced
Aeration Solutions).
4.3.2.2 Sorubin
Although somewhat similar in outer appearance to the diffusaire technology,
OptusAir developed by a Sweedish company Sorub in creates a vortex in the centre
of the vertically mounted tube through the use of an impellor placed at the bottom
of the tube. The vortex entraps air from the surface and disperses this air liquid
mixture from the bottom of the vertical tube into the wastewater treatment tank.
Optusflow could enhance the effect of aeration and other treatment methods by
up to 50%.
4.3.2.3 Sansox OY
The OxTubeBySansox (Finland), also increases the contact time and relatively
velocity between the gas and liquid, but do so in a pipe. Using a specially developed
static mixer to break up the flow pattern and create lots of eddies the OxTube
minimises the boundary layer diffusional resistance between the liquid and gas
bubble (Sansox, Oxytube).
Although the initial goal for the development of this technology was to increase the
oxygen transfer efficiency to the wastewater, the subsequent biofilm which formed on
the gas permeable membrane surface, which was initially seen as barrier to oxygen
transfer, was later discovered to take an active role in the wastewater treatment process
and lead to the development of the Membrane Aerated Biofilm Reactor (MABR)
(Timberlake et al. 1988). As a means to increase the oxygen transferred from the
air to the wastewater, by placing the oxygen containing gas inside of amembrane,
the residence time of the air in the wastewater is decoupled from the buoyancy
forces which normally limit the contact time between the air and water. Therefore
it is possible to dramatically increase the Oxygen Transfer Efficiency (OTE) up to a
theoretical maximum of 100%. By controlling the flowrate and the partial pressure
of Oxygen through the membranes the OTE and the Oxygen Transfer Rate (OTR)
can be controlled. This allows for an additional level of control over the oxygen
transfer process for an operator. Although a biofilm colonises the membrane surface
preventing the transfer of oxygen to the wastewater the membrane supported biofilm
takes an active part in the wastewater treatment and because the consumption of
oxygen occurs locally to the membrane surface the active biofilm increases the rate
of oxygen flux across the membrane (Shanahan & Semmens, 2006).
The unique counter diffusional concentration profile which is created through
the membrane aeration, also results in the natural development of aspatially unique
biofilm. This membrane aerated biofilm is a counter diffusional biofilm (Figure 4.1)
as opposed to a conventional co-diffusional biofilm.
Along with increasing the OTE the gas pressure required to supply the oxygen
to the MABR is dramatically reduced and subsequently the overall energy
requirement. The structural integrity of the membranes themselves prevents the
hydrostatic pressure closing off the lumen of the membrane and if non-porous
membranes are used no flow of water back into the membrane lumen can occur.
Therefore the supplied gas does not have to overcome the hydrostatic head and the
only pressure which the supplied gas has to overcome is the pressure drop due to
the air flow within the membranes. It must also be noted that because oxygen is
not being transferred to the wastewater there are no wastewater surface tension
forces to be overcome, therefore the alpha factor does not play a role in the oxygen
transfer in a MABR.
While university based research groups have been examining this technology
for some time with one of the first papers identifying the wastewater treatment
potential being published by Timberlake et al. 1988, until recently the technology
had not progressed beyond the lab scale. During this time the MABR has been tested
for different applications including high rate BOD removal, tertiary nitrification
and simultaneous nitrification and denitrification. These have all been identified as
areas where the MABR has a significant Advantage over conventional systems and
due to the increased level of process control the MABR can be tailored for each of
these processes or for a multiple of different processes sequentially.
Despite this concept being around for many years 2 major stumbling blocks
have prevented it from being commercialised.
(1) The availability of cheap suitable membranes to exploit the cost saving
which results from the reduced energy requirement.
(2) The ability to achieve effective long term stable performance, this has
typically been most difficult in wastewaters with high BOD loading rates.
In an economic evaluation of the MABR (Casey et al. 2008), the two major
economic factors for the commercialisation of the MABR were identified as
membrane replacement cost and energy cost as these directly influence the increased
capital cost associated with membranes in a tank and the operational saving. Today
with more and more membrane providers coming to the market and increased
membrane operational and production knowledge, membrane technology has now
become commoditised and the cost of membranes has reduced significantly over
the past 10 years. This coupled with the increased process knowledge, higher levels
of biofilm understanding and the ability to have effective biofilm control has led to
a number of companies commercialising the MABR concept.
Two different approaches have been taken to the scale-up of the membrane
aerated biofilm concept.
Figure 4.2 Photograph of gas permeable membranes covered with biofilm (provided
courtesy of Oxymem.)
A summary of the membranes used from both systems is given in Table 4.2.
Table 4.2 Submerged MABR data taken from (1) Syron et al. (2014), (2) Stricker
et al. (2011), (3) Adams et al. (2014).
with a defined space between each coil so that air can flow all around the membrane.
The wastewater to be treated is then pumped into the top of the membrane envelope,
and flows around the spiral to the end where it exits the membrane. The biofilm
grows on the inside of the membrane and any sloughed or detached biomass leaves
with the treated wastewater to be removed in a subsequent unit operation. Energy
requirements of 0.02 kwh/m3 of water treated have been reported for this system
(Spiral Aerobic Biofilm Reactor-Emefcy).
Another similar concept which pumps the wastewater to be treated to the top
of an above ground membrane unit is the Biogill developed by Biogill Operations
Pty ltd (Australia). Biogill utilises folded Nano Ceramic membranes and the
wastewater flows downward inside the membranes with a biofilm developing on
the inside of the membrane. The nanoceramic membranes chosen by Biogill are
porous to water and some of the wastewater permeates through the membranes
creating a wet membrane surface and an environment suitable for the growth
of fungi when contributes to nutrient removal. The wastewater is then collected
in a decant tank underneath the membrane unit where the detached biomass is
allowed to settle out of the wastewater. Based on the design parmeters given in
the Biogill Technical System and specification guide the energy required for
treating the water is 0.17kWh/m 3 at the scenario given in Table 4.3.
oxygen. The product of this two stage biological process was nitrogen gas which
escaped back into the atmosphere. Depending on the exact configuration chosen this
process generally required significantly more air, additional recirculation pumps,
longer biomass retention and significantly larger tanks than the more traditional
biological carbon removal processes. Overall the energy increase for the addition of
nitrification was estimated at 69% in a case study by Menendez and Veatch, 2010.
Thanks to an increased understanding of the biological nitrogen removal process,
scientists and engineers have been able to provide the suitable conditions for the
each of the different bacterial groups which are involved with multiple stages of the
nitrogen removal process and through the control of these conditions along with
reaction time, the nitrogen removal process can be split up into its component steps.
4.5.2 Shortcut nitrification
The first step of ammonia removal is carried out by Ammonia Oxidising Bacteria
(AOB), these bacteria oxidise ammonia to nitrite, while the second group of
bacteria Nitrite Oxidising Bacteria (NOB) oxidise the nitrite to nitrate. To achieve
complete nitrogen removal it is not necessary to oxidise ammonia all the way
to nitrate, by controlling the process in such a way as to prevent the complete
oxidation of ammonia to nitrate, the intermediate produced nitrite can be
subsequently denitrified to N2 resulting in a saving of 1.17 gO2 per g of N-NH4.
Shortcut nitrification, as it known, uses effective process control to reduce the
growth rate of the NOB and minimise the amount of nitrite converted to nitrate
through a better understanding of the growth rates of the two groups of bacteria
required for the nitrification process and their kinetic parameters such as oxygen
affinity (Ciudad et al. 2006). Trials have shown that process can be applied to
wastewater with high ammonia concentrations for example, leachate (Akerman,
2005). The process has been scaled-up and is available commercially from Veolia
under the name Anita-shunt.
4.5.3 Anammox
The identification of the Anaerobic Ammonia Oxidising group of bacteria
(anammox) and the ability to construct full scale anaerobic ammonia oxidising
reactors further reduces the energy required for nitrogen removal. This anaerobic
ammonia oxidation process has been successfully scaled up by many companies
including Paques and Gronmij. To date there are many different variants of the
anammox technology including suspended culture (flocs) self-supporting biofilm
(granules) and attached biofilm based systems (MBBR).
Although the anammox bacteria do not require Oxygen, some oxygen is
required to produce the Nitrite which the anammox bacteria utilise for the
ammonia oxidation. Therefore through the implementation of anammox the
oxygen requirement for the aerobic treatment is significantly reduced.
4.7 Conclusions
Aerobic biological treatment has formed the corner stone of sewage treatment
for the past 100 years and is likely to continue to form a major part of the sewage
treatment plant of the future. Aerobic biological treatment produces clean
water suitable for discharge to surface waters or further processing into usable
water cheaply and efficiently without the use of chemicals or the production of
more concentrated streams. Despite being an economical process the delivery
of oxygen to the bacteria is still energy intensive and very inefficient, thereby
presenting the opportunity for innovative aerobic biological processes which
treat the sewage more efficiently and economically using less energy. Over the
past 10 years several processes have been developed and are beginning to be seen
at full scale sewage treatment plants. While these innovative processes reduce
operation cost the overall impact on the capital cost on a new build or a upgrade
site has yet determined.
4.8 References
Adams N., Hong Y., Ireland J., Koops G. H., Côté P. (2014). An Innovative Membrane-
Aerated Biofilm Reactor (MABR) for Low Energy Treatment of Municipal Sewage.
Conference Proceedings Singapore International Water Week, Singapore.
Akerman A. (2005). Feasibility of nitrate-shunt (nitritation) on landfill leachate. Tesis
Maestría. Lunds Tekniska Universitet. Departament of Water and Environmental
Engineering. Suecia. 79 pp.
Biogill Technical and Specification Guide. http://www.biogill.com/downloads.
Chun Liu, Hiroshi Tanaka, Jin Ma, Lei Zhang, Jing Zhang, Xia Huang and Yoshiaki
Matsuzawa. (2012). Effect of microbubble and its generation process on mixed liquor
properties of activated sludge using shirasu porous glass (SPG) membrane system.
Water Research, 46(18), 6051–6058.
Chun Liu, Hiroshi Tanaka, Jing Zhang, Lei Zhang, Jingliang Yang, Xia Huang and
Nobuhiko Kubota. (2013). Successful application of shirasu porous glass (SPG)
membrane system for microbubble aeration in a biofilm reactor treating synthetic
wastewater. Separation and Purification Technology, 103, 53–59.
Ciudad, G., Werner, A., Bornhardt, C., Munoz, C., Antileo, C. (2006). Differential kinetics
of ammonia- and nitrite-oxidizing bacteria: a simple kinetic study based on oxygen
affinity and proton release during nitrification. Process Biochemistry, 41, 1764–1772.
Diffusair (2014). Diffusaire, Available at: www.diffusaire.com.
Emefcy-BioEnergy Systems (2014). SABRE-Spiral Aerobic Biofilm Reactor-EMEFCY
Available at: http://www.emefcy.com/product.php?ID=47 (accessed 15 September 2014).
Energy Efficiency in the Water Industry: A Compendium of Best Practices and Case
Studies UKWIR Report REF No 10/CL/11/13.
Eoin Casey, Eoin Syron, John Shanahan and Michael J. Semmens (2008). Comparative
economic analysis of full scale MABR configurations. IWA North American
Membrane Research Conference.
Giesen and Thompson (2013). Aerobic Granular Biomass for Cost-Effective, Energy
Efficient and Sustainable Wastewater Treatment. 7th European Waste Water
Management Conference.
Haney P. D. (1954). Theoretical principles of aeration. Journal American Water Works
Association, 46(4), 353–376. (accessed 15 September 2014)
Ibrahim Gar Al-Almrashed, Ahmed El-Morsy, Mohamed Ayoub (2013). A new approach
for upgrading of sewage treatment plants to accommodate excess organic and hydraulic
loads. Journal of Water Sustainability, 3(3), 153–163, University of Technology
Sydney & Xi’an University of Architecture and Technology.
Issy Caffoor (2008). Environmental Knowledge Transfer Network, Business Case 3: Energy
EfficientWater and Wastewater Treatment – A Priority Technology for the UK.
Kadar Y. and Siboni G. (2006). Optimization of Energy Economy in the Design and
Operation of Wastewater Treatment Plants, Mekorot Water Company, Ltd. and DHV-
MED, Ltd. Tel-Aviv, Israel, 17th Congress of world energy congress, Houston, Texas.
Koichi Terasakan, Ai Hirabayashi, Takanori Nishino, Satoko Fujiokaa, Daisuke Kobayashi
(2011). Development of microbubble aerator for waste water treatment using aerobic
activated sludge. Chemical Engineering Science, 66, 3172–3179.
Leon Downing and Robert Nerenberg (2008). Total nitrogen removal in a hybrid, membrane-
aerated activated sludge process. Water Research, 42(14), 3697–3708.
Lucas Seghezzo, Grietje Zeeman, Jules B. and van Lier H. V. M. (1998). Hamelers,
GatzeLettinga, a review: the anaerobic treatment of sewage in UASB and EGSB
reactors. Bioresource Technology, 65(3), 175–190.
Marco R. Menendez P. E., Black and Veatch (2010). How we use Energy at Wastewater
Plants...and How We Can Use Less. Annual Conference Technical Papers – NC
AWWA-WEA.
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.
Ronen Itzhak Shechter, Lior Eshed, Eytan Baruch Levy, Tamar Ashlagi Amiri. Diffusion
aeration for water and wastewater treatment. US 20120273414 A1, EMEFCY.
5.1 Introduction
The anthropogenic activities of the combustion of fossil fuel, the production of
nitrogen and phosphorus fertilizers, intensive cultivation as well as other actions
have resulted in accelerating the global nitrogen and phosphorus cycles (Vitousek
et al. 1997). As a result, several environmental problems have occurred to aquatic
ecosystems which include groundwater pollution by nitrate, acidification and
cultural eutrophication of water bodies and toxic effects caused to aquatic organisms
by organic and inorganic nitrogen forms (Cervantes, 2009). Consequently, nutrient
removal from sewage is important within wastewater treatment plants (WWTPs)
and can help alleviate the aforementioned environmental problems related to the
discharge of nutrient rich effluents.
Wastewater facilities are increasingly being required to implement treatment
process improvements in order to meet stricter discharge limits with respect to
nitrogen and phosphorus. Wastewater treatment is an energy intensive process,
accounting for approximately 3% of electricity use in developed countries. The
greenhouse gases (GHG) emissions, energy consumption and the cost of wastewater
treatment increase when nutrient removal is implemented (Keller & Hartley, 2003;
USEPA, 2007). Balancing between nutrient removal efficiency and the cost of
treatment is critical in order to implement a solution that will meet the required
limits at an acceptable cost (WERF, 2011).
This chapter reviews the conventional organic carbon and nutrient removal
bioprocesses and the latest technologies for nutrient removal, which include partial
nitritation coupled with the anoxic ammonium oxidation (anammox), nitritation/
denitritation and via nitrite/nitrate phosphorus accumulation. The various biological
nutrient removal (BNR) processes are evaluated based also on the energy aspect.
5.2 ReguLatory Background
The urban wastewater treatment directive (91/271/EEC and its amendment 98/15/
EEC) of the European Union (EU), is an emission oriented regulation which defines
the required treated effluent quality of municipal wastewater discharged into water
bodies in terms of chemical oxygen demand (COD), biochemical oxygen demand
(BOD5), total suspended solids (TSS), nitrogen and phosphorus concentration. The
Directive imposes nutrient limits in the treated effluent that is discharged to sensitive
water bodies subject to eutrophication; these limits are based on agglomeration
size (i.e., population equivalent PE) (Table 5.1). The EU Nitrate Directive (91/676/
EEC) imposed limits to the amount of nitrate that is applied to land within nitrate
vulnerable zones, in order to protect the surface and groundwater.
Table 5.1 EU Nutrient limits imposed for the discharge of treated effluent from
municipal WWTPs into sensitive water recipients (Council Directive 91/271/EEC).a
the WWTPs have to meet the limit of treatment technology (LOT). The LOT is
defined as the lowest economically achievable effluent quality, which for nitrogen
is <1.5 to 3 mgN/L and for phosphorus 0.07 mg/L to 0.1 mgP/L (Barnard, 2006;
Oleszkiewicz & Barnard, 2006).
Table 5.2 Examples of limit values for nutrients in the treated effluent discharged
into water bodies.
5.3 Energy CONSIDERATIONS
The potential energy available in raw wastewater entering a municipal WWTP
exceeds the electricity requirements of the treatment process. Energy contained
in organics entering the plant can be related to the COD load of the influent flow.
This organic load is subjected to aerobic and anaerobic degradation processes,
partly releasing the captured energy. Three different types of energy can be
distinguished: heat, synthesis energy and electricity. Aerobic metabolism yields
a high amount of energy which can hardly be put to good use. Energy is produced
by catabolic biochemical reactions as the substrate is oxidised by bacteria.
The energy is used for the synthesis and maintenance of the cells. Synthesis
energy is thus used by bacteria during their anabolic biochemical reactions in
order to develop new biomass. Synthesis energy is related to high excess sludge
production. The anaerobic digestion process generates much lower synthesis
energy and less heat than aerobic processes, which has much higher impact
because of high concentrations in the solids train. A major part of the energy
content is captured in methane. This amount of energy can be transformed by a
combined heat power (CHP) unit to electrical energy and heat, which can then be
used. These energy products can be recycled to the wastewater treatment line and
thus provide energy for the aeration system and heat the digesters. The process
should divert as much organics as possible, from the wastewater treatment line
to the anaerobic solids train in order to increase as much as possible energy
recovery. However, compliance with nutrient removal requirements remains the
overriding objective.
Aerobic conditions
VFAs
Energy
Poly-P
PHB CO2+
Poly-P PHB H2O
Energy
PO43- O2
PO43- Anaerobic conditions H2O
Figure 5.3 Attached and suspended biomass configurations which can be applied
to remove nitrogen (SND = simultaneous nitrification/denitrification, C = dosing of
external organic carbon).
Several configurations for the EBPR have been invented, developed and
implemented (Figure 5.4). In practice, the EBPR is successfully implemented when
the alternation of anaerobic/aerobic is ensured. Any bioprocess in which the nitrate/
nitrite is prevented from entering the anaerobic bioreactor and an adequate readily
biodegradable COD is provided during the anaerobic phase can accomplish EBPR
(Barnard, 2006). Therefore, EBPR can be integrated within any of the configurations
described above for nitrogen removal, provided a dedicated anaerobic phase is
established. The phosphorus removal configurations shown in Figure 5.4 place the
anaerobic reactor upstream of the biological process. This has the advantage that
volatile fatty acids present in sewage can be used by the phosphorus accumulating
organisms (PAOs). The simplest configuration is the Phoredox (A/O) process with
one anaerobic reactor placed before the aerobic one. In the SBR process, EBPR can
be accomplished by including an anaerobic phase just after the filling and before the
aerobic reaction. Thus, the sequence of anaerobic, aerobic, anoxic phase in an SBR
can successfully remove phosphorus and nitrogen from wastewater. The two and
four stage Bardenpho processes can be modified to remove phosphorus by inserting
an anaerobic reactor upstream. These configurations are known as the three and five
stage modified Bardenpho processes (Wentzel et al. 2008).
In the University of Cape Town (UCT) process and its modified version (which has
two anoxic tanks), an internal recycle of nitrate is carried out from the aerobic tank to
the anoxic one in order to achieve effective denitrification. Furthermore, to minimize
the entrance of nitrates in the anaerobic zone, the mixed liquor is recycled from the
anoxic tank to the anaerobic one and the settled sludge is returned to the anoxic tank.
By manipulating the nitrate recycle ratio, the nitrate concentration in the anoxic tank
can drop to zero. As a result, the anoxic recycle will not introduce any nitrate into
the anaerobic tank. The modified version includes two anoxic tanks in order to have
better control of the recycle. In the Johannesburg configuration, the elimination of
nitrates in the mixed liquor recycle to the anaerobic tank is accomplished by a second
anoxic tank which removes the nitrate from the recycle. The Biodenipho/Biodenitro
process is a phased isolation ditch system where nitrogen removal is accomplished by
the phased aerobic/anoxic reactors which are interchanged. The anaerobic conditions
are accomplished upstream in a separate tank (Wentzel et al. 2008).
fermentation of primary sludge to produce short chain fatty acids that can be applied
as carbon source to the BNR process is increasingly being applied in WWTPs in
Europe. MBBR and IFAS plants are often implemented when a small footprint is
required (Oleszkiewicz & Barnard, 2006). These processes use suspended carriers
onto with the biofilm develops and are compact leading to significant reduction in
the bioreactor volumes.
Many WWPTs in Central North America still practice only carbon removal or
have recently been upgraded for BNR as a response to the imposed regulations.
The tendency for upgrading is to conduct chemical precipitation of phosphorus
and extend the aerobic phase in order to accomplish complete ammonium
oxidation. In WWTPs where a low total nitrogen concentration limit must be
met, post-denitrification using denitrifying filters is usually carried out (de
Barbadillo et al. 2005). In several US plants the activated sludge BNR system is
increasingly being adopted either as Bardenpho-like process or as MLE coupled
with chemical precipitation. In western Canada the Westbank process being the
most popular process. In Eastern Canada the high rate activated sludge process
is typically used together with chemical precipitation. Since denitrification
is not practiced, the mixed liquor contains nitrates and thus EBPR cannot be
carried out. Finally, in Quebec attached growth processes (BAF) are usually
employed (Oleszkiewicz & Barnard, 2006). In the last several years, nitritation/
denitritation and partial nitritation with the anoxic ammonium oxidation (i.e.,
the completely autotrophic nitrogen removal process) are being implemented as
alternative low cost treatment options for nitrogen removal. These processes are
further discussed in Section 5.5.
Pre-treatment, 1%
Disinfec on, 4%
Other, 5%
Dewatering, 9% Primary Sedimenta on, 2%
Pumping, 9%
Secondary
sedimenta on, 2%
Table 5.3 Costs for upgrading WWTPs with BNR for Maryland (US EPA, 2007).
(Continued)
Table 5.3 Costs for upgrading WWTPswith BNRfor Maryland (US EPA, 2007)
(Continued).
Table 5.4 Costs for upgrading WWTPs with BNR for Connecticut (US EPA,
2007).
Table 5.5 Average BNR Costs for new small scale WWTPs (US EPA, 2007).
the denitrification process (Mayer et al. 2009; Ji & Chen, 2010; Longo et al. 2015).
Short chain carbon sources produced from fermentation can promote nitritation/
denitritation and denitrifying phosphorus removal via nitrite from the anaerobic
supernatant (Frison et al. 2013a).
Table 5.6 Comparison of the conventional BNR with the advanced BNR processes.
Enhanced biological removal via nitrite can also be integrated with nitritation/
denitritation in order to remove phosphorus biologically. Studies have documented
the occurrence of denitrifying phosphorus accumulating organisms (DPAOs) that
can utilize nitrate or nitrite as electron acceptors instead of oxygen (Carvalho
et al. 2007). The denitrifying phosphorus uptake can reduce the requirements for
5.7 Conclusion
Several configurations are currently applied for BNR in full scale WWTPs. Both
attached and suspended growth processes are implemented. The adoption of the
most appropriate BNR scheme is closely related to the treated effluent requirements
for phosphorus and nitrogen. In cases where treated effluents are discharged to
sensitive water bodies, the nutrient limits can be very low and the technology
selection is critical. Nitritation/denitritation and the completely autotrophic nitrogen
removal have emerged as alternative BNR processes compared to the conventional
nitrification/denitrification. Such processes have the advantages of lower aeration
requirements, resulting in lower operating expenses. They are increasingly being
adopted for the treatment of the nitrogenous sludge reject water. In BNR processes
it is also important to consider the GHG emissions and particularly nitrous oxide
emissions. The aforementioned innovative processes should be implemented with
care in order not to increase the overall carbon footprint of the WWTP.
5.8 ACKNOWLEDGEMENT
This work was supported by the Marie Curie FP7-PEOPLE-2012 project with
title Low environmental footprint biological treatment processes for waste and
wastewater treatment, LEF-BIOWASTE (Number 322333).
5.9 References
Ahn J. H., Kim S., Park H., Rahm B., Pagilla K. and Chandran K. (2010). N2O emissions
from activated sludge processes, 2008–2009: results of a national monitoring survey
in the United States. Environmental Science and Technology, 44, 4505–4511.
Ahn J. H., Kwan T. and Chandran K. (2011). Comparison of partial and full nitrification
processes applied for treating high-strength nitrogen wastewaters: microbial ecology
through nitrous oxide production. Environmental Science and Technology, 45, 2734–2740.
Anthonisen A. C., Loehr R. C., Prakasam T. B. and Srinath E. G. (1976). Inhibition of nitrification
by ammonia and nitrous acid. Water Pollution Control Federation, 48, 835–852.
Barnard J. L. (2006). Biological nutrient removal: where we have been, where we are going?
Water Environment Federation’s Technical Exhibition and Conference (WEFTEC),
Dallas, US.
Blackburne R., Yuan Z. and Keller J. (2008a). Partial nitrification to nitrite using low dissolved
oxygen concentration as the main selection factor. Biodegradation, 19, 303–312.
Blackburne R., Yuan Z. Q. and Keller J. (2008b). Demonstration of nitrogen removal via
nitrite in a sequencing batch reactor treating domestic wastewater. Water Research,
42, 2166–2176.
Carvalho G., Lemos P. C., Oehmen A. and Reis M. A. M. (2007). Denitrifying phosphorus
removal: linking the process performance with the microbial community structure.
Water Research, 41, 4383–4396.
Cervantes F. J. (2009). Anthropogenic Sources of N-Pollutants and their Impact on the
Environment and on Public Health, Chapter 1. In: Environmental Technologies to
Treat Nitrogen Pollution, F. J. Cervantes (ed.), IWA publishing, London, UK.
Commission Directive 98/15/EC of 27 February 199 Amending Council Directive 91/271/
EEC with Respect to Certain Requirements Established in Annex I there of (Text with
EEA relevance).
Water Environment Federation (WEF) and American Society of Civil Engineers (ASCE)/
Environmental and Water Resources Institute (EWRI) 2006. Biological Nutrient
Removal (BNR) Operation in Wastewater Treatment Plants. McGraw Hill, New York.
Wentzel M. C., Comeau Y., Ekama G. A., van Loosdrecht M. C. M. and Brdjanovic D.
(2008). Chapter 7: Phoshporus removal. In: Biological Wastewater Treatment:
Principles Modelling and Design, M. Henze, M. van Loosdrecht, G. Ekama and D.
Brdjanovic (eds), IWA, UK.
WERF (2011). When do the costs of wastewater treatment out weigh the benefits of nutrient
removal? Striking the Balance Between Nutrient Removal in Wastewater Treatment
and Sustainability (NUTR1R06n). www.werf.org/a/ka/Search/ResearchProfile.
aspx?ReportId=NUTR1R06n (accessed 10 March 2014).
Wett B., Buchauer K. and Fimml C. (2007). Energy Self-Sufficiency as a Feasible Concept
for Wastewater Treatment Systems. Proceeding of IWA Leading Edge Technology
Conference, Singapore, Asian Water, 21–24 September. http://www.araconsult.at/
site_04en.html (accessed 5 May 2014).
Yang Q., Peng Y. Z., Liu X. H., Zeng W., Mino T. and Satoh H. (2007). Nitrogen removal
via nitrite from municipal wastewater at low temperatures using real-time control
to optimize nitrifying communities. Environmental Science and Technology, 41,
8159–8164.
6.1 Introduction
It has already been more than 20 years since the first studies were performed
in the field of aerobic granular biomass (Mishima & Nakamura, 1991). By that
time, aerobic granules were developed in a UASB type reactor using pure oxygen.
However, this work was not really appreciated at that moment. It was only in the
late 90’s when the interest on the aerobic granular systems appeared again. Initially
only organic matter was removed (Morgenroth et al. 1997; Dangcong et al. 1999;
Beun et al. 1999, 2002) and later research works indicated that the simultaneous
removal of organic matter, nitrogen and eventually phosphorus compounds
was feasible in the same unit (Arrojo et al. 2004; de Kreuk et al. 2005a; Bassin
et al. 2012; Isanta et al. 2012). The interest for this kind of technology increased
when its advantages compared to the conventional activated sludge systems were
identified: excellent biomass settling properties, large biomass retention, the
ability to withstand shock and toxic loadings, the presence of aerobic and anoxic
zones inside the granules to perform different biological processes, and so on.
(Morgenroth et al. 1997; Dangcong et al. 1999; Beun et al. 1999; Peng et al. 1999;
Tay et al. 2001a; Lin et al. 2003; Liu et al. 2003; Yang et al. 2003; Campos et al.
2009; Khan et al. 2013).
Due to this great interest the ‘1st IWA-Workshop Aerobic Granular Sludge’ took
place in Munich in 2004, where a definition for aerobic granules was stated as
(de Kreuk et al. 2005b):
“Granules making up aerobic granular activated sludge are to be understood as
aggregates of microbial origin, which do not coagulate under reduced hydrodynamic
shear, and which settle significantly faster than activated sludge flocks.”
Furthermore, aerobic granules should fulfil the following requirements: (i) the
values of the sludge volume index after 10 and 30 minutes of settling (SVI10 and
SVI30, respectively) do not differ more than a 10% (Schwarzenbeck et al. 2004a);
(ii) the position of microorganisms inside the granules is fixed and it does not
change due to the existence of a matrix of biomass and exopolymeric substances
(EPS); (iii) no carrier material is intentionally involved or added; (iv) the minimum
size of the granules is considered to be around 0.2 mm (de Kreuk et al. 2005b); (v)
it is a general convention to consider those bacterial aggregates with an SVI10 value
equal or lower than 60–70 mL/g TSS as to be aerobic granules.
Nowadays, much research work has been performed producing significant
knowledge. From this knowledge, some patents belonging to different countries about
aerobic granular systems have been elaborated and some full-scale plants are running
in different countries. However, to the authors’ knowledge, regarding this full-scale
application, only one of all the technologies is a reality at the moment, the Nereda®
one. In the following sections the gathered knowledge up to date in the field of aerobic
granulation has been summarised. Information related to the full-scale implementation
of the aerobic granular systems is also provided, along with information regarding the
advantages of this technology compared to the conventional ones.
6.2.1.1 Feast–famine regime
Conventional activated sludge systems are facilities operated in continuous
mode with biomass grown in suspension. The concentration of organic matter in
the liquid media of these systems is always low. When high loaded wastewater
enters the system, problems related to the dissolved oxygen concentration and,
consequently, with biomass floatation frequently appear. Under these conditions,
to produce aggregated biomass like aerobic granules is not possible. In order to
produce the granular biomass in aerobic conditions, the heterotrophic biomass
must be cyclically subjected to periods of availability (feast phase) and absence
(famine phase) of organic substrate in the liquid phase. This is achieved if the fed
organic matter is supplied to the system, operated in sequencing mode, within a
very short time period (Figure 6.1a). In this way, the organic matter concentration,
after feeding addition, is increased, leading to high concentration gradients. The
dependence of the biomass aggregation capacity on the organic matter availability
in the liquid is explained by Chudoba’s theory of the kinetic selection (Chudoba
et al. 1994).
Feeding Aeration
Anaerobic feeding Settling Withdrawal
Figure 6.1 Distribution of the phases within a cycle of operation of aerobic granular
systems in the case of simultaneous (a) organic matter and nitrogen removal and
(b) organic matter, nitrogen and phosphorous removal.
This theory states that, due to the higher maximum specific growth rate (μmax)
and half saturation constant (Ks) values (according to Monod kinetics), the growth
of the granule forming microorganisms is favoured at high substrate concentrations.
When the substrate concentration is low, the filamentous microorganisms, with
low half saturation constant, grow at higher rates than the granule forming
organisms and the settling properties of the sludge are deteriorated.
Furthermore, it has been observed that, during the feast phase, the organic
biodegradable substrates present in the wastewater are stored inside the biomass in
the form of PHA, while, during the famine phase, the PHA are used for biomass
growth (Beun et al. 2002; Val del Río et al. 2013). Under these conditions, when
stable granules are formed in the system, organic matter and nitrogen can be
removed simultaneously.
The aforementioned procedure can also be applied to produce granular
biomass when the system is operated under anaerobic-aerobic alternating
conditions (Figure 6.1b). In this alternative option, the granulation occurs due
to the development of slow growing organisms like the phosphorous removing
bacteria (de Kreuk et al. 2004, 2005a). In this case, the feast phase occurs under
anaerobic conditions, when the phosphorous is released from the cells while
PHA is accumulated, and the famine period takes place under aerobic conditions,
where the phosphate is uptaken at the expense of the stored PHA. As in the first
option, the feast-famine alternating conditions function as the selective pressure
to favour the development of phosphorous removing bacteria, which are slow
growing organisms known to form aggregates like granular biomass (de Kreuk &
van Loosdrecht, 2004). In this case organic matter, nitrogen and phosphorous are
removed in the unit.
75 mL/g TSS in the case of aerobic granules compared to the 100–200 mL/g
TSS obtained in the case of biomass from conventional activated sludge systems.
These low values allow the achievement of high concentrations of compact
biomass inside the reactors and facilitate the separation of the effluent after
settling.
FEEDING
Influent
AERATION
IDLE
Air
Effluent
WITHDRAWAL SETTLING
However, to take benefit from the advantages of a SBR system, some special features
should be considered in the design stage of the full scale plants such as the fact that SBR
implies higher costs in control, data acquisition and instrumentation in comparison
to continuous processes due to its high level of sophistication. Furthermore, the use
of several SBRs in parallel is necessary to cope with high flow rates. A previous
tank is usually installed to homogenise the peaks of flow to be treated along the
working day. Although this homogenisation tank presents the advantage to avoid high
fluctuations in the influent composition of the SBR, it also means an increase in the
installation and control system costs. Another point to have in mind is the quality of
the produced effluent which depends on the settling properties of the granular sludge.
The detriment of the settling properties can occur if the denitrification process takes
place during the sedimentation phase, with the consequent production of nitrogen gas
which causes the sludge bed flotation. To finalize, an equalization step after the SBR
might be a potential requirement depending on the downstream processes.
6.2.3.1 Substrate composition
Aerobic granules can be produced in SBR systems when the readily biodegradable
organic matter content of the treated wastewater is larger than 75 mg COD/L
(Mosquera-Corral et al. 2011). Lower concentrations do not provide the minimum
concentration gradient conditions for the establishment of the required feast-famine
regime. Taking into account that the aerobic granules are a special kind of ‘biofilm’,
the knowledge obtained regarding the effect of the organic matter composition on
the biofilm formation and characteristics could be applied in the case of the aerobic
granule formation. In this sense, previous works on biofilms have demonstrated
the relationship between the degree of reduction of the substrate and the biofilm
density (Mosquera-Corral et al. 2003). For example, acetate results in high
maximum biomass growth rates (µmax = 9.6 d−1) and produce less dense biofilms
(18 g VSS/Lbiofilm) than methanol resulting in slow maximum biomass growth rates
(µmax = 5.26 d−1) and high biofilm density (80 g VSS/Lbiofilm). In the case of aerobic
granules it is necessary to take into account that the organic compounds present
in the wastewater are first transformed into storage compounds (PHA), during the
feast period, and these are latter used for growth during the famine period. In these
conditions, slow growing microorganisms are developed because the biomass growth
rate using an intracellular polymer (PHA) is slower than using a simple extracellular
compound. Furthermore, some authors have proposed to operate SBR systems to
select for Poly-P accumulating bacteria (PAO) and glycogen accumulation organisms
(GAO) which are known to grow at slow rates (de Kreuk & van Loosdrecht, 2004).
From a microscopic point of view, previous works have also indicated that
substrates containing carbohydrates favour the uncontrolled growth of filamentous
bacteria (Schwarzenbeck et al. 2005), while acetate is more suitable to produce aerobic
granules. In general, simple substrates are better than complex ones because are easy
to store by the biomass. Aerobic granules have been successfully cultivated on a wide
variety of substrates (Figure 6.3) including glucose, acetate, ethanol, peptone, phenol
and industrial wastewater (Morgenroth et al. 1997; Beun et al. 1999; Yi et al. 2003;
Arrojo et al. 2004; Schwarzenbeck et al. 2004b; de Kreuk & van Loosdrecht, 2006;
Sun et al. 2006; Figueroa et al. 2008; Ho et al. 2010; Val del Río et al. 2012a).
Figure 6.3 Aerobic granules formed in SBRs fed with acetate solution (a), pig
manure (b) and effluent from the seafood industry (c). The bar in the images
corresponds to 2 mm.
6.2.3.3 Exopolymeric substances
EPS produced in the granules include different kinds of compounds in variable
amounts comprising proteins, polysaccharides, lipids, glycoproteins, humic-like
substances, nucleic acids, among others. These compounds are produced under
different conditions such as substrate or oxygen limiting conditions, ionic strength,
temperatures of operation and so on. There are conflicting results reported in
literature regarding the function of the EPS on the formation of aerobic granules
(de Kreuk et al. 2005b). It is believed that one of the EPS functions is to act as
a ‘glue’ among the microorganisms present in an aggregate due to its physical
properties (Tay et al. 2001a). Moreover, according to previous research, the EPS
content increases with granulation, there are differences in loosely bound and
tightly bound EPS and, within the granule structure, insoluble versus soluble
polysaccharide gradients occur.
The polysaccharides have been identified as the major compounds forming
the granular EPS and the gel-like structures responsible for bacterial granule
formation (Lin et al. 2010). Furthermore, two different exoplysaccharides have
been identified by isolation techniques as the responsible for the gel-forming
matrix of the granules: alginate-like polysaccharide (Lin et al. 2010) and granulan
(Seviour et al. 2011). At the moment it is still not clear which EPS is responsible
for the formation of aerobic granules and, even new compounds could arise as
important components of these structures (Seviour et al. 2012).
that the addition of Ca2+ accelerated the aerobic granulation process. In the case
of other cations, such as Mg2+, it was demonstrated that the presence of this ion in
the feeding media enhanced the sludge granulation process in sequencing batch
reactors (Li et al. 2009). The external addition of these ions should not be an
option, since they are often already present in the wastewater and may have a
positive role during the granulation process.
FEAST FAMINE
AEROBIC
CO2 O2 Organic AEROBIC
ANOXIC source
PHA Bacteria
ANOXIC
ANAEROBIC O2
NO2- O2 ANAEROBIC
PHA
O2 NO3-
PHA NH4+ O2
CO2
NO3-
Figure 6.4 Transformation of the organic matter and nitrogen inside a granule
during the feast and famine periods.
phase surrounding the granules. The nitrate produced is removed during feast
phase of the next operational cycle.
In those cases where phosphate is simultaneously removed, the distribution of
the different processes is shown in Figure 6.5. During the feast phase, anaerobic
conditions are imposed to the granules, and organic matter (volatile fatty acids,
VFA) is uptaken to accumulate PHA, while the Poly-P is hydrolysed as phosphate
to the liquid medium to provide the energy needed for PHA accumulation (de
Kreuk et al. 2005a). If the organic matter is in excess, glycogen accumulating
organisms (GAO) will use the VFA left to produce PHA consuming the already
accumulated glycogen produced for energy production (not shown in the Figure
6.5 to simplify the process). During the famine phase, the reverse process takes
place: the stored PHA is consumed for biomass growth (or for eventual glycogen
production) and the released energy is stored in the phosphate bonds during the
poly-P formation. If nitrification also takes place during this period, the stored
PHA is used for denitrification.
Organic Biomass
NO3-
source PHA
Biomass
O2 PHA
PO4-
PO4-
Figure 6.5 Transformation of the organic matter, nitrogen and phosphorous inside
a granule during the feast and famine periods.
while required surface for implantation decreases to 25% and energy consumption
to 65–75% (de Bruin et al. 2004). These preliminary calculations indicated that the
annual costs of the granular SBR including primary and post treatment are 17%
lower while mechanical/electrical works account for 40–45% of the capital costs.
Further studies in pilot plant confirmed some of the previous estimations. Di
Iaconi et al. (2010a, b) determined a biomass productivity of 0.04–0.14 kg TSS/kg
CODremoved which represent 74% less sludge production than conventional activated
sludge systems (0.4–0.6 kg TSS/kg CODremoved).
Regarding energy associated to aeration, data collected from pilot-plants
operation do not provide reliable information due to the highly oversized air flows
as shown in Table 6.1. Information gathered indicates that pilot plants operate at
shorter HRT, around 6 hours, and higher applied organic loading rates (OLR) as
high as 9 kg COD/m3 ⋅ d much higher than activated sludge systems.
With regards to CAPEX and OPEX of industrial scale plants not much
information is available as in the case of activated sludge due to the novel application
of this technology (Table 6.2).
Table 6.2 Comparison of aerobic granular pilot and industrial scale plants to
activated sludge systems.
Pilot plants
Standard Dutch – – 7–17 – [1]
wastewater
composition
Primary effluent 0.8 – – 74 [2]
from a municipal
WWTP
Primary effluent 0.8 – 60a 87a [3]
from a tannery
WWTP
Indrustrial installations
Gansbaai Municipal + 3 × 1600 20 50 – [4]
Industrial
Epe Municipal + 3 × 4500 25 25 –
Industrial
Frielas Municipal + 1000 – 30b/50c –
Industrial
Wemmershoek 2 × 1800 – 50/75c –
Municipal
aAfter ozonation and compared to Activated Sludge+ Fenton process. bElectricity for aeration. cTotal
electricity consumption. References: [1] De Bruin et al. (2004); [2] Di Iaconi et al. (2010a); [3] Di Iaconi
et al. (2010b); [4] Giesen and Thompson (2013).
Regarding the performance some aspects have to be taken into account different
from activated sludge systems. During the start-up, large quantities of the added
inoculum are washed out from the reactor, temporarily decreasing the quality of the
produced effluent. This can be avoided by inoculating the reactor with previously
developed granular biomass (Liu et al. 2005a) or with a mixture of crushed aerobic
granules and floccular sludge (Pijuan et al. 2011).
The solid content of the effluent has to be reduced previously to its discharge to
natural water bodies by means of filtration systems (membrane systems, settlers,
sand filters, etc.).
Another aspect to take into account during operation is the aeration cost.
This is relatively high due to the need of high air flows to keep the required DO
concentration and the mixing. The air requirements can be reduced by the use
of slow-growing microorganisms, such as the nitrifiers or phosphorous removing
bacteria (de Kreuk & van Loosdrecht, 2004).
The first pilot plant based on aerobic granular biomass was started up in the
Netherlands in October 2003 to treat urban wastewater (de Bruin et al. 2004,
2005), this was the origin of the Nereda® process (Nereda® 2013). Since 2005,
over 10 full-scale aerobic granular sludge technology (AGS) systems had been
implemented in the Netherlands, Portugal and South Africa, for the treatment of
both industrial and municipal wastewater based on the Nereda® process.
In Gansbaai (South Africa) a demonstration plant, comprising three SBRs of
1600 m3, was constructed in 2006, to minimize the risks of implementation of
the technology, for the treatment of sewage and industrial wastewater to handle
capacities up to 5000 m3/day.
The Nereda® process was also implemented as a demonstration plant in a WWTP
in Portugal in 2012 (Frielas WWTP). The plant has a capacity of 70,000 m3/day
for the treatment of domestic wastewater from 250,000 inhabitants. It consists of a
conventional activated sludge system with six complete mix biological reactors and
12 settlers. One of them was retrofitted into a Nereda® pilot reactor with a volume
around 1000 m3, that is working in parallel with the remaining five activated
sludge reactors. The granular excess sludge from the Nereda® reactor is pumped
to the activated sludge lines, improving the sludge characteristics and settling
performances of the existing activated sludge plant.
After the success of the Gansbaai WWTP, the first municipal full-scale installation
in the Netherlands was constructed in Epe to treat a flow rate of 1500 m3/h. This
plant consisting of three SBRs of 4500 m3 was inaugurated in May 2012. Nowadays,
three new plants are being started up in the Netherlands with this technology
(Garmerwolde, Vroomshoop and Dinxperlo WWTPs). Also a second Nereda®
plant with a capacity of 5000 m3/day is under construction in Wemmershoek (South
Africa). The design equivalent inhabitants of these plants ranged from 15,700 and
140,000 and the COD concentrations from 500 to 875 mg COD/L.
Furthermore, there are about 20 new plants scheduled to be built in different
countries including Australia, China, Brazil, India, the Middle-East, Belgium, UK,
Poland, Ireland and the USA.
Besides the application to treat municipal wastewater, a number of Nereda®
plants had been constructed in the Netherlands to treat industrial wastewater from:
cheese speciality industry (Ede), edible oil company (Rotterdam), food industry
(Oosterwolde) and pharmaceutical industry. In this case the organic matter
concentration ranged from 1200 to 10,000 mg COD/L.
With respect to the operational conditions these plants can cope with organic
loading rates as high as 10 kg COD/m3 ⋅ d similar to those obtained from pilot
scale experiments (Table 6.1) at HRTs as short as 0.4 days. Shorter HRTs of 0.08
days are applied at Vroomshoop WWTP treating municipal wastewater with a
COD concentration of 800 mg COD/L.
Although Nereda® process is the clearest example of AGS technology application,
other works were also performed at pilot scale. Tay et al. (2005) operated a pilot plant
for the development of aerobic granular biomass with a height of 1.6 m and a diameter
of 0.19 m (working volume of 34 L). These authors treated a synthetic effluent. Inizan
et al. (2005) performed experiments at pilot scale with a reactor of 1.8 m of height
and 0.2 m of diameter (working volume of 40 L) and using either a synthetic medium
6.5 Acknowledgements
This work was funded by the Spanish Government through the NOVEDAR_
Consolider (CSD2007-00055) and Plasticwater (CTQ2011-22675) projects. Anuska
Mosquera-Corral, Ángeles Val del Río, Ramón Méndez and José Luis Campos
belong to the Competitive Group of Reference GRC2013/032, Xunta de Galicia.
6.6 References
Adav S., Lee D. J. and Lai J.-Y. (2010). Potential cause of aerobic granular sludge
breakdown at high organic loading rates. Applied Microbiology and Biotechnology,
85(5), 1601–1610.
Arrojo B., Mosquera-Corral A., Garrido J. M. and Méndez R. (2004). Aerobic granulation with
industrial wastewater in sequencing batch reactors. Water Research, 38(14–15), 3389–3399.
Bassin J. P., Winkler M. K. H., Kleerebezem R., Dezotti M. and van Loosdrecht M. C. M.
(2012). Improved phosphate removal by selective sludge discharge in aerobic granular
sludge reactors. Biotechnology and Bioengineering, 109(8), 1919–1928.
Beun J. J., Hendriks A., van Loosdrecht M. C. M., Morgenroth E., Wilderer P. A. and
Heijnen J. J. (1999). Aerobic granulation in a sequencing batch reactor. Water Research,
33(10), 2283–2290.
Beun J. J., van Loosdrecht M. C. M. and Heijnen J. J. (2001). N-removal in a granular sludge
sequencing batch airlift reactor. Biotechnology and Bioengineering, 75(1), 82–92.
Beun J. J., van Loosdrecht M. C. M. and Heijnen J. J. (2002). Aerobic granulation in a
sequencing batch airlift reactor. Water Research, 36(3), 702–712.
Campos J. L., Mendez R. and Lema J. M. (2000). Operation of a nitrifying activated
sludge airlift (NASA) reactor without biomass carrier. Water Science and Technology,
41(4–5), 113–120.
Campos J. L., Figueroa M., Mosquera-Corral A. and Mendez R. (2009). Aerobic sludge
granulation: state-of-the-art. International Journal of Environmental Engineering,
1(2), 136–151.
Chudoba J. and Pujol R. (1994). Kinetic selection of microorganisms by means of a
selector – twenty years of progress: history, practice and problems. Water Science and
Technology, 29(7), 177–180.
Dangcong P., Bernet N., Delgenes J.-P. and Moletta R. (1999). Aerobic granular sludge-a
case report. Water Research, 33(3), 890–893.
de Bruin L. M. M., de Kreuk M. K., van der Roest H. F. R., Uijterlinde C. and van
Loosdrecht M. C. M. (2004). Aerobic granular sludge technology: an alternative to
activated sludge? Water Science and Technology, 49(11–12), 1–7.
de Bruin L. M. M., van der Roest H. F., de Kreuk M. K. and van Loosdrecht M. C. M.
(2005). Promising results pilot research aerobic granular sludge technology at WWTP
ede. In: Aerobic Granular Sludge. Water and Environmental Management Series, IWA
Publishing, Munich, pp. 135–142.
de Kreuk M. K. and van Loosdrecht M. C. M. (2004). Selection of slow growing organisms
as a means for improving aerobic granular sludge stability. Water Science and
Technology, 49(11–12), 9–17.
de Kreuk M. K. and van Loosdrecht M. C. M. (2006). Formation of aerobic granules with
domestic sewage. Journal of Environment Engineering, 132(6), 694–697.
de Kreuk M., Heijnen J. J., and van Loosdrecht M. C. M. (2005a). Simultaneous COD,
nitrogen, and phosphate removal by aerobic granular sludge. Biotechnology and
Bioengineering, 90(6), 761–769.
de Kreuk M. K., McSwain B. S., Bathe S., Tay S. T. L., Schwarzenbeck N. and Wilderer P. A.
(2005b). Discussion outcomes. In: Aerobic Granular Sludge, Water and Environmental
Management Series, S. Bathe, M. K. de Kreuk, B. McSwain and N. Schwarzenbeck
(eds), IWA Publishing, Munich, pp. 165–169.
Di Iaconi C., Ramadori R., Lopez A. and Passino R. (2004). Hydraulic shear stress calculation
in a sequencing batch biofilm reactor with granular biomass. Environmental Science
and Technology, 39(3), 889–894.
Di Iaconi C., Ramadori R., López A. and Passino R. (2006). Influence of hydrodynamic
shear force on properties of granular biomass in a sequencing batch biofilter reactor.
Biochemical Engineering Journal, 30(2), 152–157.
Di Iaconi C., De Sanctis M., Rossetti S. and Ramadori R. (2008). Technological transfer
to demonstrative scale of sequencing batch biofilter granular reactor (SBBGR)
technology for municipal and industrial wastewater treatment. Water Science and
Technology, 58(2), 367–372.
Di Iaconi C., Moro G. D., Pagano M. and Ramadori R. (2009). Municipal landfill leachate
treatment by SBBGR technology. International Journal of Environment and Waste
Management, 4(3–4), 422–432.
Di Iaconi C., De Sanctis M., Rossetti S. and Ramadori R. (2010a). SBBGR technology for
minimising excess sludge production in biological processes. Water Research, 44(6),
1825–1832.
Di Iaconi C., Del Moro G., De Sanctis M. and Rossetti S. (2010b). A chemically enhanced
biological process for lowering operative costs and solid residues of industrial
recalcitrant wastewater treatment. Water Research, 44(12), 3635–3644.
Figueroa M., Mosquera-Corral A., Campos J. L. and Mendez R. (2008). Treatment of saline
wastewater in SBR aerobic granular reactors. Water Science and Technology, 58(2),
479–485.
Giesen A., van Loosdrecht M. C. M., de Bruin B., van der Roest H. and Pronk M.
(2013). Full-scale experiences with aerobic granular biomass technology for
treatment of urban and industrial wastewater. International Water Week Amsterdam,
November.
Ho K. L., Chen Y. Y., Lin B. and Lee D. J. (2010). Degrading high-strength phenol
using aerobic granular sludge. Applied Microbiology and Biotechnology, 85(6),
2009–2015.
Inizan M., Freval A., Cigana J. and Meinhold J. (2005). Aerobic granulation in a sequencing
batch reactor (SBR) for industrial wastewater treatment. Water Science and Technology,
52(10–11), 335–343.
Isanta E., Suárez-Ojeda M. E., Val del Río Á., Morales N., Pérez J. and Carrera J. (2012).
Long term operation of a granular sequencing batch reactor at pilot scale treating a
low-strength wastewater. Chemical Engineering Journal, 198–199(0), 163–170.
Jiang H. L., Tay J. H., Liu Y. and Tay S. T. L. (2003). Ca2+ augmentation for enhancement
of aerobically grown microbial granules in sludge blanket reactors. Biotechnology
Letters, 25(2), 95–99.
Jungles M. K., Figueroa M., Morales N., Val del Río Á., da Costa R. H. R., Campos J. L.,
Mosquera-Corral A. and Méndez R. (2011). Start up of a pilot scale aerobic granular
reactor for organic matter and nitrogen removal. Journal of Chemical Technology and
Biotechnology, 86(5), 763–768.
Khan M. Z., Mondal P. K. and Sabir S. (2013). Aerobic granulation for wastewater
bioremediation: a review. Canadian Journal of Chemical Engineering, 91(6),
1045–1058.
Li X.-M., Liu Q.-Q., Yang Q., Guo L., Zeng G.-M., Hu J.-M. and Zheng W. (2009). Enhanced
aerobic sludge granulation in sequencing batch reactor by Mg2+ augmentation.
Bioresource Technology, 100(1), 64–67.
Lin Y. M., Liu Y. and Tay J. H. (2003). Development and characteristics of phosphorous-
accumulating granules in sequencing batch reactor. Applied Microbiology and
Biotechnology, 62(4), 430–435.
Lin Y., de Kreuk M., van Loosdrecht M. C. M. and Adin A. (2010). Characterization of
alginate-like exopolysaccharides isolated from aerobic granular sludge in pilot-plant.
Water Research, 44(11), 3355–3364.
Liu Y., Xu H. L., Yang S. F. and Tay J. H. (2003). Mechanisms and models for anaerobic in
upflow anaerobic sludge blanket reactor. Water Research, 37(3), 661–673.
Liu Q. S., Liu Y., Tay S. T. L., Show K. Y., Ivanov V., Benjamin M. and Tay J. H.
(2005a). Start-up of pilot-scale aerobic granular sludge reactor by stored granules.
Environmental Technology, 26(12), 1363–1369.
Liu Y., Wang Z. W. and Tay J. H. (2005b). A unified theory for upscaling aerobic granular
sludge sequencing batch reactors. Biotechnology Advances, 23(5), 335–344.
Liu Y-Q. and Tay J-H. (2006). Variable aeration in sequencing batch reactor with aerobic
granular sludge. Journal of Biotechnology, 124(2), 338–346.
Mishima K. and Nakamura M. (1991). Self-immobilization of aerobic activated sludge – a
pilot study of the aerobic upflow sludge blanket process in municipal sewage treatment.
Water Science and Technology, 23(4–6), 981–990.
Morales N., Figueroa M., Mosquera-Corral A., Campos J. L. and Méndez R. (2012). Aerobic
granular-type biomass development in a continuous stirred tank reactor. Separation
and Purification Technology, 89(0), 199–205.
Morales N., Figueroa M., Fra-Vázquez A., Val del Río A., Campos J. L., Mosquera-Corral
A. and Méndez R. (2013). Operation of an aerobic granular pilot scale SBR plant to
treat swine slurry. Process Biochemistry, 48(8), 1216–1221.
Morgenroth E., Sherden T., van Loosdrecht M. C. M., Heijnen J. J. and Wilderer P. A.
(1997). Aerobic granular sludge in a sequencing batch reactor. Water Research, 31(12),
3191–3194.
Mosquera-Corral A., Montras A., Heijnen J. J. and van Loosdrecht M. C. M. (2003).
Degradation of polymers in a biofilm airlift suspension reactor. Water Research,
37(3), 485–492.
Mosquera-Corral A., de Kreuk M. K., Heijnen J. J. and van Loosdrecht M. C. M. (2005a).
Effects of oxygen concentration on N-removal in an aerobic granular sludge reactor.
Water Research, 39(12), 2676–2686.
Mosquera-Corral A., Vázquez-Padín J. R., Arrojo B., Campos J. L. and Méndez R.
(2005b). Nitrifying granular sludge in a sequencing batch reactor. In: Aerobic
Granular Sludge. Water and Environmental Management Series, S. Bathe, M. K.
de Kreuk, B. McSwain and N. Schwarzenbeck (eds), IWA Publishing, Munich,
pp. 63–70.
Mosquera-Corral A., Arrojo B., Figueroa M., Campos J. L. and Méndez R. (2011). Aerobic
granulation in a mechanical stirred SBR: treatment of low organic loads. Water
Science and Technology, 64(1), 155–161.
Moy B. Y. P., Tay J. H., Toh S. K. Liu Y. and Tay S. T. L. (2002). High organic loading
influences the physical characteristics of aerobic sludge granules. Letters in Applied
Microbiology, 34(6), 407–412.
Nereda® http://www.royalhaskoningdhv.com/nereda (accesed 25 March 2014)
Ni B.-J., Xie W.-M., Liu S.-G., Yu H.-Q., Wang Y.-Z., Wang G. and Dai X.-L. (2009).
Granulation of activated sludge in a pilot-scale sequencing batch reactor for the
treatment of low-strength municipal wastewater. Water Research, 43(3), 751–761.
Peng D., Bernet N., Delgenes J. P. and Moleta R. (1999). Aerobic granular sludge – a case
report. Water Research, 33(3), 890–893.
Pijuan M., Werner U. and Yuan Z. (2011). Reducing the startup time of aerobic granular
sludge reactors through seeding floccular sludge with crushed aerobic granules. Water
Research, 45(16), 5075–5083.
Qin L. Liu Y. and Tay J-H. (2004). Effect of settling time on aerobic granulation in
sequencing batch reactor. Biochemical Engineering Journal, 21(1), 47–52.
Schwarzenbeck N. and Wilderer P. A. (2005). Treatment of food industry effluents in
a granular sludge SBR. In: Aerobic Granular Sludge. Water and Environmental
Management Series, IWA Publishing, Munich, pp. 95–102.
Schwarzenbeck N., Erley R. and Wilderer P. A. (2004a). Aerobic granular sludge in an
SBR-system treating wastewater rich in particulate matter. Water Science and
Technology, 49(11–12), 41–46.
Schwarzenbeck N., Erley R., Mc Swain B. S., Wilderer P. A. and Irvine R. L. (2004b).
Treatment of malting wastewater in a granular sludge sequencing batch reactor (SBR).
Acta Hydrochimica et Hydrobiologica, 32(1), 16–24.
Seviour T., Lambert L., Pijuan M. and Yuan Z. (2011). Selectively inducing the synthesis
of a key structural exopolysaccharide in aerobic granules by enriching for candidatus
“competibacter phosphatis”. Applied Microbiology and Biotechnology, 92(6),
1297–1305.
Seviour T., Yuan Z., van Loosdrecht M. C. M. and Lin Y. (2012). Aerobic sludge granulation:
a tale of two polysaccharides? Water Research, 46(15), 4803–4813.
Shin H. S., Lim K. H. and Park H. S. (1992). Effect of shear stress on granulation in oxygen
aerobic upflow sludge reactors. Water Science and Technology, 26(3–4), 601–605.
Sun F. Y., Yang C. Y., Li J. Y. and Yang Y. J. (2006). Influence of different substrates on
the formation and characteristics of aerobic granules in sequencing batch reactors.
Journal of Environmental Sciences, 18(5), 864–871.
Tay J.-H., Liu Q.-S. and Liu Y. (2001a). The role of cellular polysaccharides in the formation
and stability of aerobic granules. Letters in Applied Microbiology, 33(3), 222–226.
Tay J.-H., Liu Q.-S. and Liu Y. (2001b). The effects of shear force on the formation, structure
and metabolism of aerobic granules. Applied Microbiology and Biotechnology,
57(1–2), 227–233.
Tay J-H, Liu Q-S. and Liu Y. (2001c). Microscopic observation of aerobic granulation in
sequential aerobic sludge blanket reactor. Journal of Applied Microbiology, 91(1),
168–175.
Tay J. H., Pan S., He Y. X. and Tay S. T. L. (2004). Effect of organic loading rate on
aerobic granulation. II: characteristics of aerobic granules. Journal of Environment
Engineering-ASCE, 130(10), 1102–1109.
Tay J. H., Liu Q.-S., Liu Y., Show K.-Y., Ivanov V. and Tay S. T.-L. (2005). A comparative
study of aerobic granulation in pilot- and labratory scale SBRs. In: Aerobic Granular
Sludge. Water and Environmental Management Series, IWA Publishing, Munich,
pp. 125–133.
Val del Río A., Figueroa M., Arrojo B., Mosquera-Corral A., Campos J. L., García-Torriello
G. and Méndez R. (2012a). Aerobic granular SBR systems applied to the treatment of
industrial effluents. Journal of Environmental Management, 95(Supplement), S88–S92.
Val del Río A., Morales N., Figueroa M., Mosquera-Corral A., Campos J. L. and Mendez R.
(2012b). Effect of coagulant-flocculant reagents on aerobic granular biomass. Journal
of Chemical Technology and Biotechnology, 87(7), 908–913.
Val del Río A., Morales N., Figueroa M., Mosquera-Corral A., Campos J. L. and Méndez
R. (2013). Effects of the cycle distribution on the performance of SBRs with aerobic
granular biomass. Environmental Technology, 34(11), 1463–1472.
Wang F., Lu S., Wei Y. and Ji M. (2009). Characteristics of aerobic granule and nitrogen and
phosphorus removal in a SBR. Journal of Hazardous materials, 164(2–3), 1223–1227.
Wilderer P. A., Irvine R. L. and Goronszy M. C. (2001). Sequencing Batch Reactor
Technology. IWA publishing, London, UK.
Winkler M. K., Kleerebezem R., Strous M., Chandran K. and van Loosdrecht M. C.
M. (2013). Factors influencing the density of aerobic granular sludge. Applied
Microbiology and Biotechnology, 97(16), 7459–7468.
Yang S. F., Tay J-H. and Liu Y. (2003). Effect of substrate N/COD ratio on the formation of
aerobic granules. Journal of Environment Engineering, 131(1), 86–92.
Yi S., Tay J. H., Maszenan A. M. and Tay S. T. L. (2003). A culture-independent approach
for studying microbial diversity in aerobic granules. Water Science and Technology,
47(1), 283–290.
Zhou D., Liu M., Gao L., Shao C. and Yu J. (2013). Calcium accumulation characterization
in the aerobic granules cultivated in a continuous-flow airlift bioreactor. Biotechnology
Letters, 35(6), 871–877.
Zima B. E., Diez L., Kowalczyk W. and Delgado A. (2007). Sequencing batch reactor (SBR)
as optimal method for production of granular activated sludge (GAS) – fluid dynamic
investigations. Water Science and Technology, 55(8–9), 151–158.
7.1 Introduction
Anaerobic digestion has been closely related to energy production since the major
final product of the process is a gaseous mixture of carbon dioxide and methane
(biogas). From this aspect, anaerobic digestion is by default an energy producing
technology and, depending on the substrate and the operating conditions, the
energy produced may exceed any energy required so that the net energy balance
may be positive.
Anaerobic digestion is practically an ‘energy transfer’ process. The energy
captured in the chemical bonds of the organic compounds of solid or liquid media
is transferred mainly to the C-H bonds of the gaseous methane where the carbon
atom is at its utmost reduced state, meaning that methane is the highest energy
density organic compound. This energy transfer is carried out by microorganisms
growing in the absence of oxygen by degrading a great range of organic substrates
or feedstocks. The microorganisms belong to groups of different physiology and
substrate affinity; they co-operate to break complex compounds into successively
smaller ones in a balanced anaerobic environment of natural ecosystems or robust
bioreactors.
The common application of the anaerobic digestion is the treatment of sewage
sludge and slurries in digesters. Landfills are also gigantic bioreactors where
biogas emissions indicate the anaerobic biological activity upon the organic
fraction of the municipal solid wastes. Other feedstocks considered to be ideal
substrates for anaerobic microorganisms are the high organic loaded liquid or
solid wastes of industries such as food, pulp, petrochemical and so on industries,
livestock establishments, and so on. Recently, dedicated crops or crop residues
7.2 The Process
Anaerobic digestion is a bioprocess consisting of a complex network of individual
steps catalysed by different groups of microorganisms. The microorganism
groups generally grow at different rates and become sensitive at a different degree
when exposed to certain environmental conditions (pH, ammonia, concentration
of metabolites such as volatile fatty acids, hydrogen etc.). The basic steps of the
process and the breakdown of the organic content of the initial feedstock (expressed
in Chemical Oxygen Demand, COD, units) are depicted in Figure 7.1. A brief
description of the steps follows:
• Disintegration. This step lumps processes such as lysis, non enzymatic decay,
phase separation and physical breakdown (e.g., shearing) (Batstone et al.
2002) and is responsible for the initial separation of the organic complex
(e.g., sludge particles) into carbohydrates, proteins and lipids as well as inert
material. It is regarded as the slowest step when the feedstock consists of
solid particles.
• Hydrolysis. After disintegration, hydrolytic enzymes are excreted by
microorganisms to breakdown the organic polymers (carbohydrates, proteins
and fats) into their respective monomers (sugars, amino acids, lipids), so that
they can be taken up by the microorganisms for further degradation.
Figure 7.1 COD flux for a particulate waste consisting of 10% inerts and 30% each
of the main organic polymers (in terms of COD) (Batstone et al. 2002).
7.3 The Technology
Due to the variety of factors influencing the anaerobic process efficiency, many
aspects of technology have been developed to improve and make the process
cost-effective: reactor engineering, operation practices – integration with other
processes, control and automation. All these aspects are based on the deep
comprehension of the biochemical processes of anaerobic digestion (Angelidaki
et al. 2012).
Reactor engineering in anaerobic digestion aims at (a) maintaining the
microorganisms inside the reactor and sustaining sufficient contact between them
and the substrate (b) increasing the reaction rates and eliminating the limiting
transport phenomena, and (c) providing the microorganisms with a suitable
environment to adapt and coexist under the operating conditions (Lettinga, 1995).
Based on these principles, high-rate configurations were developed based on the
contact process and the ability of the microorganisms to aggregate into granules
or form biofilms (filters, fluidised beds, sludge blanket reactors).
The innovation connected with the boost of anaerobic digestion technology
came with the development of the upflow anaerobic sludge blanket reactors
(UASBR) by Lettinga et al. (1980). They found that, under certain conditions, the
microorganisms tend to aggregate with a dense structure forming the granules.
Due to the high settling velocity, the granules are maintained in the interior of
the reactor in the lower part. UASB is a vertical column filled partially with
granules which are kept in suspension by introducing the wastewater from the
bottom at an appropriate upward velocity (to keep suspension but not break or
wash out the granules from the reactor). On the top of the reactor there is the
three phase separator device, the design of which is very important. When the
granules hit on the separator as they rise up, the biogas bubbles attached get
separated from the granule and as a result, they fall back on the bottom while the
effluent is removed from the top of the reactor. According to Tiwari et al. 2006)
there are more than 1000 full scale applications in the world based on UASB
technology. Since then, there are a lot of modifications developed combining the
features of UASBR, filters etc.).
The anaerobic membrane bioreactors (AnMBR) is another type of bioreactors
focusing on the complete retention of microorganisms by combinations of
microfiltration (MF) or ultrafiltration (UF) modules. High conversion rates are
achieved but the main disadvantages (membrane fouling and the concomitant
operating cost as well as the cost due to energy required for the pressure-driven
membranes to function) still remain and hinder the widespread application of this
technology (Lin et al. 2013; Stuckey, 2012).
Besides the basic reactor configuration, other important details of an
anaerobic digestion plant are the necessary number of reactors-units in series,
the organic loading of each reactor, the pH and the temperature, the addition
of trace metals and nutrients and so on (Stamatelatou et al. 2010; Takashima
& Speece, 1989; Zandvoort et al. 2006; Lv et al. 2010; Demirel & Yenigün,
2002). Other tools such as molecular techniques, developments on analytical
chemistry, modeling and simulation have contributed to a deeper insight of the
process itself (Vanwonterghem et al. 2014; Batstone et al. 2002; Pavlostathis &
Giraldo-Gomez, 1991), while developments on control and automation make the
anaerobic systems more robust (Pind et al. 2001).
the major component of the total sludge produced, since some sewage treatment
plants may not use primary treatment.
Figure 7.2 Sewage sludge route in a typical layout of a sewage treatment plant.
where the VSdes is the destruction of the volatile solids (%) and SRT the solid
retention time.
The production of biogas VQH4 (in m3/d) can be estimated according to:
Y ⋅ CODdes ⋅ Q
PX = (7.3)
1 + kd ⋅ SRT
where Y is the yield coefficient (0.05–0.1 kg VSS/kg COD destroyed and kd is the
biomass decay constant (0.02–0.04 d−1)
The biogas can be transformed either to thermal energy in boilers or both
thermal and electrical energy in combined heat units (CHP). Other alternatives are
upgrading it to biomethane to resemble the natural gas.
The main barrier for the anaerobic digestion of sewage sludge is the slow
hydrolysis step of the particulate matter; especially in the case of WAS, the
extracellular polymeric substances (EPS) contribute into forming a complex of a
slurry mass consisting of organics, inorganics and microorganisms. As a result,
the enhancement of the anaerobic digestion process has been often correlated
with the development of cost efficient pretreatment steps that would increase
the accessibility of the hydrolytic enzymes to the extracellular complex network
(formed by EPS) as well as the intracellular compounds.
Besides the energy recovery potential, anaerobic digestion of WAS results in the
improvement of sludge dewaterability, especially when combined with pretreatment
methods (Xu et al. 2011). EPS have been shown to influence the dewaterability
of sludge; although they favor the flocculation of sludge, their high water affinity
promote the hydration and decrease the dewaterability (Neyens & Baeyens, 2003).
The pretreatment methods considered to be suitable for enhancing the
anaerobic biodegradability of sewage sludge are ultrasonic treatment, mechanical
disintegration, chemical oxidation, treatment with alkali, thermal hydrolysis
and biological treatment (Stamatelatou et al. 2012). Combinations of the above
methods have also been studied (Dhar et al. 2012). However, with the introduction
of a new technology, new problems may arise; the energy consumption is usually
increased, the cost (operation and capital) is higher, environmental or other
economical problems result because of the harsh conditions applied. Consequently
the development of a pretreatment method is intriguing and the scientific interest
on this is growing. Some of these scientific developments have been proved in
numerous sewage treatment plants in the world (Table 7.1) according to a study of
Jolly and Gillard (2009). In the sequel, the principles of the most common studied
and evaluated economically methods are presented.
Table 7.1 Experience on sewage sludge technologies at full scale (Jolly & Gillard,
2009).
7.4.1 Sonication
During sonication, the ultrasound waves (sound waves at a frequency higher than
20 kHz) are emitted and transmitted into the slurry medium and create alternating
regions of positive and negative pressure (compression and rarefraction regions
respectively). When negative pressure is developed in a liquid medium, the gas
dissolved forms bubbles (cavitation bubbles) which collide due to the vibrations of
the ultrasound and become large up to a size that become unstable and collapse.
This phenomenon is known as cavitation. The bubble collapse is followed by sever
conditions of temperature and pressure (5000°C and pressure of 500 atmospheres)
which last a few microseconds (Pilli et al. 2011). These harsh conditions cause
high shear forces on the cells and disrupt them. Another effect of cavitation is
the generation of oxygen hydroxide radicals (•OH) which also contribute into the
disintegration of sludge through oxidation. However, Wang et al. (2005) concluded
that the main mechanism for disintegration by ultrasonication is the shear forces
induced by cavitation.
The ultrasonic pretreatment of sludge has been reviewed and evaluated with
respect to the effect of frequency, duration, the sonication density (power supplied
per volume of sludge treated), specific sonication energy (energy supplied per kg
of sludge solids) on the particle size reduction, the dewaterability and settlability
of sludge, the COD and nitrogen solubilization and the biogas production when
anaerobic digestion is applied on the sonicated sludge (Carrere et al. 2010; Pilli
et al. 2011). It would be expected that the positive effect of sonication would be
enhanced as the sonication density increases. However there seems to be a certain
limit beyond which, the intracellular polymers released will cause the flocculation
of the small particles through hydroxyl and negatively charged carboxyl groups.
Another negative effect of breaking the particles into smaller parts is that their
specific surface is increased and so is the water bounded per surface unit. In any
case, sonication favors COD solubilisation and, therefore, digestability. The increase
of biogas up to 50% has been observed after sludge pretreatment with sonication in
lab and full scale digesters (Carrere et al. 2010; Barber, 2005; Hogan et al. 2004).
7.4.2 Microwave
The microwave pretreatment is based on the electromagnetic radiation emitted
at a frequency of 0.3 to 200 GHz into a polar medium. The dipoles of the polar
medium (water molecules in the case of sludge) rotate to align with the alternating
electromagnetic field. The molecule rotation causes friction which produced
thermal energy. Due to reported results on the cell lysis and the change in the
structure of the proteins of the microorganisms, microwave has been considered to
be a potential method for sludge pretreatment. Compared to ultrasonic pretreatment,
it was found by Park et al. (2004) that the same level of COD solubilisation could
be achieved by providing a 3 time less energy per mass of sludge solids than in the
case of ultrasonic treatment.
7.4.3 Thermal hydrolysis
During thermal hydrolysis the sludge is subjected to elevated temperature levels
(150–200°C) under high pressure (600–2500 kPa). These severe conditions disrupt
the microbial cell membrane and the structure of other solids present in the sludge
mixture. As a result the soluble COD, which is more accessible to microorganisms,
is increased. The effectiveness of thermal hydrolysis depends on the combination
of temperature and pressure levels applied as well as the duration of pretreatment.
Generally, thermal pretreatment is considered suitable to be implemented before
mesophilic digestion and not thermophilic digestion because in the latter case,
the digestion process is much more efficient than the mesophilic digestion and the
thermal pretreatment does not result in substantial higher methane yields (Appels
et al. 2008). There are several full scale applications of thermal hydrolysis as a
pretreatment step of sewage sludge, developed by companies such as Cambi and
Kruger Inc (a subsidiary of Veolia Water).
7.4.5 Other methods
The application of pressure has been used to enhance the cell disruption of the
WAS. Specifically in the MicroSludge process, alkaline pre-treatment and milling
P ⋅ ∆t
SE = (7.4)
V ⋅ TSS
where P is the power (kW), Δt is the duration of sonication (s), V is the volume of
the sludge under sonication (L) and TSS is the concentration of the total suspended
solids of sludge (kg/L).
The energy imparted to the sludge volume, however, is lower, due to
losses from the electrical energy of the ultrasonic generator to the acoustical
energy transmitted in the medium. The series of energy transformations
during ultrasonic treatment is (Kobus & Kusinska, 2008): electrical →
mechanical → acoustical → cavitation → thermal. The imparted energy is
expressed as the acoustical energy and can be estimated from the thermal energy,
assuming that the acoustical energy will finally result in heat when cavitation and
bubble collapse occur. The thermal energy is the most used method for estimating
the acoustic energy in the medium and is based on the temperature change of the
medium with time (Kobus & Kusinska, 2008).
On the other hand, the thermal energy imparted in the medium during thermal
hydrolysis can be calculated directly based on the energy (Qs, kJ) needed to elevate
the temperature of the sludge (To, °C) to the temperature of the pretreatment (T, °C)
according to eq. (7.5):
where ρsl is the sludge density (kg/m3), Vsl is the volume of the sludge under thermal
hydrolysis and Cp is the specific thermal capacity of the sludge (4.18 kJ/kg °C). The
actual thermal energy consumption is calculated based on the heat losses during
heating. Moreover the recovery of thermal energy from the heated sludge should
be taken into account when estimating the cost of the process.
Dhar et al. (2012) studied the effect of sonication (from 1000 to 10000 kJ/kg),
thermal hydrolysis (from 50 to 90°C), and sonication at elevated temperature
(combined ultrasonic and thermal hydrolysis treatment) on sludge and found that the
increase of the ratio of soluble COD to total COD (Y, %) correlated to the imparted
energy to the sludge (X, kJ): Y = 0.247X + 7.056, R2 = 0.0801 regardless the
pretreatment method used. However, the increase in the biogas production did not
follow a linear correlation with the soluble to total COD ratio; Ultrasonic treatment
was more effective than thermal pretreatement under the conditions tested in terms
of biogas production, but the opposite was observed in terms of COD solubilisation.
Thermal pretreatment results in agglomeration and increase in particle size and
this could have influenced the methane yield (Bougrier et al. 2005). The increase
in temperature and the SE input did not seem to affect the methane yield either,
but the combination of both methods resulted in higher yields. Similar results were
obtained with volatile sulfur compounds generated. Sludge dewaterability was not
improved by temperature raise and was rather decreased at the highest SE input.
The assumptions for the cost estimation of the pretreatment technologies (ultrasonic
and thermal hydrolysis) were (a) the sludge temperature was 25°C, the heat recovery
from the thermally pretreated sludge was 80%. The cost for dewatering, transportation
and landfilling was $250/ton TSS, while for electricity and natural gas was $0.07/
kWh and $0,28/m3 respectively. The cost for biogas purification and specifically H2S
removal through non regenerable KOH-AC bed was $0.0005/m3 biogas (cost per unit
of biogas purification) and $12/kg H2S (cost absorbent per unit of H2S removal). The
results of this economic assessment showed that (Dhar et al. 2012):
(c) the thermal hydrolysis (at 50–90°C) combined with the untrasound
pretreatment (at 1000 kJ/kg TSS yielded a net saving from $44/ton TSS to
$66/ton TSS.
Other aspects which have an economic impact in the long term, but were not
included in the assessment of Dhar et al. (2012), are the prevention of erosion
of the equipment from sulfur compounds, the enhanced dewaterability of sludge
after pretreatment and digestion as well as the optimization of the polymer dose
and, finally, the investment cost for purchasing and installing the pretreatment
equipment. Jolly and Gillard (2009), on the other hand, retrieved data from full scale
applications of various technologies and estimated that pretreatment technologies
as well as thermophilic anaerobic digestion enhanced dewaterability allowing the
production of a sludge cake containing 25–32% Total Solids (TS). The cost of the
polymer dose (decreasing as the volatile solid content decreases) as well as the cost
of the anaerobic digestion liquor treatment (higher in ammonia concentration with
increasing the process efficiency) were taken into account.
An important factor for the economic evaluation is the energy balance and the
breakdown of the energy into thermal and electrical needed in each technology. The
energy required depends on the performance of each method which vary according
to the conditions (type of sludge – primary or secondary, solid content, temperature
conditions and duration of treatment, etc.). Table 7.2 summarises the estimation of
energy required in the case of some common pretreatment technologies and the
biogas yields obtained. The main assumptions are also included where available.
In some cases, the energy estimates are obscure to decipher, because it is
not clear if they concern the individual steps of the pretreatment or they refer to
the whole anaerobic digestion unit. For example the electrical energy required
for a mesophilic digester is 0.04 kWh kg−1VS or 0.032 kWh kg−1 TS according
to Carrere et al. (2010) and 0.150 kWh kg−1 TS according to Jolly and Gillard
(2009). These values are not comparable probably because Jolly and Gillard (2009)
have considered the electrical consumption of the whole plant (consisting of the
following stages: pre-digestion thickening, pre-treatment, anaerobic digestion,
CPH plant, post digestion storage, post digestion dewatering, and liquor treatment).
Comparison between the different technologies with respect to the mesophilic
digestion (as the control case) reveal that sonication requires much higher electrical
energy than the others and all three studies agree that the energy balance is negative.
In the case of thermal hydrolysis, the assumption of Pérez-Elvira (2011) that the
electrical demand is zero may be optimistic since electricity is indeed required to
drive the various units of the process. In any case the electrical energy required
is low compared to the thermal energy. All three studies concluded that the high
thermal energy demand of thermal hydrolysis has not a negative impact on the
economics since this form of energy can be recovered though the heat generated
by the CPH units, the hot streams of the process itself, and if more thermal energy
is needed, a part of biogas can be used in boilers (reducing the biogas available
by marinhoeng@hotmail.com
Technology Jolly and Gillard (2009) Pérez-Elvira (2011) Carrere et al. (2010)
Assumptions Electrical/ Assumptions Electrical/ Biogas Assumptions Electrical/ Biogas
Thermal energy Thermal yield (L Thermal yield
(kWh kg−1 TS) energy kg−1VS) Energy (kWh kg−1
(kWh m−3) (kWh kg−1 TS)
VS)
None (only Input: 6% TS 0.150/NR Input: 8% TS NR 488 Input: 6% TS, 0.04/0.5 1.9
mesophilic HRT: 18d 35°C, HRT: 17d 80% VS 35°C,
digestion TS reduction: 45% TS reduction: HRT: 20d VS
applied) 42% reduction: 40%
None (only Input: 6% 0.178/NR Input: 6% TS, 0.03/1.0 2.4
thermophilic) TS HRT: 22d 80% VS
VS reduction: 56% 55°C, HRT: 15d
VS reduction:
50%
to the CHP and thus ‘consuming’ the potential electrical energy that could be
produced. Pérez-Elvira (2011) showed that the process can be self sustained in the
case of a 13% TS, that is, the thermal energy for the pretreatment was obtained
by recovering the heat from the process itself, increasing the profitability of the
combined process. The author estimated that the economic benefit of treating
the sewage sludge from a population of 100,000 is €132,373/yr (8% TS inlet) and
€223,867/yr (13%TS). The biogas productivity reached 1.4 L L −1 d−1 compared to
0.26 observed in conventional AD systems (Pérez-Elvira et al. 2011).
The savings in energy is not the only criterion for selecting a treatment scheme.
Mills et al. (2014) studied five scenarios for conversion of sewage sludge to energy.
The core technology in all five scenarios was the anaerobic digestion process.
Thermal hydrolysis was selected as a pretreatment step in most of the scenarios.
Post treatment steps for biogas exploitation (in CHP or as biomethane) and digested
sludge disposal (land application, solid fuel production or conversion to pyrolysis
gas to be used in CHP) were considered as alternatives and evaluated against
the conventional scenario of an anaerobic digestion unit coupled with CHP and
the digested sludge be utilized in land applications. They performed a life cycle
analysis (LCA) to include both inflow and outflow materials and energy of each
scenario as well as the emissions to the environment on the assumption of treating
100 total dry solids per day. They also estimated the Capital Expenditure (CapEx)
based on the simplified equation (eq. (7.6)) as well as the annual operating expenses
(OpEx). Based on CApEx and OpEx and assumptions made for the discount rate
(8%), they calculated the internal rate of return (IRR) of the investment for each
scenario and considered both cases of the offer or absence of incentives by the
UK state.
is rather high in this case, and poses high risks for the investment if this policy
is adjusted or cancelled. The use of biomethane as vehicle fuel is a better option
since the higher prices of biomethane as vehicle fuel would make the investment
less independent on the incentives. The scenarios of using the digested sludge for
solid fuel or in CHP after pyrolysis are comparable with high IRR (14.39% and
17.46% with incentives and 8.48% and 7.64%) and the most positive environmental
impacts.
In the same line, Jolly and Gillard (2009) have concluded that the choice for
the final disposal of sludge determines the economics and the payback period.
Incineration of digested sludge, despite the high capital cost, results in more
positive cost balance and shorter payback period than land application. In their
study, they concluded that the thermal hydrolysis and thermophilic digestion are
the best choices in this respect. They admit that their conclusion on the efficiency
of thermophilic digestion should be verified by other works too, since they relied
their estimations on data taken from one single plant.
Co digestion of sludge with other feedstocks such as organic residues as glycerol
(Athanasoulia et al. 2014), landfill leachates (Pastor et al. 2013) agricultural wastes
or energy crops (Hidaka et al. 2013; Galitskaya et al. 2014) or food wastes (Serrano
et al. 2014; Belhadj et al. 2014; Powell et al. 2013; Dai et al. 2013) would increase
the efficiency of the process as well as the quality of the digested liquor and sludge.
The economic evaluation of developing a biogas unit within a sewage treatment
plant accepting more inflows than sewage to increase the methane potential could
be based on the work of Karellas et al. (2010). They developed an investment
decision tool for biogas production from biomass feedstocks. This tool requires
inputs such as the feedstock characteristics, availability and their gate fees (in
the case of outputs of industrial activities) or cost (in the case of biomass), the
market prices for the end products (electricity, heat, deigested sludge, liquor) and
additional revenues, the total capital and annual operating costs and any economic
incentives (loans, existing subsidies and grants). The output of this tool is the
economic evaluation of the investment in terms of essential economic indicators
such as the internal rate of return (IRR) and net present value (NPV) and so on.
Khalil et al. (2008) based on the large experience gained in India calculated
the net present value of sewage treatment plants adopting various technologies:
activated sludge process, trickling filters, stabilisation ponds, sequential batch
reactors, membrane bioreactors and UASB combined with a final polishing step
(ponds) or extended aeration systems. A 30 year life time of the investment with
a 12% interest was assumed and the year served as the basis for all calculations
was 2010. The net present values were in the national currency (rupee) and the
currency rate was 40 rupees per US dollar. The net present value for a plant of
a 50 MLD total capacity (in rupee) was 8100 (activated sludge process), 4800
(trickling filters), 1000 (stabilisation ponds), 7600 (sequential batch reactors),
13100 (membrane bioreactors), 2700 (UASB and aeration pond), 3300 (UASB and
extended aeration system). The stabilisation ponds is the cheapest technology but
requires large areas which may not be available in all cases. The next cost-effective
technologies are those with the UASB used as a pretreatment step, while the other
aerobic technologies were far more expensive.
The anaerobic membrane bioreactors (AnMBR) have also been studied. They
combine the features of anaerobic bioreactors and membranes which are able to
achieve separation of solids from the mixed liquor to a high degree. As a result,
the COD conversion (>85%) is similar to the one achieved in conventional aerobic
MBR without the cost of aeration, the total suspended solid removal is more than
99% (Lin et al. 2013). This is not the case for nutrient removal though; further
treatment is needed unless the treated effluent is used as a fertiliser. The two
common configurations with membranes located out of the bioreactor (external) or
inside the bioreactor (submerged) present advantages and disadvantages; the more
direct control on fouling and replacement of membranes, the higher fluxes but
more frequent cleaning, the negative effect on microbial activity due to the higher
fluxes and the high energy consumption (approximately 10 kWh/m3 product) in
the case of external An MBR, and the lower energy consumption, the more simple
and less frequent membrane cleaning and the milder operating conditions in the
case of submerged AnMBR. Lin et al. (2011) concluded that the decisive factors
for the life cycle cost of submerged AnMBR are the flux, the price and the lifetime
of membranes.
settled sludge from the A stage is very rich in organics and, in this respect, the A
stage could be applied as a preconcentration step (Verstraete et al. 2009). Other
preconcentration techniques include membrane filtration, dynamic filter filtration,
dissolved air flotation and on coagulation/flocculation by metal salt or polyelctrolyte
addition. The latter method is the chemically enhanced primary treatment (CEPT)
and has been studied with respect to the HRT, the types of suitable coaggulants, the
dose of coaggulants etc (Tchobanoglous et al. 2003; Libhaber & Jaramillo, 2012;
Harleman & Murcott, 1999).
Diamantis et al. (2013) made an economic evaluation of the CEPT process
followed by anaerobic digestion for a 2000 PE scenario, 15 years lifetime and a 6%
interest. They estimated the cost of the combined process as 0.2 €/m3 (0.1 €/m3 for
the CEPT and 0.1 €/m3 for the anaerobic digestion). The biogas produced suffices
for supplying the required energy to the digester and as a result the process can be
regarded as zero energy.
Moreover it seems that different strategies should be followed depending on
the scale of STP. For small scale STPs where anaerobic digestion of sludge is
not a feasible option, medium term storage of sludge is favored if biodegradation
is hindered. This is achieved by increasing the coagulant dose which ends up in
the concentrated stream inhibiting its biodegradation. Moreover, a high quality
supernatant is produced, simplifying the post treatment steps and reducing the
cost. On the other hand, for medium to large scale facilities, lower doses of
coagulant would not cause any problem in the digestability of the energy rich
concentrated stream.
Verstraete et al. (2009) introduced the term ‘used water’ for sewage regarding
sewage as a resource of energy and matter and not as something useless that can
be wasted. If energy, water and nutrients are recovered from the ‘used water’, then
the whole process for treating sewage can be economically viable with no waste
streams generated (Zero WasteWater). They estimated that the order of the total
cost for the combined preconcentration (dissolved air flotation or dynamic sand
filtration followed by ultrafiltration and reverse osmosis) and anaerobic digestion
steps vary between €0.66–0.95/m3. On the other hand, Verstraete and Vlaeminck
(2011) estimated that almost €1/m3 can be gained as profit from (a) recovering
water, heat, nitrogen and phoshorous and (b) producing energy from biogas and
biochar from digested sludge. This means that the zero wastewater approach can
be economically viable.
7.6 Conclusions
The evaluation of a technology developed or improved to yield high efficiency and
productivity should not be based solely on energy saving criteria or economic indices.
It is evident that a holistic approach is imposed in the case of sewage treatment. There
is a growing number of researchers that considered the sewage a resource and not
as waste. Based on this concept, there is an attempt to benefit from sewage as much
as possible with the least cost and minimum environmental impacts. As a result,
concepts such as the ‘zero wastewater’ (Verstraete & Vlaeminck, 2011), the energy
self-sufficient sewage treatment plants (Jenicek et al. 2013; Frijns et al. 2013), zero
carbon footprint (Novotny, 2011, 2012) and so on, indicate the desired targets for the
future sewage treatment plant: recover anything recoverable from a STP, no energy
consumption, no environmental impacts. However, a sewage treatment plant based
on such integrated concepts may be a good option for new installations, otherwise
retrofitting existing conventional facilities to novel, anaerobic based facilities seems
to be costly (McCarty et al. 2011).
The anaerobic digestion process has a key role in all these schemes since it
has been related with energy and matter recovery as well as economic profit.
Cost efficient technologies or practices that improve the efficiency and biogas
productivity of the anaerobic digestion process are in the core of such schemes. This
justifies the continuously growing effort on anaerobic digestion which, although
regarded as mature technology, still remains on the top of scientific interest. It
should be noted however, that the conclusion on the economic sustainability of the
sewage treatment plants of the future is based on assumptions for some units of the
concept. Whatever the risk of false estimation is, it is evident that a vision leading to
a less energy intensive and costly sewage management is under shape and becomes
inspiring. Following this vision, the modifications on existing conventional sewage
plants are carried out and the positive results are demonstrated in numerous case
studies (see second part of the present book as well as Dewettinck et al. 2001;
Zeeman et al. 2008; Jenicek et al. 2013 and many others).
7.7 References
Angelidaki I., Karakashev D., Batstone D. J., Plugge C. M., Alfons J. M. and Stams A. J.
M. (2011). Biomethanation and its potential. Methods in Enzymology, 494, 327–351.
Appels L., Baeyens J., Jan Degreve 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.
Athanasoulia E., Melidis P. and Aivasidis A. (2014). Co-digestion of sewage sludge and
crude glycerol from biodiesel production. Renewable Energy, 62, 73–78.
Barber W. P. (2005). The effects of ultrasound on sludge digestion. Chartered Institution of
Water and Environmental Management, 19, 2–7.
Batstone D. J. and Virdis B. (2014). The role of anaerobic digestion in the emerging energy
economy. Current Opinion in Biotechnology, 27, 142–149.
Batstone D. J., Keller J., Angelidaki I., Kalyuzhnyi S. V., Pavlostathis S. G., Rozzi A.,
Sanders W. T. M., Siegrist H. and Vavilin V. A. (2002). Anaerobic Digestion Model
No. 1 (ADM1), IWA, Task Group for Mathematical Modelling of Anaerobic Digestion
Processes. IWA Publishing, London.
Belhadj S., Joute Y., El Bari H., Serrano A., Gil A., Siles J. Á., Chica A. F. and Martín
M. Á. (2014). Evaluation of the anaerobic co-digestion of sewage sludge and tomato
waste at mesophilic temperature. Applied Biochemistry and Biotechnology, 172(8),
3862–3874.
Guo Js. and Xu Yf. (2011). Review of enzymatic sludge hydrolysis. Journal of Bioremediation
& Biodegradation, 2, 130.
Harleman D. R. F. and Murcott S. (1999). The role of physical-chemical wastewater
treatment in the mega-cities of the developing world. Water Science and Technology,
40(4–5), 75–80.
Harper S. R. and Pohland F. G. (1987). Enhancement of anaerobic treatment efficiency through
process modification. Journal of Water Pollution Control Federation, 59, 152–161.
Hidaka T., Arai S., Okamoto S. and Uchida T. (2013). Anaerobic co-digestion of sewage
sludge with shredded grass from public green spaces. Bioresource Technology, 130,
667–672.
Hogan F., Mormede S., Clark P. and Crane M. (2004). Ultrasonic sludge treatment for
enhanced anaerobic digestion. Water Science and Technology, 50(9), 25–32.
Jenicek P., Kutil J., Benes O., Todt V., Zabranska J. and Dohanyos M. (2013). Energy self-
sufficient sewage wastewater treatment plants: is optimized anaerobic sludge digestion
the key? Water Science and Technology, 68(8), 1739–1744.
Jolly M. and Gillard J. (2009). The Economics of Advanced Digestion. 14th European
Biosolids and Organic Resources Conference and Exhibition, 9–11 November, The
Royal Armouries, Leeds, UK.
Karellas S., Boukis I. and Kontopoulos G. (2010). Development of an investment decision
tool for biogas production from agricultural waste. Renewable and Sustainable Energy
Reviews, 14(4), 1273–1282.
Khalil N., Sinha R., Raghav A. K. and Mittal A. K. (2008). UASB Technology for Sewage
Treatment in India: Experience, Economic Evaluation and its Potential in other
Developing Countries. Twelfth International Water Technology Conference, IWTC12
2008, Alexandria, Egypt.
Khan A. A., Gaur R. Z., Tyagi V. K., Khursheed A., Lew B., Mehrotra I. and Kazmi A.
A. (2011). Sustainable options of post treatment of UASB effluent treating sewage: a
review. Resources Conservation and Recycling, 55, 1232–1251.
Kobus Z. and Kusinska E. (2008). Influence of physical properties of liquid on acoustic
power of ultrasonic processor. TEKA Komisji Motoryzacji i Energetyki Rolnictwa –
OL PA 8a, 71–78.
Lettinga G. (1995). Anaerobic digestion and wastewater treatment systems. Antonie van
Leeuwenhoek, 67, 3–28.
Lettinga G., Hobma S. W., Klapwijk A., Van Velsen A. F. M. and De Zeeuw W. J. (1980).
Use of the Upflow Sludge Blanket (UAS) reactor concept for biological wastewater
treatment. Biotechnology and Bioengineering, 22, 699–734.
Lew B., Tarre S., Belavski M. and Green M. (2004). UASB reactor for domestic wastewater
treatment at low temperatures: a comparison between a classical UASB and hybrid
UASB-filter reactor. Water Science and Technology, 49(11–12), 295–301.
Libhaber M. and Jaramillo A. O. (2012). Sustainable Treatment and Reuse of Municipal
Wastewater For Decision Makers and Practicing Engineers, IWA Publishing, UK.
Lin H., Peng W., Zhang M., Chen J., Hong H. and Zhang Y. (2013). A review on anaerobic
membrane bioreactors: applications, membrane fouling and future perspectives.
Desalination, 314, 169–188.
Lv W., Schanbacher F. L. and Yu Z. (2010). Putting microbes to work in sequence: recent
advances in temperature-phased anaerobic digestion processes (review). Bioresource
Technology, 101(24), 9409–9414.
Mahmoud N. (2008). High strength sewage treatment in a UASB reactor and an integrated
UASB-digester system. Bioresource Technology, 99(16), 7531–7538.
Mahmoud N., Zeeman G., Gijzen H. and Lettinga G. (2004). Anaerobic sewage treatment
in a one-stage UASB reactor and a combined UASB-digester system. Water Research,
38(9), 2348–2358.
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, 7100–7106.
Mills N., Pearce P., Farrow J., Thorpe R. B. and Kirkby N. F. (2014). Environmental
& economic life cycle assessment of current & future sewage sludge to energy
technologies, Waste Management, 34, 185–195.
Neyens E. and Baeyens J. (2003). A review of thermal sludge pre-treatment processes to
improve dewaterability. Journal of Hazardous Materials, B98, 51–67.
Novotny V. (2011). Water and energy link in the cities of the future – achieving net zero
carbon and pollution emissions footprint. Water Science and Technology, 63(1),
184–190.
Novotny V. (2012). Water and energy link in the cities of the future – achieving net zero
carbon and pollution emissions footprint. In: Water/Energy Interactions of Water
Reuse, V. Lazarova, K. H. Choo and P. Cornel (eds), IWA Publishing, London.
Park B., Ahn J.-H., Kim J. and Hwang S. (2004). Use of microwave pretreatment for enhanced
anaerobiosis of secondary sludge. Water Science and Technology, 50(9), 17–23.
Pastor L., Ruiz L., Pascual A. and Ruiz B. (2013). Co-digestion of used oils and urban
landfill leachates with sewage sludge and the effect on the biogas production. Applied
Energy, 107, 438–445.
Pavlostathis S. G. and Giraldo-Gomez E. (1991). Kinetics of anaerobic treatment: a critical
review. Critical Reviews in Environmental Control, 21(5–6), 411–490.
Pérez-Elvira S. I., Fdz-Polanco M. and Fdz-Polanco F. (2011). Enhancement of the
conventional anaerobic digestion of sludge: comparison of four different strategies.
Water Science and Technology, 64(2), 375–383.
Pilli S., Bhunia P., Yan S., LeBlanc R. J., Tyagi R. D. and Surampalli R. Y. (2011). Ultrasonic
pretreatment of sludge: review. Ultrasonics Sonochemistry, 18, 1–18.
Pind P. F., Angelidaki I., Ahring B. K., Stamatelatou K. and Lyberatos G. (2001).
Monitoring and control of anaerobic reactors. Advances in Biochemical Engineering/
Biotechnology, Springer, Berlin, 82, 135–182.
Powell N., Broughton A., Pratt C. and Shilton A. (2013). Effect of whey storage on biogas
produced by co-digestion of sewage sludge and whey. Environmental Technology
(United Kingdom), 34(19), 2743–2748.
Schellinkhout A. and Collazos C. J. (1992). Full-Scale Application of the UASB Technology
for Sewage Treatment. Water Science and Technology, 25(7), 159–166.
Seghezzo L., Zeeman G., van Lier J. B., Hamelers H. V. M. and Lettinga G. (1998). A
review: the anaerobic treatment of sewage in UASB and EGSB reactors. Bioresource
Technology, 65, 175–190.
Serrano A., Ángel Siles López J., Chica A. F., Ángeles Martin M., Karouach F., Mesfioui A.
and El Bari H. (2014). Mesophilic anaerobic co-digestion of sewage sludge and orange
peel waste. Environmental Technology (United Kingdom), 35(7), 898–906.
Stamatelatou K., Antonopoulou G. and Lyberatos G. (2010). Production of biogas via anaerobic
digestion. In: Handbook of Biofuels Production: Processes and Technologies, R. Luque,
J. Campelo and J. Clark (eds), Woodhead Publishing Series in Energy No. 15, UK.
Stamatelatou K., Antonopoulou G., Ntaikou I. and Lyberatos G. (2012). The effect
of physical, chemical and biological pretreatments of biomass on its anaerobic
digestibility and biogas production in Biogas Production: Pretreatment Methods in
Anaerobic Digestion, Scrivener Publishing, USA.
Stuckey D. C. (2012). Recent developments in anaerobic membrane reactors. Bioresource
Technology, 122, 137–148.
Subtil E. L., Cassini S. T. A. and Goncalves R. F. (2012). Sulfate and dissolved sulfide
variation under low COD/sulfate ratio in up-flow anaerobic sludge blanket (UASB)
treating domestic wastewater. Ambi-Agua Taubate, 7, 130–139.
Takashima M. and Speece R. E. (1989). Mineral requirements for methane fermentation
(review). Critical Reviews in Environmental Control, 19(5), 465–479.
Tchobanoglous G., Burton F. and Stensel H. D. (2003). Wastewater Engineering: Treatment
and Reuse. 4th edn, Metcalf and Eddy Inc., McGraw Hill, New York, USA.
Tiwari M., Guha S., Harendranath C. and Tripathi S. (2006). Influence of extrinsic factors
on granulation in UASB reactor. Applied Microbiology and Biotechnology, 71(2),
145–154.
Turovskiy I. S. and Mathai P. K. (2006). Wastewater Sludge Processing. Wiley-Interscience.
John Wiley & Sons, Inc., Hoboken, New Jersey.
Vanwonterghem I., Jensen P. D., Ho D. P., Batstone D. J. and Tyson G. W. (2014). Linking
microbial community structure, interactions and function in anaerobic digesters using
new molecular techniques, Current Opinion in Biotechnology, 27, 55–64.
Verstraete W. and Vlaeminck S. E. (2011). Zero WasteWater: short-cycling of wastewater
resources for sustainable cities of the future, International Journal of Sustainable
Development & World Ecology, 18(3), 253–264.
Verstraete W., van de Caveye P. and Diamantis V. (2009). Maximum use of resources
present in domestic ‘used water’. Bioresource Technology, 100, 5537–5545.
Vieira S. M. M. and Garcia A. D. Jr. (1992). Sewage treatment by UASB reactor. Operation
results and recommendations for design and utilization. Water Science and Technology,
25(7), 143–157.
Vieira S. M. M., Carvalho J. L., Barijan F. P. O. and Rech C. M. (1994). Application of the
UASB technology for sewage treatment in a small community at Sumare, Sao Paulo
state. Water Science and Technology, 30(12), 203–210.
Wang F., Wang Y. and Ji M. (2005). Mechanisms and kinetic models for ultrasonic waste
activated sludge disintegration. Journal of Hazardous Materials, 123, 145–150.
Xu H., He P., Yu G. and Shao L. (2011). Effect of ultrasonic pretreatment on anaerobic
digestion and its sludge dewaterability. Journal of Environmental Sciences, 23(9),
1472–1478.
Zandvoort M. H., van Hullebusch E. D., Fermoso F. G. and Lens P. N. L. (2006). Trace
metals in anaerobic granular sludge reactors: bioavailability and dosing strategies
(review). Engineering in Life Sciences, 6(3), 293–301.
Zhang L., Hendrickx T. L. G., Kampman C., Temmink H. and Zeeman G. (2013).
Co-digestion to support low temperature anaerobic pretreatment of municipal sewage
in a UASB-digester. Bioresource Technology, 148, 560–566.
Rep. of Ireland
2School of Biosystems Engineering, University College Dublin, Co. Dublin,
Rep. of Ireland
3Teagasc Environment Centre, Johnstown Castle, Co. Wexford, Rep. of Ireland
T.E. Laboratories Ltd., Loughmartin Industrial Estate, Tullow, Co. Carlow,
4
Rep. of Ireland
5 Danone Nutrition Ireland, Rocklands, Co. Wexford, Rep. of Ireland
6 Águas do Algarve S.A., Rua do Repouso, 10, 8000-302, Faro, Portugal
7AdP Energias, Rua Visconde de Seabra n°3, 1700-421 Lisboa, Portugal
8.1 Introduction
More than 10 million tons of sewage sludge was produced in the European Union
(EU) in 2010 (Eurostat, 2014). For the disposal of sewage sludge (solid, semisolid,
or liquid residue generated during the treatment of domestic sewage), chemical,
thermal or biological treatment, which may include composting, aerobic and
anaerobic digestion, solar drying, thermal drying (heating under pressure up to
260°C for 30 min), or lime stabilisation (addition of Ca(OH)2 or CaO such that pH
is ≥12 for at least 2 h), produces a stabilised organic material.
The Waste Framework Directive (2008/98/EC; EC, 2008) lays down measures
to protect the environment and human health by preventing or reducing adverse
impacts resulting from the generation and management of waste. Under the directive,
a hierarchy of waste is applied: prevention, preparing for re-use, recycling, other
recovery and disposal. The objective of the Directive is to maximise the resource
value and minimise the need for disposal (EC, 2008). This has prompted efforts
within sewage sludge management to utilise sewage sludge as a commodity. The
terminology ‘biosolids’ reflects the effort to consider these materials as potential
resources (Isaac & Boothroyd, 1996). Biosolids may be used in the production
of energy, bio-plastics, polymers, construction materials and other potentially
useful compounds. However, as the disposal of sewage sludge is commonly
achieved by recycling treated sludge to land, nutrient recovery, particularly in the
context of pressure on natural resources, and potential barriers to its reuse on land
(environmental, legislative), deserves particular attention.
The aim of this chapter is to examine the recovery of nutrients and other
compounds, such as volatile fatty acids (VFA), polymers and proteins, from sewage
sludge. Due to the increasing awareness regarding risks to the environment and
human health, the application of sewage sludge, following treatment, to land as a
fertilizer in agricultural systems has come under increased scrutiny. Therefore, any
potential benefits accruing from the reuse of sewage sludge are considered against
possible adverse impacts associated with its use. Finally, the potential costs and
benefits arising from its re-use are examined.
differences in sewage sludge treatments can be observed between the EU, USA and
Canada. With regards to sludge stabilization, aerobic and anaerobic treatments are
the most widely used methods of sewage sludge treatment. Within the EU, anaerobic
and aerobic wastewater treatments appear to be the most common methods, with 24
countries out of 27 applying this method (Kelessidis & Stasinakis, 2012).
Anaerobic digestion (AD) is most commonly used in Spain, Italy, United
Kingdom and Czech Republic (Table 8.1). Within the USA and Canada, biosolids
are classed according to their pathogenic levels. Class A biosolids contain
minute levels of pathogens and must undergo heating, composting, digestion, or
increased pH. Thus, these methods are more commonly employed (Table 8.1).
Class B biosolids have less stringent parameters for treatment and contain small,
but compliant, amounts of bacteria (USEPA, 2011). In order to achieve Class A
biosolids, the sewage sludge must undergo stringent treatment. Stabilization
methods such as aerobic, anaerobic, liming and composting, are the recommended
options in both the USA and Canada.
by marinhoeng@hotmail.com
Denmarka Francea Germanya Greecea,b Irelanda Italya Spaina Swedena UKa Czech Polanda Portugald USAc
Rep.a
Stabilisation
Aerobic
Anaerobic
Lime
Composting
Conditioning
Lime
Inorganics
Polymers
Thermal
Drying belts
Dewatering
Others
Thermal
Solar drying
Pasteurisation
Long-term storage
Cold fermentation
bag filling
Common use most common use
aKelessidis and Stasinakis (2012); bTsagarakis et al. (1999); c Lu et al. (2012); dMartins and Béraud, pers. comm.
Resource recovery from sewage sludge 143
Figure 8.1 Trends in unit cost of nitrogen (N), phosphorus (P) and potassium (K) in
chemical fertilisers in Ireland from 1980 to 2011 (Lalor et al. 2012).
One of the main concerns associated with the use of treated sewage sludge as an
organic fertiliser on grassland are the loss of nutrients, metals and pathogens along
a transfer continuum (Wall et al. 2011) to a waterbody via direct discharges, surface
and near surface pathways and/or groundwater discharge. More recently, so-called
‘emerging contaminants’, which may include antibiotics, pharmaceuticals and
other xenobiotics, have been considered, as they have health risks associated with
them. Therefore, nutrient recovery from treated sewage sludge must be considered
against possible adverse impacts associated with its use.
8.3.3 Polymers
Extracellular polymeric substances (EPS) are the major constituents of organic
matter in sewage sludge floc, which comprises polysaccharides, proteins, nucleic
acids, lipids and humic acids (Jiang et al. 2011). They occur in the intercellular space
of microbial aggregates, more specifically at or outside the cell surface (Neyens
et al. 2004), and can be extracted by physical (centrifugation, ultrasonication and
heating, for example) or chemical methods (using ethylenediamine tetraacetic
acid, for example), although formaldehyde plus NaOH has proven to be effective
in extracting EPA from most types of sludge (Liu & Fang, 2002). Extracellular
polymeric substances perform an important role in defining the physical properties
of microbial aggregates (Seviour et al. 2009). There are many biotechnical uses of
EPS, including the production of food, paints and oil drilling ‘muds’; their hydrating
properties are also used in cosmetics and pharmaceuticals. Furthermore, EPS
may have potential uses as biosurfactants for example, in tertiary oil production,
and as biological glue. Extracellular polymeric substances are an interesting
component of all biofilm systems and still hold large biotechnological potential
(Flemming & Wingender, 2001). A relatively new method for treatment of sewage
sludge is aerobic granular sludge technology (Morgenroth et al. 1997). A special
8.3.4 Proteins
Vermicomposting (sludge reduction by earthworms) is a relatively common
technology, especially in developing countries with small scale settings. The main
product of this process is vermicompost, which consists of earthworm faeces that
can be used as a fertilizer due to its high N content, high microbial activity and
lower heavy metal content (Ndegwa & Thompson, 2001). Vermicomposting results
in bioconversion of the waste streams into two useful products: the earthworm
biomass and the vermicompost. In a study by Elissen et al. (2010), aquatic worms
grown on treated municipal sewage sludge, produced high protein values with a
range of amino acids. These proteins can be used as animal feed for non-food
animals, such as aquarium fish or other ornamental aquatic fish. Other outlets for
the protein could be technical applications such as coatings, glues and emulsifiers.
The study also revealed that the dead worm biomass can be utilized as an energy
source in anaerobic digestion. Experiments have shown that biogas production of
worms is three times that of sewage sludge. Other applications include fats and
fatty acid extraction. Treatment of sewage sludge using earthworms has been well
documented; however, research studies on protein extraction of earthworms grown
on sewage sludge are very limited.
Bioconversion of biosolids using fly larvae has been studied for years. Organic
waste has a high nutritional and energy potential and can be used as a feed substrate
for larvae. Apart from significantly reducing organic waste, grown larvae make an
excellent protein source in animal feed. The insect protein could be used in animal
feed to replace fishmeal (Lalander et al. 2013). One of the most studied species is
the larvae of the Black Soldier fly (Hermetia illucens L.). The larvae of this non-
pest fly feed on, and thereby degrade, organic material of different origin (Diener
et al. 2011a). The 6th instar, the prepupa, migrates from the sludge to pupate
and can therefore easily be harvested. Since prepupae contain on average 44%
crude protein and 33% fat, it is an appropriate alternative to fishmeal in animal
feed (St-Hilaire et al. 2007). Proposals for other uses for the pupae other than
animal feed have been put forward. The other components of the pupae (protein,
fat, and chitin) could be fractioned and sold separately. The extracted fat can be
converted to biodiesel; chitin is of commercial interest due to its high percentage
8.5.3 Pathogens
During wastewater treatment, the sludge component of the waste becomes
separated from the water component. As the survival of many microorganisms
and viruses in wastewater is linked to the solid fraction of the waste, the numbers
of pathogens present in sludge may be much higher than the water component
(Straub et al. 1992). Although treatment of municipal sewage sludge using lime,
AD, or temperature, may substantially reduce pathogens, complete sterilisation is
difficult to achieve (Sidhu & Toze, 2009) and some pathogens, particularly enteric
viruses, may persist. Persistence may be related to factors such as temperature, pH,
water content (of treated sludge), and sunlight (Sidhu & Toze, 2009). Also, there
is often resurgence in pathogen numbers post-treatment, known as the ‘regrowth’
phenomenon. This may be linked to contamination within the centrifuge,
reactivation of viable, but non-culturable, organisms (Higgins et al. 2007), storage
conditions post-centrifugation (Zaleski et al. 2005), and proliferation of a resistant
sub-population due to newly available niche space associated with reduction in
biomass and activity (McKinley & Vestal, 1985).
The risk associated with sludge-derived pathogens is largely determined by their
ability to survive and maintain viability in the soil environment after landspreading.
Survival is determined by both soil and sludge characteristics. The major physico-
chemical factors that influence the survival of microorganisms in soil are currently
considered to be soil texture and structure, pH, moisture, temperature, UV radiation,
nutrient and oxygen availability, and land management regimes (reviewed in van
Elsas et al. (2011)), whereas survival in sludge is primarily related to temperature,
pH, water content (of treated sewage sludge), and sunlight (Sidhu & Toze, 2009).
Pertinent biotic interactions include antagonism from indigenous microorganisms,
competition for resources, predation and occupation of niche space (van Elsas et al.
2002). Pathogen-specific biotic factors that influence survival include physiological
status and initial inoculum concentration (van Veen et al. 1997).
Following landspreading, there are two main scenarios which can lead to human
infection. First, pathogens may be transported via overland or sub-surface flow to
surface and ground waters, and infection may arise via ingestion of contaminated
water or accidental ingestion of contaminated recreational water (Jaimeson et al.
2002; Tyrrel & Quinton, 2003). Alternatively, it is possible that viable pathogens
could be present on the crop surface following biosolid application, or may become
internalised within the crop tissue where they are protected from conventional
sanitization (Itoh et al. 1998; Solomon et al. 2002). In this case, a person may
become infected if they consume the contaminated produce. Therefore, it is
critical to accurately determine the pathogen risk associated with land application
of sewage sludge to fully understand the potential for environmental loss and
consequently, human transmission.
However, survival patterns of sludge-derived pathogens in the environment are
complex, and a lack of a standardised approach to pathogen measurement makes
it difficult to quantify their impact. For example, Avery et al. (2005) spiked treated
and untreated sludge samples with a known concentration of E. coli to quantify the
time taken to achieve a decimal reduction. The pathogen response was variable and
ranged from 3 to 22 days, depending on sludge properties. Lang and Smith (2007)
investigated indigenous E. coli survival in dewatered, mesosphilic anaerobically
digested (MAD) sludge, and in different soil types post MAD sludge application.
Again, decimal reduction times proved variable, ranging from 100 days when applied
to air-dried sandy loam, to 200 days in air-dried, silty clay. This time decreased to 20
days for both soil types when field moist soil was used, demonstrating the importance
of water content in regulating survival behaviour. Therefore, in order to quantify
pathogen risk in a relevant, site-specific manner, it is necessary to incorporate both
soil and treated sewage sludge characteristics in risk assessment modelling. This
has been done previously by conducting soil, sludge and animal slurry incubation
studies, where pathogens are often spiked to generate a survival response (Vinten
et al. 2004; Lang & Smith, 2007; Moynihan et al. 2013). Pathogen decay rate is
then calculated based on decimal reduction times, or a first-order exponential decay
model previously described by Vinten et al. (2004), and has been shown to be highly
contingent on soil type and sludge or slurry combinations. Currently, the Safe
Sludge Matrix provides a legal framework for grazing animals and harvesting crops
following landspreading of treated sewage sludge, and stipulates that a time interval
of three weeks and 10 months should be enforced to ensure safe practice, respectively
(ADAS, 2001). However, further work is required to determine if these regulations
are overly stringent, particularly in light of the comparatively higher pathogen
concentrations reported for animal manures and slurries. For example, E. coli
concentrations ranged from 3 × 102 to 6 × 104 CFU g−1 in sludge (Payment et al.
2001), compared to 2.6 × 108 to 7.5 × 104 CFU g−1 in fresh and stored cattle slurry,
respectively (Hutchison et al. 2004). Therefore, environmental losses associated with
treated sewage sludge application may not be as extensive as previously thought, and
further comparisons on pathogen risk should form the basis of future research.
8.5.4 Pharmaceuticals
Pharmaceuticals comprise a diverse collection of thousands of chemical
substances, including prescription and over-the-counter therapeutic drugs and
veterinary drugs (USEPA, 2012). Pharmaceuticals are specifically designed to
alter both biochemical and physiological functions of biological systems in humans
and animals (Walters et al. 2010). Pharmaceuticals are referred to as ‘pseudo-
persistent’ contaminants (i.e., high transformation/removal rates are compensated
by their continuous introduction into the environment) (Barceló & Petrovic, 2007).
Pharmaceuticals are likely to be found in any body of water influenced by raw or
treated waste water, including river, lakes, streams and groundwater, many of which
are used as a drinking water source (Yang et al. 2011). Between 30 and 90% of an
administered dose of many pharmaceuticals ingested by humans is excreted in the
urine as the active substance (Cooper et al. 2008). In a survey conducted by the US
Environmental Agency (see McClellan & Halden, 2010), the mean concentration
of 72 pharmaceuticals and personal care products were determined in 110 treated
sewage sludge samples. Composite samples of archived treated sewage sludge,
collected at 94 U.S. wastewater treatment plants from 32 states and the District
by marinhoeng@hotmail.com
(adapted from RPA, Milieu Ltd. And WRc., 2008; Fytili & Zabaniotou, 2008; Astals et al. 2012; Cao & Pawłowski, 2012).
Depending on the type of treatment applied, costs associated with the re-use
of sewage sludge may include, amongst other issues, drying, lime amendment,
thermal drying costs, along with costs of installation of storage facilities in which
to carry out these treatments; labour, energy and transport costs; and where the
treated sewage sludge is re-used on land, soil and sewage sludge analysis costs and
other professional service costs (Table 8.3). Potential benefits accruing from the
land application of treated sewage sludge may be enhanced nutrient availability to
crops and enhanced crop yield, and in countries where sewage sludge, treated or
untreated, is considered a waste material (e.g., Ireland), there is a substantial saving
for the farmer.
By application of this protocol, energy costs can be reduced. Through the re-use of
energy produced during wastewater treatment, the long-term sustainability of the
WWTPs is enhanced, while also contributing to offset installation and on-going
operational costs.
In Southern European countries, including the Mediterranean area, cultural,
social and economic reasons means that the management of the sewage sludge is
not necessarily the same as in other EU countries. Here, recycling to agriculture
is the main route for final disposal. For example, in Portugal and Spain about
50% of the sewage sludge is recycled in agriculture (Milieu et al. 2013a, 2013b,
2013c). Therefore, sewage sludge management in these countries should be
governed by the following objectives (Martins & Béraud, pers. comm.): (1) pro
vision of solutions that are technically and economically adapted to the economic
realities of these countries (lower investment and operating costs); (2) full legal
compliance, including the ability to adapt to future restrictions, which may be
placed on the disposal of treated sludge in agriculture; (3) diversification of the
final disposal of sludge with new sludge treatment systems; (4) reduction in the
quantity of sewage sludge to be disposed of; (5) optimization of the utilisation of
weather conditions for sludge treatment, which makes solar drying an appealing
solution.
8.7 Acknowledgements
The authors wish to acknowledge funding from the Irish EPA (Project reference
number 2012-EH-MS-13). M.G. Healy and A. Martins are members of EU COST
Action Water_2020.
8.8 References
ADAS (2001). Safe Sludge Matrix – Guidelines for the Application of Sewage Sludge to
Agricultural Land, UK.
Alam M. Z., Molla A. H. and Fakhru’l-Razi A. (2005). Biotransformation of domestic
wastewater treatment plant sludge by two-stage integrated processes – LSB & SSB.
IIUM Engineering Journal, 6, 23–40.
Andersen A. and SEDE (2002). Disposal and Recycling Routes for Sewage Sludge.
Synthesis Report (22 February 2002). http://ec.europa.eu/environment/waste/sludge/
pdf/synthesisreport020222.pdf (accessed 17 September 2013).
Apedaile E. (2001). A perspective on biosolids management. Canadian Journal of
Infectious Diseases, 12(4), 202–204.
Astals S., Venegas C., Peces M., Jofre J., Lucena F. and Mata-Alvarez J. (2012). Balancing
hygienization and anaerobic digestion of raw sewage sludge. Water Research, 46(19),
6218–6227.
Avery L. M., Killham K. and Jones D. L. (2005). Survival of E. coli O157: H7 in organic
wastes destined for land application. Journal of Applied Microbiology, 98(4), 814–822.
Barceló D. and Petrovic M. (2007). Pharmaceuticals and personal care products (PPCPs)
in the environment. Analytical and Bioanalytical Chemistry, 387(4), 1141–1142.
EC (2010). Environmental, Economic and Social Impacts of the Use of Sewage Sludge on
Land. Summary Report 1. http://ec.europa.eu/environment/waste/sludge/pdf/part_iii_
report.pdf (accessed 11 June 2014).
EC (2012). Sewage Sludge. http://ec.europa.eu/environment/waste/sludge/index.htm
(accessed 11 June 2014).
EEC (1986). Council Directive of 12 June 1986 on the Protection of the Environment, and in
Particular of the Soil, when Sewage Sludge is used in Agriculture (86/278/EEC). http://
www.efma.org/PRODUCT-STEWARDSHIP-PROGRAM-10/images/86278EEC.pdf
(accessed 11 June 2014).
Elissen H. J. H., Mulder W. J., Hendrickx T. L. G., Elbersen H. W., Beelen B., Temmink H.
and Buisman C. J. N. (2010). Aquatic worms grown on biosolids: biomass composition
and potential applications. Bioresource Technology, 101(2), 804–811.
Epstein E. (2003). Land Application of Sewage Sludge and Biosolids. Lewis, Boca Raton, Fl.
European Commission (2009). Directive 2009/28/EC of the European Parliament and of the
Council of 23 April 2009 on the promotion of the use of energy from renewable sources
and amending and subsequently repealing Directives 2001/77/EC and 2003/30/EC.
European Commission (2011). EUROSTAT. http://epp.eurostat.ec.europa.eu/tgm/table.do?t
ab=table&init=1&language=en&pcode=ten00030&plugin=0 (accessed 19 November
2013).
Eurostat (2014). Sewage Sludge Production and Disposal. http://appsso.eurostat.ec.europa.
eu/nui/show.do?dataset=env_ww_spd&lang=en (accessed 11 June 2014).
Evanylo G. K., Haering K. C., Sherony C. and Christian A. H. (2011). Mid-Atlantic
composting directory. Virginia Cooperative Extension Publication, Virginia Tech,
Blacksburg, VA, 452–230. http://pubs.ext.vt.edu/452/452-230/452-230.pdf (accessed
21 November 2014).
Fehily Timoney and Company (1999). Codes of good practice for the use of biosolids
in agriculture – guidelines for farmers. http://www.environ.ie/en/Publications/
Environment/Water/FileDownLoad,17228,en.pdf (accessed 11 June 2014).
Fenton O., Schulte R. P. O., Jordan P., Lalor S. T. J. and Richards K. G. (2011). Time
lag: a methodology for the estimation of vertical and horizontal travel and flushing
timescales to nitrate threshold concentrations in Irish aquifers. Environmental Science
and Policy, 14, 419–431.
Flemming H. C. and Wingender J. (2001). Relevance of microbial extracellular polymeric
substances (EPSs) – Part II: Technical aspcets, 43, 9–16.
Food Safety Authority of Ireland (2008). Food safety implications of land-spreading
agricultural, municipal and industrial organic materials on agricultural land used for
food production in Ireland. www.fsai.ie/WorkArea/DownloadAsset.aspx?id=8226
(accessed 11 June 2014).
Fytili D. and Zabaniotou A. (2008). Utilization of sewage sludge in EU application of old
and new methods – a review. Renewable and Sustainable Energy Reviews, 12(1),
116–140.
Galbally P., Ryan D., Fagan C. C., Finnan J., Grant J. and McDonnell K. (2013). Biosolid
and distillery effluent amendments to Irish short rotation coppiced willow plantations:
impacts on groundwater quality and soil. Agricultural Water Management, 116, 193–203.
Gerba C. P. and Smith J. E. (2005). Sources of pathogenic microorganisms and their
fate during land application of wastes. Journal of Environmental Quality, 34,
42–48.
Girovich M. J. (1996). Biosolids Treatment and Management: Processes for Beneficial Use.
CRC Press, New York.
Higgins M. J., Chen Y.-C., Murthy S. N., Hendrickson D., Farrel J. and Schafer P. (2007).
Reactivation and growth of non-culturable indicator bacteria in anaerobically digested
biosolids after centrifuge dewatering. Water Research, 41, 665–673.
Hogendoorn A. (2013). ‘Enhanced digestion and alginate-like-exopolysaccharides extraction
from Nereda sludge’. Masters in Science Civil Engineering, University of Technology,
Delft. http://www.citg.tudelft.nl/fileadmin/Faculteit/CiTG/Gezondheidstechniek/doc/
Afstudeerrapporten/final_thesis_anthonie_hogendoorn.pdf (accessed 21 November
2014).
Hutchison M. L., Walters L. D., Avery S. M., Synge B. A. and Moore A. (2004). Levels
of zoonotic agents in british livestock manures. Letters in Applied Microbiology, 39,
207–214.
Hytiris N., Kapellakis I. E., La Roij de R. and Tsagarakis K. P. (2004). The potential use
of olive mill sludge in solidification process. Resources, Conservation and Recycling,
40(2), 129–139.
Ibrahim T. G., Fenton O., Richards K. G., Fealy R. M. and Healy M. G. (2013). Loads
and forms of nitrogen and phosphorus in overland flow and subsurface drainage on a
marginal land site in south east Ireland, Biology and Environment: Proceedings of the
Royal Irish Academy, 113B(2), 169–186.
Isaac R. A. and Boothroyd Y. (1996). Beneficial use of biosolids: Progress in controlling
metals. Water Science and Technology, 34(3–4), 493–497.
Itoh Y., Sugita-Konishi Y., Kasuga F., Iwaki M., Hara-Kudo Y., Saito N., Noguchi Y.,
Konuma H. and Kumagai S. (1998). Enterohemorrhagic Escherichia coli O157: H7
present in radish sprouts. Applied and Environmental Microbiology, 64, 1532–1535.
Jamieson R. C., Gordon R. J., Sharples K. E., Stratton G. W. and Madani A. (2002).
Movement and persistence of fecal bacteria in agricultural soils and sub-surface
drainage water: a review. Canadian Biosystems Engineering, 44, 1–1.9.
Jeng A. S., Haraldsen T. K., Grønlund A. and Pedersen P. A. (2006). Meat and bone meal
as nitrogen and phosphorus fertilizer to cereals and rye grass. Nutrient Cycle in
Agroecosystems, 76, 183–191.
Jiang J., Zhao Q., Wei L., Wang K. and Lee D.-J. (2011). Degradation and characteristic
changes of organic matter in sewage sludge using microbial fuel cell with ultrasound
pretreatment. Bioresource Technology, 102(1), 272–277.
Kapshe M., Kuriakose P. N., Srivastava G. and Surjan A. (2013). Analysing the co-benefits:
case of municipal sewage management at Surat, India. Journal of Cleaner Production,
58, 51–60.
Kelessidis A. and Stasinakis A. S. (2012). Comparative study of the methods used
for treatment and final disposal of sewage sludge in European countries. Waste
Management, 32(6), 1186–1195.
Knacker T. and Metcalfe C. (2010). Introduction to the special issue on environmental
risk assessment of pharmaceuticals. Integrated Environmental Assessment and
Management, 6(S1), 511–513.
Lalander C., Diener S., Magri M. E., Zurbrügg C., Lindström A. and Vinnerås B.
(2013). Faecal sludge management with the larvae of the black soldier fly (hermetia
illucens) — from a hygiene aspect. Science of the Total Environment, 458–460,
312–318.
dynamics and soil biochemical and microbiological properties. Soil Biology and
Biochemistry, 40, 462–474.
Morgenroth E., Sherden T., van Loosdrecht M. C. M., Heiinen J. J. and Wilderer P. A.
(1997). Aerobic granular sludge in sequencing batch reactors. Water Research, 31(12),
3191–3194.
More T. T., Yan S., Tyagi R. D. and Surampalli R. Y. (2010). Potential use of filamentous
fungi for wastewater sludge treatment. Bioresource Technology, 101(20), 7691–7700.
Mouri G., Takizawa S., Fukushi K. and Oki T. (2013). Estimation of the effects of chemically-
enhanced treatment of urban sewage system based on life-cycle management.
Sustainable Cities and Society, 9(0), 23–31.
Moynihan E. L., Richards K. G., Ritz K., Tyrrel S. and Brennan F. P. (2013). Impact of
soil type, biology and temperature on the survival of non-toxigenic Escherichia coli
O157. Biology and Environment, 113(B), 1–6.
Navas A., Machn J., and Navas B. (1999). Use of biosolids to restore the natural vegetation
cover on degraded soils in the badlands of Zaragoza (NE Spain). Bioresource
Technology, 69, 199–205.
Ndegwa P. M. and Thompson S. A. (2001). Integrating composting and vermicomposting
in the treatment and bioconversion of biosolids. Bioresource Technology, 76(2),
107–112.
Neyens E., Baeyens J., Dewil R. and De heyder B. (2004). Advanced sludge treatment
affects extracellular polymeric substances to improve activated sludge dewatering.
Journal of Hazardous Materials, 106(2–3), 83–92.
Payment P., Plante R. and Cejka P. (2001). Removal of indicator bacteria, human enteric
viruses, giardia cysts, and cryptosporidium oocysts at a large wastewater primary
treatment facility. Canadian Journal of Microbiology, 47,188–93.
Pérez-Cid B., Lavilla I. and Bendicho C. (1999). Application of microwave extraction for
partitioning of heavy metals in sewage sludge. Analytica Chimica Acta, 378(1–3),
201–210.
Robinson K. G., Robinson C. H., Raup L. A. and Markum T. R. (2012). Public attidudes and
risk perception toward land application of biosolids within the south-eastern United
States. Journal of Environmental Management, 98, 29–36.
RPA, Milieu Ltd. And WRc. (2008). Environmental, Economic and Social Impacts of the
use of Sewage Sludge on Land. Final report. Part II: report on options and impacts.
http://ec.europa.eu/environment/waste/sludge/pdf/part_ii_report.pdf (accessed 6
November 2013).
Seviour T., Pijuan M., Nicholson T., Keller J. and Yuan Z. (2009). Gel-forming
exopolysaccharides explain basic differences between structures of aerobic sludge
granules and floccular sludges. Water Research, 43(18), 4469–4478.
Siddique M. T. and Robinson J. S. (2004). Differences in phosphorus retention and release
in soils amended with animal manures and sewage sludge. Soil Science Society of Am
Journal, 68, 1421–1428.
Sidhu J. P. S., and Toze S. G. 2009. Human pathogens and their indicators in biosolids: a
literature review. Environment International, 35, 187–201.
Singh R. P. and Agrawal M. (2008). Potential benefits and risks of land application of
sewage sludge. Wate Management, 28, 347–358.
Smith S. R. and Durham E. (2002). Nitrogen release and fertiliser value of thermally-dried
biosolids. Water and Environment Journal, 16(2), 121–126.
van Leeuwen J. H., Rasmussen M. L., Sankaran S., Koza C. R., Erickson D. T., Mitra D.
and Jin B. (2012). Fungal treatment of crop processing wastewaters with value-added
co-products. In: Sustainable Bioenergy and Bioproducts, K. Gopalakrishnan, J. (Hans)
van Leeuwen and Robert C. Brown (eds), Springer, Germany, pp. 13–44.
van Veen J. A., van Overbeek L. S., and van Elsas J. D. (1997). Fate and activity of
microorganisms introduced into soil. Microbiology and Molecular Biology Reviews,
61(2),121–135.
Vinten A. J. A., Douglas J. T., Lewis D. R., Aitken M. N. and Fenlon D. R. (2004). Relative
risk of surface water pollution by E. coli derived from faeces of grazing animals
compared to slurry application. Soil Use and Management, 20(1), 13–22.
Wall D., Jordan P., Melland A. R., Mellander P.-E., Buckley C., Reaney S. M. and Shortle G.
(2011). Using the nutrient transfer continuum concept to evaluate the European Union
Nitrates Directive National Action Programme. Environmental Science and Policy,
14(6), 664–674.
Walters E., McClellan K. and Halden R. U. (2010). Occurrence and loss over three years
of 72 pharmaceuticals and personal care products from biosolids–soil mixtures in
outdoor mesocosms. Water Research, 44(20), 6011–6020.
Wett B., Buchauer K. and Fimml C. (2007). Energy self-sufficiency as a feasible concept
for wastewater treatment systems. In: IWA Leading Edge Technology Conference,
Singapore.
Yan S., Subramanian S., Tyagi R. and Surampalli R. (2008). Bioplastics from waste activated
sludge-batch process. Practice Periodical of Hazardous, Toxic, and Radioactive
Waste Management, 12(4), 239–248.
Yang X., Flowers R. C., Weinberg H. S. and Singer P. C. (2011). Occurrence and removal
of pharmaceuticals and personal care products (PPCPs) in an advanced wastewater
reclamation plant. Water Research, 45(16), 5218–5228.
Yang X., Du M., Lee D.-J., Wan C., Zheng L. and Wan F. (2012). Improved volatile fatty
acids production from proteins of sewage sludge with anthraquinone-2,6-disulfonate
(AQDS) under anaerobic condition. Bioresource Technology, 103(1), 494–497.
Zaleski K. J., Josephson K. L, Gerba C. P. and Pepper I. L. (2005). Survival, growth and
regrowth of enteric indicator and pathogenic bacteria in biosolids, compost, soil and
land applied biosolids. Journal of Residuals Science and Technology, 2(1), 49–63.
9.1 Introduction
Over the past few decades, atmospheric pollution has become more important
since recent investigations have consistently demonstrated it poses a threat for
human health and natural ecosystems. In October 2013, the specialized cancer
agency of the World Health Organization (WHO) classified outdoor air pollution
as a human carcinogenic and related it to lung and bladder cancer (WHO, 2013).
The harmful consequences derived from polluted gas emissions have resulted in
an increasing public concern and the enforcement of the stricter environmental
legislations (Stuetz et al. 2001; Inrapour et al. 2005).
Odorous emissions constitute an important contributor to atmospheric pollution
and represent a significant contribution to photochemical smog formation and
particulate secondary contaminant emission (Sucker et al. 2008). Moreover, at
wastewater treatment plants (WWTPs), the accumulation of specific odorants
such as H2S in confined spaces may reach lethal concentrations, entailing a severe
occupational risk to the operators (Vincent, 2001). H2S also leads to the corrosion
of valuable assets, reducing the life of the WWTP’s infrastructure.
Malodors are the main cause of the public complaints received by environmental
regulatory agencies worldwide (Kaye & Jiang, 2000). Malodorous emissions from
WWTPs rank among the most unpleasant ones and their nuisance on the nearby
emission, they might increase the odour concentration of the emission even above
regulatory limits (Decottignies et al. 2007; Bilsen & De Fre, 2008).
Finally, odour control technologies are implemented when neither prevention
nor mitigation are viable or sufficiently efficient. It is important to highlight that
the capital and operating costs associated with odour treatment using traditional
technologies (biofilters, biotrickling filters, adsorption and chemical scrubbing)
might represent from 5% up to 15% of the total costs of WWTPs (Kiesewetter et al.
2012). Odour abatement often involves covering and extracting foul air from the
odour emission source and subsequently treating this air using a specific process
unit prior to atmospheric discharge. These treatment technologies are based either
on physical-chemical or biological principles (Lebrero et al. 2011).
In this chapter, an energy/economic efficiency analysis of the most typically
employed odour treatment technologies (chemical scrubbing, activated carbon
filtration, biofiltration and biotrickling filtration) together with three hybrid systems
(chemical scrubbing with activated carbon filtration, biotrickling filtration with
activated carbon filtration and biotrickling filtration with chemical scrubbing) and
three emerging technologies recently applied for odour abatement (activated sludge
diffusion, activated sludge recycling and step-feed biofiltration) will be presented.
The sensitivity of this energy/economic efficiency towards design parameters such
as the length-to-diameter ratio (L/D) in packed based systems will be also evaluated.
Figure 9.1
Schematic illustration of a chemical scrubber (a) and an activated
carbon adsorption unit with steam regeneration (b).
activated sludge (Figure 9.2c). Odour REs > 99% have been recorded in large-
scale WWTPs treating the odorous emissions in their aeration tanks (Kiessewetter
et al. 2012; Lebrero et al. 2011; Barbosa et al. 2006). ASD has been applied for
more than 30 years mainly in North America, and only in the past decade started
to be perceived as a real engineered alternative for odour abatement all over the
world. Recent works have ruled out the traditional concerns of detrimental effects
on wastewater treatment caused by the sparging of malodours in the aeration
basin due to pH modification by the H2S present in the malodorous stream or to
possible alterations in the structure of the biological communities responsible for
wastewater treatment (Barbosa & Stuetz, 2013). A more detailed description of
these odour control technologies can be found elsewhere (Delhomenie & Heitz,
2005; Lebrero et al. 2011; Estrada et al. 2013a).
Figure 9.2 Schematic illustration of a biofilter (a), a biotrickling filter (b) an activated
sludge diffusion system (c) and activated sludge recycling (d).
Chemical 2m Intalox Saddles 4s 10 years 1370 EUR 137 EUR m−3 20700 EUR 1000 Pa Liquid recirculation:
scrubbing (2 s per m−3 per media 180 L m−3 min−1
(CS) stage) substitution
two-stage
NaOH-NaClO
Activated 0.6 m Granular 2.5 s 6 months 5.5 EUR 137 EUR m−3 20700 EUR 2250 Pa No regeneration of
carbon impregnated AC kg−1 year−1 the activated carbon
filtration (450 kg m−3) was considered in the
(AC) estimation of lifespan and
costs
Biotrickling 4 m (2 m Inert polyurethane 15 s 10 years 1370 EUR 137 EUR m−3 20700 EUR 1000 Pa Liquid recirculation:
Filtration per stage) foam (PUF) m−3 per media 7.2 L m−3 min−1.
(BTF) substitution Liquid renewal rate: 2.5 L
per g of H2S removed
(Continued)
169
170
by marinhoeng@hotmail.com
Table 9.1 Data compilation for the design parameters and operating costs for the different technologies evaluated (Continued).
Technology Height Packing material EBRT Packing Packing Packing Labour, Pressure Others
material Material Material transport Drop
lifespan Purchase Disposal and
Costs Costs handling
costs
1 stage 2 m + 0.6 m PUF + Standard 9s+ PUF: 10 1370 EUR 137 EUR m−3 22770 EUR 2500 Pa Extended lifespan of
BTF + AC AC 2.5 s years. m−3 for PUF. year−1 2 years for AC due to the
AC: 2 4.1 EUR lower concentration of
years kg−1 for AC odorants to be treated in
the adsorption unit.
1 stage 2 m + 0.6 m Intalox Saddles + 2s+ Intalox 1370 EUR 137 EUR m−3 22770 EUR 2500 Pa Extended lifespan of 2
CS + AC Standard AC 2.5 s Saddles: m−3 for year−1 years for AC due to the
10 years. Intalox lower concentration of
Table 9.2 Data compilation for the operating costs based on previous studies by
Estrada et al. 2012 (updated to 2012).
Chemicals Caustic soda EUR 0.175 kg−1 50% w/w, density 1.53 kg l−1
Hypochlorite EUR 0.210 kg−1 12.5% w/w, density 1.22 kg l−1
Activated Virgin EUR 4.11 kg−1 density 0.45 kg l−1
Carbon Impregnated EUR 5.50 kg−1 density 0.45 kg l−1
Water Potable EUR 1.12 m−3
secondary effluent EUR 0.56 m−3
Activated carbon filtration (AC). The adsorbent selected for the filtration process
was granular impregnated activated carbon, characterized by a density of 450 kg
m−3, purchase cost of 5.5 EUR kg−1 and a lifespan of 6 months (no regeneration of
the activated carbon was considered). The adsorption filter consisted of a 0.6 m
height column operated at an EBRT of 2.5 s, resulting in a total pressure drop
of 2250 Pa including grease filters and ductwork. The disposal costs of activated
carbon were 137 EUR m−3, while AC transport and renewal costs added up to
20700 EUR year−1.
Biofiltration (BF). A 1 m biofilter packed with compost with a lifespan of
2 years was selected in this study. The system operated at an EBRT of 60 s and
a total pressure drop of 1500 Pa (including the pressure drop of the humidifier
and ductwork). Irrigation of the biofilter was performed by means of 2 water nets
located at the top of the unit, each of them provided with 49 drips m−2 and each drip
irrigating 1.9 L h−1 for 3 min day−1. Total humidification requirements were estimated
to be 0.02 kg water (kg air)−1. The packing purchase costs were 82 EUR m−3, while
the costs associated to its disposal and transport-handling were 48 EUR m−3 and
33130 EUR year−1, respectively.
Biotrickling filtration (BTF). A two-stage BTF operated at acid (~2) and neutral
pH, respectively, with a total height of about 4 m (2 m per stage) was considered
as model BTF. Inert PUF, with a lifespan of 10 years and a cost of 1370 EUR m−3,
was selected as the packing material. Disposal and labour costs were estimated
at 137 EUR m−3 and 20700 EUR per media substitution, respectively. The total
pressure drop in the system and ductwork was 1000 Pa. The EBRT, liquid recycling
and liquid renewal rate were set at 15 s, 7.2 L m−3 min−1 and 2.5 L (gH2S removed)−1,
respectively.
BTF + AC. This hybrid technology consisted of a single stage BTF and an AC
filter acting as a polishing step. Similar operating parameters as in stand-alone
technologies were used in the hybrid technology except for a shorter EBRT of 9 s
in the BTF and the use of standard activated carbon with a price of 4.1 EUR kg−1
and an extended lifespan of the packing material of up to 2 years due to the lower
concentration of odorants to be treated in the unit after the treatment in the BTF.
The total pressure drop of this combined system was 2500 Pa.
CS + AC. A single-stage CS operated at an EBRT of 2 s was coupled with an
AC as a polishing step. The rest of the operating parameters were maintained as
in the stand-alone technologies except for the use of standard activated carbon
packing with a price of 4.1 EUR kg−1 and extended lifespan of 2 years due to the
lower concentration of odorants to be treated in the unit. A total pressure drop of
2500 Pa was considered.
BTF + CS. A single-stage CS acts as the polishing stage after a single-stage
BTF in this hybrid technology. EBRTs in the BTF and the CS stages can be set
at 9 s and 2 s, respectively, to fulfil the target odour and H2S REs. The rest of
the operating parameters were maintained as in the stand-alone technologies. The
total pressure drop of this combined system was 1500 Pa.
Step-feed biofilter (Step BF). The operation of the standard BF was modified
by supplying the odourous emission in three different locations along the BF
height (Estrada et al. 2013b). This configuration allows for an increased packing
lifespan of 25% compared to the conventional BF, while reducing the overall
pressure drop of the bed by 25% (total pressure drop in the step-feed BF of
1250 Pa).
Activated sludge diffusion (ASD). The odorous emission is sparged into an
activated sludge tank (devoted to wastewater treatment) with a depth of 4 m. An
additional pressure drop of 500 Pa was considered to take into account the ducting
required to conduct the emission to the aeration basin. Grease filters, corrosion
resistant blower, upgrade to fine bubble diffusers and instrumentation were
included as capital costs.
Activated sludge recycling (ASR). A sludge flowrate of 625 m3 h−1 is pumped
from the secondary settler of the activated sludge tank to the head of the WWTP,
representing 5% of the total wastewater flowrate treated in the plant (a model
WWTP treating 300 megaliter per day was considered, typical size of a WWTP
with approximately 50000 m3 h−1 malodorous air emission). Piping, sludge pumps,
dispensers, valves, instrumentation and automation needed were included in the
costs of this technology.
The CO2 footprint for each technology was calculated according to the data
shown in Table 9.3. The following transportation distances were assigned to
the different materials required in the technologies evaluated: 50 km of road
transportation to the compost needed in BF based on the possibility of locally
purchasing this packing material; 500 km of road transportation to the polyurethane
foam (PUF) and Intalox Saddles required in the BTF and CS, respectively, based
on the possibility of purchasing these materials inside the country; 5000 km of sea
transportation + 200 km of road transportation to the activated carbon, according
the present trend of activated carbon purchase from Asian manufacturers. Finally,
15 km of road transportation were considered for the disposal of all spent packing
materials in local landfills.
technology with the highest energy requirements among the biological techniques,
mainly due to the significant pressure drop in the system. Step feed biofiltration,
where the total gas flow is split and fed at different heights along the biofilter
bed, was shown to reduce the overall pressure drop across the bed without
significantly impacting BF performance, which resulted in similar pressure drops
to those recorded in BTFs (Estrada et al. 2013b). In ASR, the energy requirements
were exclusively devoted to the pumping of the return activated sludge from the
secondary settler at a flow rate of ≈5% of the total wastewater volume treated in
the plant. This rate was estimated based on recent experimental results and should
be able to reduce the hydrogen sulphide concentration in the inlet wastewater by
80–95% (Zhang et al. 2011).
Finally, ASD presented the lowest energy requirements among the technologies
evaluated, since it only accounts for the energy needed to overcome the pressure
drop of the piping to conduct the air to the aeration basin (500 Pa). The energy
consumption associated to the bubbling of the air emission into the reactor was
not considered as an extra energy need for the technology, since these energy
requirements are usually already taken into account in the overall consumption in
the water treatment line of the WWTP (Estrada et al. 2011).
abatement that a technology can offer, are usually most relevant when selecting
technologies for full-scale applications. Therefore, the application of ‘cost/benefit
parameters’ accounting for different aspects of the odour abatement process and
lumping information into a single parameter would allow for a fairer comparison
among odour abatement technologies.
In this context, the cost/benefit Energy Efficiency Parameter (EEP) was devised
to account for three key features of the evaluated techniques: one cost and two
benefits. The Net Present Value in 20 years (NPV20, EUR) was considered as the
‘technology cost’. The NPV20 accounts for the total amount of money spent in the
installation and operation of a technology for a period of 20 years and includes
both investment and operating costs as defined elsewhere (Estrada et al. 2011).
Any change reducing the operating or investment costs, such as more economic
construction or packing materials, lower salaries in work costs or an increased
packing material lifespan would reduce the NPV20 values as described in depth
in Estrada et al. (2012). The first ‘benefit’ considered was the odour abatement
performance quantified as odour RE for the corresponding technology. The
second relevant ‘benefit’ selected was the inverse of the power required (P, kW),
considering a low energy consumption as a potential benefit in a technology. Thus,
the EEP was defined as follows:
NPV20
EEP = (9.1)
RE ⋅ ( P )−1
A technology with a low EEP value will be therefore preferred since ideally this
would entail a low NPV20, high odour REs and low energy consumption.
The main conclusions derived from the EEP analysis were in agreement
with those drawn from the total energy consumption data (Figure 9.3), with
physical/chemical and the hybrid technologies also exhibiting the highest
EEP values (Figure 9.4). Compared to the overall energy consumption, the
differences among technologies got sharpened using this parameter, CS + AC
being the less efficient technology in terms of costs and energy use. This finding
confirmed previous research on the field that concluded that physical/chemical
technologies are both more expensive and energy demanding than their biological
counterparts (Estrada et al. 2011). In the particular case of CS + AC, the high
operating costs of both individual technologies combined with their intrinsic
high energy requirements, ranked this technology as the less preferred with an
EEP value of 3.37 × 108 EUR kW. AC and CS ranked as the second and third
less efficient technologies (EEP values of 2.35 × 108 and 2.08 × 108 EUR kW,
respectively) based on the previously mentioned rationales. In this context, the
hybrid technologies involving BTF performed better than the above discussed
technologies despite exhibiting similar power requirements. Both BTF + AC
and BTF + CS exhibited EEP values below 1.5 × 108 (Figure 9.4) even though
they involved similar energy requirements to CS, AS and CS + AC (Figure 9.3).
This enhanced EEP derived from the lower NPV20 of these technologies, since
the upfront biological technology significantly counter-balanced the overall
costs of the AC or the CS employed as polishing steps (mainly due to the lower
frequency of activated carbon purchase and replacement, and lower chemical
usage, respectively).
Figure 9.4 Energy Efficiency Parameter for the evaluated odour treatment
technologies.
Interestingly, BF was the less efficient technology among the purely biological
techniques due to its high operating costs caused by the low lifespan of the
organic packing material (EEP value of 6.75 × 107 EUR kW). BTF and step feed
BF exhibited similar EEPs of 4.26 × 107 and 4.48 × 107 EUR kW, respectively,
as a result of their low operating costs and low energy requirements. ASR and
ASD emerged as the most efficient technologies in terms of economic and energy
efficiency. Despite their lower odour REs (75% for ASD and 50% for ASR vs 95%
considered for the rest of technologies), the low NPV20 inherent to ASD and ASR
ranked these techniques as the top performing technologies for odour abatement
in terms of EEP. However, it is important to remark that they might not be able to
achieve the odour removal required in some particular full scale applications, and
in these scenarios complementary solutions to odour control in WWTP should be
considered. In this context, ASR has been scarcely tested to date and ASD might
not be able to treat the whole malodorous emission when its flow rate exceeds
the aeration requirements of the wastewater aerobic treatment (Kiesewetter et al.
2012), which would entail the implementation of an additional treatment. In brief,
these wastewater treatment-associated technologies should be considered in future
WWTP designs and emerge as promising techniques, with limitations that must
be gradually overcome in on-site research programs.
NPV20
SEP = (9.2)
RE ⋅ (CO2 ftp)−1
A technology with a low SEP value will be therefore preferred since ideally
this would entail a low NPV20, high odour REs and low CO2 footprint. Any
parameter reducing the NPV20 would therefore be beneficial for the SEP. In this
context, the use of renewable energy sources would decrease the CO2 footprint
also improving the SEP results, specially benefiting those technologies with higher
energy demands.
The SEP (Figure 9.5) was strongly correlated to the EEP previously reported
(Figure 9.4) since the main contributor to the CO2 footprint was the CO2 associated
to energy consumption, and most technologies were designed to achieve similar
odour and H2S REs. The CO2 footprint associated to energy usage was at least one
order of magnitude higher than the CO2 footprint from other sources, except for
BF and Step BF, where it was only 3 and 5 times higher, respectively (Table 9.4).
Packing material manufacturing, transport and the untreated odour emission
contributed marginally to the overall CO2 footprint among the technologies
evaluated. On the other hand, chemicals manufacture became the second most
important contributor in CS and CS + AC. In technologies with low packing
material lifespan such as AC, BF and Step BF, packing material manufacture
constituted the second contributor to the total CO2 footprint. Finally, the untreated
odorous emission became the second most important contributor to the total CO2
footprint in technologies with low packing material and transport requirements
and/or low RE, such as BTF, ASD and ASR (Table 9.4).
The CS + AC hybrid technology was the least efficient technology due to
its high energy requirements and chemical usage, contributing to a higher CO2
footprint (SEP value of 7.34 EUR t CO2-eq m −3 treated). CS and AC ranked second
and third in terms of sustainability efficiency with SEP values of 4.76 and 4.63
EUR t CO2-eq m −3treated, respectively, confirming the relatively poor performance
of physical/chemical techniques in terms of environmental impact. Despite the
absence of CO2 footprint associated to chemical usage, AC presented the highest
CO2 footprint associated to packing material manufacture due to its low lifespan.
Figure 9.5 Sustainability Efficiency Parameter for the odour treatment technologies
evaluated.
short lifespan of their organic packing material, which increased packing material
transportation frequency and compost utilization, with the subsequent increase in the
CO2 emissions associated to those aspects. In our particular analysis, BF exhibited a
SEP value of 1.66 EUR t CO2-eq m −3 treated, very close to that estimated for the hybrid
BTF + CS system (SEP = 1.99 EUR t CO2-eq m −3 treated). BTF constituted the preferred
option in terms of sustainability among the conventional end-of-the-pipe odour
treatment technologies, with a SEP value of 0.79 EUR t CO2-eq m −3 treated. However, the
wastewater treatment-associated technologies ASD and ASR were indeed the most
cost-sustainable with SEP values of 0.53 and 0.09 EUR t CO2-eq m −3 treated, respectively.
This confirmed ASD and ASR are extremely cost-efficient, while entailing a low
environmental impact in terms of energy use and CO2 emissions.
Despite providing a limited knowledge of the overall environmental impact
of odour abatement technologies, the use of CO2 footprint as environmental
benefit parameter confirmed the widely accepted best performance of biological
techniques for odour abatement (Estrada et al. 2011). In addition, the SEP could be
formulated to include other environmental or social impacts relevant in the future.
Similarly, weighted factors could be added to tune the relevance of a certain cost
or benefit in the SEP definition. For instance, in a sensitive scenario where the
odour RE constitutes the most important parameter, the RE could be multiplied
by a sensitivity factor.
Technology
BTF + AC
BTF + CS
CS + AC
Step BF
ASD
ASR
BTF
AC
CS
BF
This estimated robustness (R) can be then included as a benefit in the previous cost/
benefit parameter SEP in order to have a more complete comparative parameter.
NPV20
SREP = (9.3)
RE ⋅ (CO2 ftp)−1 ⋅ R −1
Chemical scrubbing was the least efficient technology according to the SREP
as a result of its relatively low robustness due to the key importance of water and
chemicals for its correct operation (Figure 9.6). The high robustness of adsorption
systems derived from its relative simplicity and the fact that it does not rely on
water or process control to operate may explain the worldwide acceptance of AC
and all the AC-involving technologies among the WWTP operators and resulted in
SREPs comparable to biofiltration. The consideration of process robustness among
technology selection criteria entailed that the hybrid technology CS + AC would be
preferred over CS (SREP values of 118 vs 185 EUR t CO2-eq m −3 treated, respectively).
The hybrid technology BTF + AC, combining the low energy consumption and
environmental impact of BTF and the robustness of AC, remained at the level of the
step BF in terms of SREP (41 vs 35 respectively EUR t CO2-eq m −3 treated). BTF would
be the preferred conventional technology according to the SREP analysis due to
its balanced costs and sustainability despite not showing the highest robustness.
Figure 9.6 Sustainability and Robustness Efficiency Parameter for the evaluated
odour treatment technologies.
from their relative simplicity and ease of operation. Despite practical experience
available is scarce, especially for ASR, the high robustness of ASD for odour
abatement was recently confirmed (Lebrero et al. 2010). In addition, both pilot and
field studies suggest that the diffusion of malodorous emissions into the activated
sludge process does not negatively affect the efficiency of wastewater treatment
(Barbosa & Stuetz, 2013), but ASD might promote the development of filamentous
bacteria and induce a poor biomass sedimentation under high H2S concentration
scenarios (Kiesewetter et al. 2012).
On the other hand, the combination of increased packing material and chemical
consumption at 45 ppm of H2S boosted the SREP value of the CS + AC. Finally,
the good performance of BTF + AC and BTF + CS even at high H2S concentration
must be highlighted, which confirmed that physical/chemical technologies highly
reduced their overall costs and environmental impacts when implemented in
combination with a biological technology.
special trucks; (iii) the high gas velocities achieved in units with high L/D ratios
mediate an enhanced odorant mass transfer from the gas to the aqueous phase
or to the adsorbent surface, which is a key operational issue since mass transfer
limitations are commonly encountered in odour treatment applications (Kim &
Deshusses, 2008; Iranpour et al. 2005). Energy efficiency considerations apply
also in the selection of the optimum L/D ratio in biofilters, where a packed bed
height of approximately 1 m is commonly accepted as a maximum in order to limit
the pressure drop across the bed, and the diameter is then determined by the EBRT
(Iranpour et al. 2005).
However, in the current scenario of increasing energy prices around the
world, a modification in conventional design parameters such as the L/D ratio
could be considered. In this regard, technologies exhibiting the highest pressure
drops (i.e., AC, BF, CS + AC and BTF + AC) presented the highest benefits in
terms of reductions in power consumption derived from a decrease in the L/D ratio
(Figure 9.8). For instance, a reduction of 30% in the L/D ratio resulted in energy
savings of up to 22% in AC, 20% in both BF and Step-feed BF, 18% in CS + AC
and 21% in BTF + AC, mainly due to a decreased pressure drop in each reactor
configuration. In the particular case of CS, the benefits derived mainly from the
savings in energy consumption for liquid recycling. However, the energy savings
in a CS accounted only for 11% when the L/D ratio is reduced by a 30% due
to the low pressure drop across the packing material and the lower size of the
reactors employed. BTFs exhibited moderate savings of 15% when the L/D ratio
was reduced by 30%.
Figure 9.8 Influence of the L/D ratio on the power consumption of the evaluated
technologies, maintaining the rest of the parameters constant.
These results suggest that relatively small variations in the L/D ratio can strongly
impact on the energy consumption of the treatment technologies. Moreover,
unnecessary increases in L/D ratios, even by a moderate 30%, can derive in a
9.4 Conclusions
In brief, physical/chemical technologies exhibited an inferior performance than
their biological counterparts in terms of energy consumption, economic and
environmental efficiency.
Among biological technologies, biotrickling filtration (BTF) exhibited one
of the lowest energy requirements and better overall efficiency when process
economic, environmental impacts and robustness are taken into account.
Biotrickling filtration technology can be backed up by an activated carbon
filtration unit (AC) in order to increase its robustness, providing better sustainability
than AC in standalone applications. When backed-up by chemical scrubbing (CS),
the hybrid BTF + CS technology showed similar sustainability results to those of
BTF + AC, but a lower process robustness.
Standard biofiltration (BF) showed low energy requirements and environmental
impacts, however, its sustainability and robustness efficiency decreased to values
similar to those of AC when robustness was considered. However, biofiltration
performance can be increased in both economic and environmental terms when a
step-feed configuration is employed.
The lack of sufficient reliable data on activated sludge diffusion (ASD) and
activated sludge recycling (ASR) performance limited a complete comparison of
both technologies with the rest of well-established systems. However, preliminary
results showed their potential as low cost environmentally-friendly technologies
for odour control in WWTPs.
Physical/chemical technologies are the most sensitive technologies to variations
in H2S concentration. Overall, an increase in H2S concentration highly impacted
the chemical and/or packing material needs of physical/chemical technologies,
which at the same time entailed higher environmental impacts and operating costs.
On the other hand, biological techniques were less influenced by H2S concentration
fluctuations, which rendered biotechniques more predictable over time in economic
and environmental terms.
Finally, variations in the design of conventional technologies must be evaluated
to reduce power consumption and increase efficiency based on the increasing
energy prices. Limited reductions of 30% in the L/D ratio can yield reductions
in the energy requirements of ≈20% for most of the odour treatment technologies
here evaluated. However, there is still a lack of experimental data on the influence
of L/D ratio on RE of odour treatment technologies.
9.5 References
Agentschap (2012). Translation of I&E Monitoring Protocol 12-014.
Alfonsín C., Hernández J., Omil F., Prado Ó. J., Gabriel D., Feijoo G. and Moreira M. T.
(2013). Environmental assessment of different biofilters for the treatment of gaseous
streams. Journal of Environmental Management, 129(0), 463–470.
Barbosa V. L. and Stuetz R. M. (2013). Performance of activated sludge diffusion
for biological treatment of hydrogen sulphide gas emissions. Water Science and
Technology, 68(9), 1932–1939.
Barbosa V. L., Burgess J. E., Darke K. and Stuetz R. M. (2002). Activated sludge
biotreatment of sulphurous waste emissions. Reviews in Environmental Science and
Biotechnology, 1, 345–362
Barbosa V. L., Hobbs P. J., Sneath R. W., Callen J. and Stuetz R. M. (2006). Investigating the
capacity of an activated sludge process to reduce odour emissions. Water Environment
Research, 78, 842–851.
Bilsen I. and De Fré R. (2008). Evaluation of a Neutralizing Agent Applied on a Cooling
Tower at a Citric Acid Production Plant. 3rd IWA International Conference on Odour
and VOCs, Barcelona, Spain.
Boldrin A., Andersen J. K., Møller J., Christensen T. H. and Favoino E. (2009). Composting
and compost utilization: accounting of greenhouse gases and global warming
contributions. Waste Management & Research, 27(8), 800–812.
CEPCI Chemical Engineering Plant Cost Index (2012). Chemical Engineering, New York.
Capodaglio A. G., Conti F., Fortina L., Pelosi G. and Urbini G. (2002). Assessing the
environmental impact of WWTP expansion: Odour nuisance and its minimization.
Water Science and Technology, 46(9), 339–346.
Card T. (2001). Chemical odour scrubbing systems. In: odours in Wastewater Treatment:
Measurement, Modelling and Control, R. Stuetz and F.-B. Frenchen (eds), IWA
Publishing, Padstow, Cornwall, UK.
Decottignies V., Filippi G. and Bruchet A. (2007). Characterisation of odour masking
agents often used in the solid waste industry for odour abatement. Water Science and
Technology, 55(5), 359–364.
Delhoménie M. C. and Heitz M. (2005). Biofiltration of air: a review. Critical Reviews in
Biotechnology, 25(1), 53–72.
Estrada J. M., Kraakman N. J. R., Muñoz R. and Lebrero R. (2011). A comparative analysis
of odour treatment technologies in wastewater treatment plants. Environmental
Science and Technology, 45(3), 1100–1106.
Estrada J. M., Kraakman N. J. R., Lebrero R. and Muñoz R. (2012). A sensitivity analysis
of process design parameters, commodity prices and robustness on the economics of
odour abatement technologies. Biotechnology Advances, 30(6), 1354–1363.
Estrada J. M., Lebrero R., Quijano G., Kraakman N. J. R. and Muñoz R. (2013a). Strategies
for odour control. In: Odour Impact Assessment Handbook, V. Belgiorno, V. Naddeo
and T. Zarra (eds), John Wiley & Sons, Ltd., Hoboken, NJ, USA.
Estrada J. M., Quijano G., Lebrero R. and Muñoz R. (2013b). Step-feed biofiltration: a low
cost alternative configuration for off-gas treatment. Water Research, 47(13), 4312–4321.
Husband J. A., Phillips J., Coughenour J. R., Walz T. and Blatchford G. (2010). Innovative
approach to centrate nitrification accomplishes multiple goals: nitrogen removal and
odour control. Water Science and Technology, 61(5), 1097–1103.
IChem (2002). The Sustainability Matrics. The Institution of Chemical Engineers, Rugby,
UK.
IMO (2009). Second IMO GHG Study. International Maritime Organization, London, UK.
Iranpour R., Cox H. H. J., Deshusses M. A. and Schroeder E. D. (2005). Literature Review
of air pollution control biofilters and biotrickling filters for odor and volatile organic
compound removal. Environmental Progress, 24(3), 254–267.
Jehlickova B., Longhurst P. J. and Drew G. H. (2008). Assessing effects of odour: a
critical review of assessing annoyance and impact on amenity. 3rd IWA International
Conference on Odour and VOCs, Barcelona, Spain.
Kaye R. and Jiang K. (2000). Development of odour impact criteria for sewage treatment
plants using odour complaint history. Water Science and Technology, 41(6), 57–74.
Kiesewetter J., Kraakman N. R. J., Cesca J., Trainor S. and Witherspoon J. (2012).
Expanding the use of Activated Sludge at Biological Wastewater Treatment Plants
for Odor Control. Proceedings of the WEF Odors and Air Pollutants Conference.
Louiseville, Kentucky, USA.
Kim S. and Deshusses M. A. (2008). Determination of mass transfer coefficients for
packing materials used in biofilters and biotrickling filters for air pollution control. 1.
Experimental results. Chemical Engineering Science, 63(4), 841–855.
Kraakman N. J. R. (2003). Robustness of a full-scale biological system treating industrial
CS2 emissions. Environmental Progress, 22(2), 79–85.
Kraakman N. J. R., Roche Rios J. and van Loosdrecht M. C. M. (2011). Review of mass
transfer aspects for biological gas treatment. Applied Microbiology and Biotechnology,
91, 873–886.
Lebrero R., Rodríguez E., Martin M., García-Encina P. A. and Muñoz R. (2010). H2S and
VOCs abatement robustness in biofilters and air diffusion bioreactors: a comparative
study. Water Research, 44(13), 3905–3914.
Lebrero R., Bouchy L., Stuetz R. and Muñoz R. (2011). Odor assessment and management
in wastewater treatment plants: a review. Critical Reviews in Environmental Science
and Technology, 41(10), 915–950.
Mattinen M. and Nissinen A. (2011). Carbon Footprint Calculators for Public Procurement.
The Finnish Environment 36/2011, Helsinki, Finland.
Owen W. F. (1982). Energy in Wastewater Treatment. Prentice Hall, USA.
Prado Ó. J., Gabriel D. and Lafuente J. (2009). Economical assessment of the design,
construction and operation of open-bed biofilters for waste gas treatment. Journal of
Environmental Management, 90(8), 2515–2523.
Sanchez C., Couvert A., Laplanche A. and Renner C. (2006). New compact scrubber for odour
removal in wastewater treatment plants. Water Science and Technology, 54(9), 45–52.
Shareefdeen Z. and Singh A. (2005). Biotechnology for Odor and Air Pollution Control.
Springer-Verlag, Heidelberg, Germany.
Stuetz R. M., Gostelow P. and Burgess J. E. (2001). Odour perception. In: Odours in
Wastewater Treatment: Measurement, Modeling and Control, R. Stuetz and F. B.
Frechen (eds), IWA Publishing, London, UK.
Sucker K., Both R. and Winneke G. (2008). Review of adverse health effects of odours
in field studies. 3rd IWA International Conference on Odour and VOCs, Barcelona,
Spain.
Tchobanoglous G., Burton F. L. and Stensel H. D. (2003). Wastewater Engineering:
Treatment and Reuse. McGraw Hill, New York, USA.
Turk A. and Bandosz T. J. (2001). Adsorption systems for odour treatment. In: Odours in
Wastewater Treatment: Measurement, Modeling and Control, R. Stuetzand and F. B.
Frechen (eds), IWA Publishing, London, UK.
UBS – United Bank of Switzerland (2012). Prices and Earnings: A Comparison of
Purchasing Power Around the Globe. UBS CIO WM Research, Switzerland.
Vincent A. J. (2001). Sources of odours. In: Odours in Wastewater Treatment: Measurement,
Modeling and Control, R. Stuetzand and F. B. Frechen (eds), IWA Publishing, London,
UK.
WHO, World Health Organization International Angency for Research on Cancer (2013).
Press Release Number 221, 17 October.
Zarra T., Naddeo V., Belgiorno V., Reiser M. and Kranert M. (2008). Odour monitoring of
small wastewater treatment plant located in sensitive environment. Water Science and
Technology, 58(1), 89–94.
Zhang L., De Gusseme B., Cai L., De Schryver P., Marzorati M., Boon N., Lens P. and
Verstraete W. (2011). Addition of an aerated iron-rich waste-activated sludge to control
the soluble sulphide concentration in sewage. Water and Environment Journal, 25(1),
106–115.
10.1 Introduction
Over the last decades, the operation of urban wastewater treatment plants
(WWTPs) has been mainly focused on maintaining the amount of active biomass
required for removing the influent organic matter within the system. For this
purpose, the most common operation and control strategies were oriented toward
regulating the suspended solids and dissolved oxygen concentrations needed
to guarantee biodegradation (Olsson, 1987). However, stricter requirements in
effluent quality (especially for nutrient removal in sensitive areas) have been
promoting new, more flexible and efficient WWTP plant layouts. New treatment
plants incorporate complex combinations of aerobic, anoxic, anaerobic and
facultative tank reactors, and their optimum operation cannot be based on the
traditional strategies (Andrews, 1993), which are frequently based on rigid
and conservative rules that do not consider the interactions between processes
and their dynamic characteristics. For these reasons, new Instrumentation,
Control and Automation (ICA) tools are being progressively incorporated into
plants in order to optimally govern (both from an environmental and economic
point of view) all the dynamic and interrelated mechanisms that appear in an
urban WWTP.
The need to meet stricter effluent requirements at a minimal cost is not the
only reason for the significant increase in ICA implementation in recent years.
New advances in monitoring and actuation equipment are also crucial. On the
one hand, new on-line monitoring sensors and analysers are more reliable and
economical, and on the other hand the new actuators incorporate enough flexibility
for implementing more efficient control strategies. In this sense, it is very important
to note that the dynamic behaviour of the WWTP processes should be taken
into account in the design and dimensioning stages in order to optimize future
operational costs.
Computation and data acquisition tools have also progressed significantly in
recent years. However, the lack of appropriate data management tools is presently
a limiting factor for a broader implementation and a more efficient use of sensors
and analysers, monitoring systems and process controllers in WWTPs. Therefore,
there is the challenge to develop new tools that are able to synthesize useful
information from the processing and analysis of combined data gathered from
several interrelated processes and deal with their heterogeneous characteristics
and storage decentralization (Beltrán et al. 2012).
It is clear that optimizing WWTP operation using automatic controllers
makes it possible to reduce operational costs. However, this reduction cannot
be generically evaluated because it depends on many factors like the size and
flexibility of the plant, the amount of available information, the wastewater
characteristics and the effluent requirements. This chapter focuses firstly on
the state-of-the-art of instrumentation for monitoring and control purposes.
Then it describes the main real-time control strategies for efficient sewage
treatment plant operation: control of the aeration system, control of chemical
addition, control of the internal, external and wastage flow-rates, and control
of the anaerobic processes. Finally, plant-wide control is addressed and several
conclusions are highlighted.
Properties Constituents
Conductivity Ammonium
pH Biogas production
Redox Dissolved oxygen
Sludge blanket Nitrate
Temperature Organic matter
Turbidity Phosphate
Solids concentration
Aeration is one of the key systems for WWTP stability, performance and
operational costs, and consequently dissolved molecular oxygen (DO) sensors
are considered particularly important. The first DO determination was performed
by L. W. Winkler in 1888 by using a colorimetric method based on titration of
oxygen with thiosulfate (S2O3−2) and iodine (I2) (Winkler, 1988). The amount of
the DO is proportional to the amount of tetrathionate (S4O6−2) generated, which
is determined by reducing I2 to iodide (I−). In spite of being difficult to use for
on-line sensing purposes, automatic measurement of DO based on potentiometric
determination of the I− produced has also been developed (Orellana et al. 2011)
and it is still employed as a reference method for calibrating DO sensors since it
is the most precise and reliable titrimetric procedure for DO analysis (Standard
Methods, 2012). Commercial electrochemical (e.g., polarographic, galvanic) DO
sensors were already being used for control on a routine basis from the early 1980s.
These sensors are protected by an oxygen-permeable plastic membrane that serves
as a diffusion barrier against impurities, which makes electrochemical DO sensors
suitable for analysis in situ and particularly convenient for field applications such
as the continuous monitoring of DO in activated sludge. However, DO control was
still far from fully utilised in the early 2000s (Ingildsen et al. 2002b; Jeppsson et al.
2002). Nowadays, even though galvanic cells are the dominant electrochemical
technology, the control of aeration systems based on luminescence-based oxygen
sensors is growing rapidly. The advantages over electrochemical devices include
the ease of miniaturization, the lack of chemical reactive agent consumption, faster
response (<60 s), robustness and insensitivity to interfering agents (e.g., H2S, CO2
or NHX-N). The low maintenance, extended operational lifetime and reliability of
optic oxygen sensors are so notable that every major manufacturer of environmental
monitors is currently offering at least one model for in situ DO measurements in
water, rapidly phasing out the membrane electrode sensors.
Nutrient analysers (e.g., NH X-N, NOX-N, PO4-P) were emerging at WWTPs in
the early 2000s (Olsson, 2005) and they now have developed into in situ sensors
(e.g., ion-selective electrodes probes for ammonia and ultraviolet probes for nitrate
and nitrite) for reasons of maintenance, costs, measurement delays and sensor
dynamics, resulting in easier control and better performance (Kaelin et al. 2008).
In the case of biological parameters, the on-line information is more reduced
and the most used analysers are the respirometers (Spanjers et al. 1998). Despite
the great expectations already present in the 1970s, there are not many controls
based on respirometry implemented in full scale (Trillo, 2004). Still, respirometry
is a viable method for in-stream early warning systems.
In anaerobic sludge digestion operation, gas production, gas quality, volatile acid
content, temperature, sludge feed rate, alkalinity and total organic carbon (TOC)
are usually measured (Spanjers & van Lier, 2006). The sludge blanket height in
the clarifiers is monitored today by reliable sensors that are used on a routine basis,
sometimes for the dynamic control of the return sludge (Vanrolleghem et al. 2006).
Further information of on-line measuring equipment in WWTPs is listed
in Rieger et al. (2003) and Vanrolleghem and Lee (2003). In addition, the
Water Environment Federation publishes an annual literature review about
instrumentation for monitoring and control purposes in WWTPs (e.g., Sweeney
& Kabouris, 2013).
In summary, it can be stated that instrumentation, including sensors, analysers
and other measuring instruments, is no longer the bottleneck for the control of
wastewater systems (Jeppsson et al. 2002). In fact, instrumentation for control
is no longer the main focus for international research. Currently, according to a
recent industrial marketing analysis, there are almost 100 sensor companies in
the world working with water (Olsson et al. 2013). Indeed, increased confidence
in instrumentation is now driven by the fact that clear definitions of performance
characteristics and standardised tests for instrumentation have become available
(ISO 15839:2003).
In a short time the research on sensors will become more focused by providing
relevant and reliable data on the problem at hand and they will deal with the
painstaking fouling problems whilst at the same time minimising maintenance
requirements. In addition, there is an exponential growth in the use of soft-sensors
for estimating process variables that considerably influence process behaviour but
cannot be measured on-line in a successful manner (e.g., active biomass, soluble
substrate). Such tools are being used as an effective utilisation for advanced control
and for the development of new optimization strategies, helping the operator or
a supervision system to take the appropriate actions to maintain the process in
good operating conditions, diagnose possible process failures or prevent accidents
(Haimi et al. 2013).
Finally, it can be said that although the instrumentation in wastewater treatment
systems has increased almost exponentially over the past decades, there is still no
standardized way to check data quality (fault detection) and know the source of
the error (diagnosis), even if a lot of progress has been made (Olsson et al. 2013).
While some years ago the scarcity of data due to the lack of reliable measuring
devices was a major limitation, at present WWTPs deal with extremely large
volumes of data. Data logging and SCADA tools manage thousands of data
from all points of the plant on a daily basis, meaning that the likelihood that
errors or faults will occur has greatly increased. Processing and managing these
heterogeneous, incomplete and frequently inconsistent data appropriately often
exceeds the capacity of the WWTP staff. Consequently, valuable plant information
for diagnosis and optimisation continues to be limited, in this case due to the excess
of data rather than the previous shortage. Although several methodologies and
tools have been developed in recent years to support decision making at WWTPs
(Beltrán et al. 2012), at present it is clear that the optimum operation of WWTPs
urgently requires advanced data management algorithms and tools to optimize the
global operation of wastewater systems by adequately managing and using all the
information available in the plant at every moment.
fluctuations in oxygen demand that appear through time and at the different
treatment zones. Another common limitation in keeping DO constant is motivated
by the need to avoid the sedimentation of the suspended solids in the bioreactors,
which requires a minimum air flow-rate value that, in some cases, could be
excessive during the low load periods, and above all at the final zones of the aerated
bioreactors. In addition to the extra energy cost, the peaks in DO associated with
these situations can be detrimental for the anoxic or anaerobic bioreactors because
of the highly oxygenated water recirculations.
The reduction of the air flow-rate that is achieved with the incorporation of a
simple feedback control loop at a constant DO set-point is difficult to evaluate in
a general way due to the high dependence of the influent load fluctuations. In any
case, a reduction of between 20% and 40% in the air flow-rate is very common
(Åmand et al. 2013), in addition to there being other possible advantages from the
point of view of process stability and possible nutrient removal.
A second level of aeration control, which is a bit more sophisticated, consists
of varying the DO concentration with the aim of optimising costs and nutrient
removal. The DO concentration in the aerated zones of the biological reactors is
the manipulated variable with quicker time response when it comes to regulating
biological activity. On the one hand, the DO concentration in the aerated zones
must be kept high enough to favour the growth of non-filamentous microorganisms,
to guarantee the needed nitrifying speed and to maintain the adequate mixing
characteristics in order to maintain solids in suspension and to assure that the oxygen
is well distributed throughout the reactors. On the other hand, the DO concentration
must be low enough to avoid wasting energy and an excessive agitation that could
fragment the biological flocs, and to minimise as much as possible the oxygenation
induced by the internal recirculations in the anoxic and anaerobic zones. It is clear
that the nitrification is enhanced for high DO concentrations, but operating a plant
with relatively low DO concentrations not only promotes denitrification in the anoxic
zones, but it also encourages a certain simultaneous denitrification-nitrification in
the aerated zones (Olsson & Newell, 1999).
When it comes to assessing the oxygen requirements and thus the reduction
in energy costs that can be achieved by manipulating the aeration, it is important
to consider that a specific load to be oxidised in aerobic conditions is associated
with an equivalent consumption of dissolved oxygen that cannot be, in principle,
substantially reduced without increasing the effluent residual load. This
requirement is purely stoichiometric, so it cannot be avoided. However, aeration
efficiency can usually be changed, since the same DO can be supplied by different
air flow-rate values, basically depending on the efficiency of the aeration devices
and the difference from the DO saturation value.
In this way, a simple control strategy for adapting the DO concentration set-point
to the nitrification needs is the incorporation of a control feedback loop (Upper
layer, Figure 10.2) that selects the DO reference needed by the lower control loop
(Lower layer) on a continuous basis. This is done by using the discrepancy between
the NHX-N set-point selected by the plant operator and the NHX-N that is measured
experimentally in the effluent. In this cascade, the three control loops work at
different time scales with the airflow control (see Figure 10.1) being the fastest.
This combined control scheme guarantees good performance.
air flow-rate in the range of 14% to 28%, depending on the NHX-N concentration
that was used as reference. The most likely improvement of the effluent nitrates
concentration was not studied.
A successful example of full-scale application was developed at the Galindo-
Bilbao WWTP, which has an average design flow-rate of 4 m3/s (1.5 MPE). The
specific design and preliminary verification of the most appropriate controllers for
the plant were based on the exhaustive and rigorous prior work of mathematically
modelling and simulating the system, which allowed the efficiency obtained to be
evaluated in a comparative way with different control strategies. The incorporation
of an automatic variation of the DO set-point in order to reach the required
24 h-averaged ammonia concentration at the end of the aerated volume enabled
the reduction of energy consumption in the range of 15% to 20% and effluent
nitrates (and consequently in total nitrogen) in 2.0 g N/m3 (Ayesa et al. 2006). The
different phases of the project lasted 8 years, combining model simulations, pilot-
plant experimentation and full-scale validation.
When there is a benefit to reacting quickly to a disturbance (e.g., to the NH X-N
influent load), such as in the case of ‘never-to-exceed’ effluent limits, feedforward
control can be used. In order to do this, more sensors and a mathematical model
of the controlled system are needed. Using feedforward without a feedback loop
is not recommended since feedback contributes to a more robust performance in
light of feedforward model uncertainty and it can compensate for non-modelled
disturbances. Essentially, there are two ways to use a feedforward control in an
effort to reduce effluent ammonia peaks: changing the aeration intensity (e.g., by
varying the DO set-point, see Figure 10.3), or adding more aerated volume (e.g., by
switching on aeration in a swing zone). However, the latter is preferred since the
nitrification capacity of the system may be exceeded (Rieger et al. 2012).
points generated by the transition among the oxic, anoxic and anaerobic phases
(e.g., the ORP ‘elbows’ that appear when the DO grows suddenly at the end of
the nitrification process and when the NOX-N run out at anoxic conditions and
anaerobic conditions start).
On the other hand, pH profiles also show variations motivated by the biochemical
transformations of nitrification and denitrification (e.g., ‘the NHX-N valley’ that
appears at the end of the nitrification and ‘the NOX-N peak’ that occurs at the end
of the denitrification). In addition to the trajectory variations, several controllers
use the absolute value of the measurements of ORP or pH, but it is a risky practice
because of the difficulty in choosing suitable values in each case. On the other hand,
it is clear that optimising the duration of each cycle phase induces an improvement
in the effluent quality, an increase in treatment capacity and therefore a reduction
in the total operating costs.
The economic impact of the pumping costs associated with the internal
recirculation is really low because the height elevation of the water is low even
though the flow-rate could be high. Moreover, the implementation of this controller
does not have a direct economic implication since its effect is more focused on the
optimisation of the process and, in particular, on the optimisation of the use of
the plant denitrification potential. However, this optimisation always has indirect
economic consequences for the aeration costs since all the COD that is used under
anoxic conditions will not consume oxygen and will allow savings in the carbon
external addition when this is necessary.
When the limited factor of the denitrification process is the organic matter
and when internal addition is not possible (sludge supernatants, acid fermentation,
etc.), the most widespread practice today consists of applying an external carbon
source (e.g., methanol, ethanol, acetic, etc.), which enhances the denitrification rate
and the nitrate removal, especially in the case of WWTPs that treat wastewater
with an unfavourable COD/N ratio.
The control strategies with constant dosage do not allow the process operation
to be optimised since at high nitrogen loads or low temperature periods, an
immediate increase of the denitrification rate would be desirable, while at low load
periods the external carbon requirements would be lower. Therefore, there is a
need for regulating the quantity of the external carbon to be dosed into the system
in an automatic way.
When automatic control of nitrate recirculation is not available, the most common
strategy consists of regulating the carbon dosage so that the nitrate concentration at
the end of the anoxic zone is kept low. In this case, the internal recirculation flow-
rate is set to a level in which the average effluent nitrate concentration is kept within
the appropriate limits. Since nitrates are in the anoxic zone when the carbon source
is added, this strategy guarantees that the carbon source is instantaneously used for
the denitrification. However, although an average effluent nitrate concentration can
be fixed beforehand, this strategy does not have direct control on the effluent nitrate
concentration as it depends on the influent variations of COD and N. Moreover, this
strategy does not guarantee the maximum use of the influent COD for denitrification.
For example, when the nitrogen load is low, part of the anoxic zone is changed to aerobic,
even when external carbon is not added, due to an inappropriate nitrate recirculation.
The disadvantages associated with individual control of the nitrate recirculation
and the dosage of external carbon can be overcome by an integral control of
both variables. The objectives of integral control are focused on meeting the
instantaneous as well as the medium- and long-term requirements on the effluent
nitrate concentration, maximising the use of influent COD for the denitrification
and minimising the dosage of external carbon. For example, the control proposal
in Yong et al. (2006) showed through simulation an average reduction of 42%–
47% in the effluent nitrate concentration.
It is not possible to make a direct quantitative evaluation of the benefits that are
obtained when an external carbon dosage control loop is implemented since they
clearly depend on the plant characteristics. It is clear that the automatic control
loops allow the dosage to be adjusted to the one that is strictly necessary for
meeting the effluent requirements, avoiding excessive dosages that could involve a
significant direct cost (because of the additive price) and an additional indirect cost
associated with the subsequent aerobic oxidation of the carbon excess.
result it has a direct effect on the solids that are attained in the process and in their
distribution. Each plant has its specific range of SRT since a longer SRT results in a
higher concentration of suspended solids and thus increases the solids loading into
the settler. A concentration that is too high could cause an elevated concentration of
suspended solids in the effluent, especially when coupled with high hydraulic loads.
On the other hand, a minimum SRT is needed to guarantee the necessary active
biomass. For example, in the case of nitrogen removal plants, a minimum SRT value
is essential in order to achieve nitrification, especially at low temperatures.
In a general way, higher SRTs mean lower sludge production, and therefore the
costs associated with sludge treatment are lower. However, the potential benefits
that can be achieved by methane production in the anaerobic digestion are also
lower. On the other hand, the oxygen required by the microorganisms increases
with sludge age. This increase is logically associated with a higher reduction in
the contaminant load of COD and nitrogen and with an increase in the biomass
endogenous respiration. Selecting the optimum SRT set-point for the different
operational conditions of a WWTP should take into account the intrinsic inevitable
limitations of the process (nitrification safety and maximum solids load to the
settler) and operating costs criteria (sludge treatment, aeration costs, etc.).
There are different strategies for controlling the sludge wastage flow-rate. One
option is to maintain a constant SRT via a feedforward loop, complemented with a
loop that stops the wastage flow-rate when the average effluent ammonia exceeds a
maximum fixed value (Olsson et al. 2005). Another option is to keep the solids mass
under a fixed range, such as the one that was successfully implemented in the Galindo-
Bilbao WWTP (Ayesa et al. 2006) and in the Mekolalde WWTP (Irizar et al. 2014).
10.6.1 Technological barriers
Although the benefits of an extensive application of advanced monitoring and
control solutions on the anaerobic processes are clear, faster development of control
products in this technology has been hampered for a number of reasons:
– Lack of on-line sensors. In the last decade, considerable research has
focused on the consolidation of a sensor technology specific to anaerobic
digesters. However, the so-called advanced instrumentation is still not
sufficiently reliable for use in full-scale real applications (Spanjers & van
Lier, 2006).
– The design and development of industrial control products involves real
aspects such as the physical and operational constraints of actuators,
disturbances in the form of sensor faults, noise, and so on, that need to
be considered at the outset (Liu et al. 2004a). This means extrapolating
control solutions that have been satisfactorily validated at laboratory or
pilot scale is not straightforward.
– The intricate non-linear nature of anaerobic digestion phenomena and the
intrinsic uncertainty of their mathematical formulations make the design of
control solutions a non-trivial issue.
In addition to this, given that the feed flow-rate is the only manipulated variable
used for control, a surprising conclusion drawn from all these works is the lack of
consensus about which output signals (sensors) must be linked to this actuator. As
shown below, the list of suggested signals is relatively long: from pH, intermediate
alkalinity, hydrogen gas, dissolved hydrogen, volatile fatty acids (VFAs), effluent
COD, to methane gas, and so on.
Cord-Ruwisch et al. (1997), for instance, propose a very simple proportional
controller that automatically adjusts the feeding flow-rate by trying to keep the
dissolved hydrogen close to a pre-set reference. Antonelli et al. (2003) suggest
a similar control scheme where the dissolved hydrogen signal is replaced by
the methane gas flow-rate. In the same way, Rodríguez et al. (2006) argue
that hydrogen gas, being an intermediate process variable, becomes especially
appropriate for the early detection of abnormal situations (overloads, drops in
reactor temperature, etc.). On the basis of this argument, they design a very simple
control law to keep the hydrogen gas concentration around a pre-set reference.
Moreover, the non-linearity of the process is taken into account through a real-time
variable gain algorithm dependent on the instantaneous measurements of hydrogen
gas concentration and methane gas flow-rate. A more sophisticated cascade control
scheme is described in Álvarez-Ramírez et al. (2002), where the control signals for
the inner and outer loops are the concentration of VFA in the reactor and effluent
COD, respectively.
Fuzzy logic is another technique that has received considerable acceptance
in control solutions for anaerobic reactors. Just to cite a few examples, Estaben
et al. (1997) implemented this technique to control the pH and the biogas flow-
rate. Similarly, García et al. (2007) applied a fuzzy-based real-time system in a
pilot plant, which controls the anaerobic digester by monitoring the following
process variables: the ratio between the intermediate and total alkalinities (IA/
TA), methane gas production, the concentration of hydrogen in the gas headspace
and its rate of change.
Adaptive control techniques have also been used in control solutions for
anaerobic processes. An example is the adaptive controllers with linearisation
designed by Bernard et al. (2001) to control the IA/TA ratio and TA in an anaerobic
filter. Similarly, by using a simplified model to reproduce plant behaviour, a non-
linear adaptive controller to control the effluent COD was developed by Mailleret
et al. (2004). Méndez-Acosta et al. (2007) suggested a model-based non-linear
controller to regulate the effluent TOC and the VFA in the reactor. Although the
above approaches are of great interest in the academic research, their application
in real-world plants currently faces a major barrier: it is extremely difficult to
develop reduced models that guarantee good predictions for all operations taking
place in a WWTP.
A common feature of all the above control solutions is that the ultimate goal is
to keep one or more process variables close to a fixed reference value. Moreover, a
different control approach for anaerobic systems also found in the literature focuses
its objective on maximising a process variable, usually the methane gas production
rate. These strategies, known as extremum-seeking, are especially appropriate when
it comes to controlling non-linear systems (Krstic, 2000). Moreover, extremum-
seeking controllers are classified in two main groups: (1) those that need a predictive
model to optimise the objective variable (Marcos et al. 2004; Guay et al. 2005);
and (2) those that apply external excitation in order to diagnose whether or not the
‘objective’ has reached its maximum value. It is actually its model-free condition
that makes this second group attractive from a practical point of view.
Reported works on model-free extremum-seeking controllers differ from each
other in terms of the pattern followed for external excitation. Thus, in Steyer et al.
(1999) a periodic pulse is superimposed on the feed rate and its effect on the
methane production observed in order to determine whether or not the organic
loading rate can be increased. In contrast, in Simeonov et al. (2004, 2007) a
sinusoidal excitation is applied in combination with an adaptation law for the
feeding flow-rate that is proportional to the gradient of the biogas production.
Liu et al. (2004c) proposed a cascade control scheme consisting of an inner loop
controlling the pH of the reactor at a pre-set value which is determined by an outer
loop. The outer loop compares the biogas production rate of the reactor with a
reference value, set by a rule-based supervision module on a regular basis in order
to automatically steer the process towards the maximum biogas production. The
validation of this control solution in a lab-scale anaerobic reactor fed with synthetic
wastewater reveals very good disturbance rejection properties as well as potential
to keep the process stable even at organic loading rates above 25 kg COD/m3 ⋅ d.
Some years later, the same control architecture was adopted and extended (Figure
10.5) by Alferes and Irizar (2010) to prove that the combination of extremum-
seeking controllers and optimum management of the equalisation tanks preceding
the anaerobic reactor leads in the medium/long-term to a significant increase in the
production of biogas. Comparative simulation results with respect to conventional
operation reveal improvement of about 20% in both the effluent quality and the
biogas yield. Nonetheless, the extent to which these estimates are valid during the
operation of a full scale plant still needs to be verified.
The economic evaluation of the automatic control of anaerobic digesters has
to be studied in function of the characteristics of each WWTP. For example, the
possibility of including pre-treatment steps that hydrolyse the sludge or digest the
sludge with other external substrates confirms the interest of having automatic
tools that allow the digester to work at its maximum efficiency. It is clear that
maximising methane production yields an important energy recovery (1.0 m3N of
methane is equivalent to 8570 kcal, i.e., it is equivalent to 1.15 l of gasoline, 1.3 kg
of coil, 0.94 m3N of natural gas or 9.7 kWh of electricity). However, it is also clear
that the potential of generating methane in WWTPs is determined by the quantity
and type of the available sludge. Therefore, a plant-wide control strategy operating
the system as a whole with a global objective should be considered instead of trying
to optimise the operation of the individual units separately.
Figure 10.5.
Diagram of a control strategy for the maximisation of methane
production.
10.7 Plant-Wide Control
In order to address the control of WWTPs in a reasonable way, it is necessary to
break the problem down into smaller parts, taking advantage of the separation
among lines and process units, as well as the different time responses of the
physical, chemical and biological mechanisms of the process. This allows the
control strategies to be organised in a hierarchy and enables the optimisation of
each subprocess to be resolved separately.
However, it is clear that the optimisation of the operation of a WWTP when
understood as a global system does not, in principle, have to be equivalent to the
result of optimising each one of the elements and process units (e.g., primary
settler, secondary biological reactors, anaerobic digesters, etc.). Therefore, the
global optimisation of the system has to take the interactions between the different
parts of the plant into account and use them to find the operational and control
strategies that optimise a global criterion. In some cases, this could mean that some
parts of the WWTP could be operated in a suboptimal way, while still contributing
into the plant operation an acceptable way or near the global optimum.
Speaking of the integrated control of sanitary systems, the first point of
improvement is taking advantage of the relation between the networks of drainage
and sanitation (including the sewers, retention tanks and storm tanks) and the
WWTPs. Taking the undesirable effects of the load variations in the WWTP into
account, an appropriate management of the hydraulic retention capacity of a sewer
network may induce a significant increase in the treatment capacity. The experience
gained with the combined operation of sewers, storm tanks and WWTPs shows
that integrating all these parts in a control strategy has great potential.
However, even in the WWTP itself, there is wide margin for designing operation
strategies and loops of automatic control that would comprise an integrated and
flexible control system. This is the so-called ‘plant-wide control’.
The interactions among the different unit processes of the WWTPs are
realised through the recirculation of flow-rates that physically link them. There
are many interactions of this type among the different units of a WWTP, such as
the recirculation of oxygen and nitrates to the anaerobic zones from the sludge
drying systems, the recirculation of oxygen to the anoxic zones because of the
nitrate recirculation or the nutrient recirculation from the supernatant of the
sludge treatment. It is clear that the disturbances each recirculation imposes on
the process units have to be taken into account for the optimal control of the
whole system. In other words, the control set-points of each unit process have to
be supervised by taking into account the most relevant interactions among the
different control loops.
This integral operation also involves rethinking of the objectives, taking
into account the whole system. A clear example could be the discussion of the
global objective of sludge production (or sludge retention time), considering the
requirements needed for appropriate nutrient elimination, the costs of the sludge
treatment or energy recovery through methane or incineration. Another point of
great interest when global plant operation is considered is the decision regarding
the optimum use of organic matter in each unit process, considering, for example,
the sedimentation in the primary settlers, the operating costs associated with
aeration in the secondary treatment, the need for carbon for nutrient elimination
and the possibilities of energy recovery in the anaerobic digestion. These examples
clearly show the need for supervision of the local control loops by a global
supervision strategy that optimally manages the flexibility of the whole system.
Finally, it is important to note that the integral control of a WWTP is the one
that provides real possibilities for increasing the capacity of the global treatment
and as a result removing or postponing new investments when the size of the
WWTP is not sufficiently large. This can lead to important economic savings.
10.8 Conclusions
The full-scale implementation of automatic control strategies in WWTPs has been
severely limited for years because of a limitation in the sturdiness and reliability
of on-line monitoring devices and the operating flexibility of plants. However, in
recent years new analysers that are more reliable and require less maintenance
have been developed. In this way, the problem of the instrumentation is no longer
the bottleneck to progressive implantation of advanced automation in WWTPs.
On the other hand, the new treatment technologies, in addition to providing higher
efficiency, also provide higher operating flexibility, which can and must be taken
advantage by the automatic control strategies.
The concept of automatic control in WWTPs is closely linked to plant-wide
optimisation of the process, with all the implications that this has for the improvement
of treatment efficiency and the reduction of the operating costs. The results of
this optimisation are starting to be evident now. On the one hand, the long list of
simulation studies have demonstrated the great advantages that are derived from
the implantation of automatic control strategies in WWTPs, which also enable the
detection of the most important points of improvement and a comparative analysis of
the efficiency of the different control alternatives. On the other hand, the results of
applying controllers are no longer limited to simulation studies. Instead the reports
that illustrate their successful implementation at full scale are increasing every
day, corroborating the improvement in the stability and efficiency of the process.
It can be expected that in a few years’ time the majority of the controllers whose
efficiency has been successfully evaluated by simulation will be experimentally
validated.
The economic implications resulting from the implantation of automatic
controllers in new WWTPs are increasingly more evident. Some control loops,
such as the ones that optimise aeration (at constant or variable set points), the
dosing of additives for physical and chemical precipitation and the dosing of
methanol for denitrification are demonstrating their capability to not only improve
process performance but also substantially reduce the operating costs and recover
initial investments in a short time. The cost reduction associated with other
automatic control loops is more difficult to quantify, since it significantly depends
on the specific characteristics of each plant, such as flexibility, treatment type,
size, effluent quality requirements, and so on. Therefore, it is highly recommended
that a simulation study be undertaken beforehand, making it possible to quantify
the expected benefits of each control strategy and prioritize the different available
alternatives, depending on the needs and limitations of each WWTP. It is also
important to note the interest in the automatic control loops for improving the
stability and general governability of the process, facilitating the final objective of
‘global optimisation’ of the WWTP.
The lines for the future developments of ICA tools are many and varied. Some
examples of the major challenges that have to be faced in the near future and are
being actively worked on now include the optimum operation of aeration systems
that minimises the energetic costs, the design of the most appropriate control
strategies that optimises the biological elimination of phosphorus, the adaptation
of the controllers to small-sized plants in which simplicity and toughness must take
precedence, and the plant-wide control that integrates the operation of the water
and sludge lines. In order to efficiently respond to these challenges it is necessary
that the research groups who are experts in the control and operation of plants
closely collaborate with the water engineering and control companies and the
entities that manage WWTPs.
10.9 References
Alferes J. and Irizar I. (2010). Combination of extremum-seeking algorithms with effective
hydraulic handling of equalization tanks to control anaerobic digesters. Water Science
and Technology, 61(11), 2825–2834.
Álvarez-Ramírez J., Meraz M., Monroy O. and Velasco A. (2002). Feedback control
design for an anaerobic digestion process. Journal of Chemical Technology and
Biotechnology, 77, 725–734.
Andrews J. F. (1993). Modeling and simulation of wastewater treatment process. Water
Science and Technology, 28(11–12), 141–150.
Antonelli R., Harmand J., Steyer J. P. and Astolfi A. (2003). Set-point regulation of an
anaerobic digestion process with bounded output feedback. IEEE Transactions on
Control Systems Technology, 11(4), 495–504.
Åmand L., Olsson G. and Carlsson B. (2013). Aeration control – a review. Water Science
and Technology, 67(11), 2374–2398.
Ayesa E., de la Sota A., Grau P., Sagarna J. M., Salterain A. and Suescun J. (2006).
Supervisory control strategies for the new WWTP of Galindo-Bilbao: the long run
from the conceptual design to the full-scale experimental validation. Water Science
and Technology, 53(4–5), 193–201.
Batstone D. J., Gernaey K. V. and Steyer J. P. (2004). Instrumentation and control in
anaerobic digestion In: Proeedings of the 2nd IWA Leading-Edge Water and
Wastewater Treatment Technologies (LET 2004), Prague, Czech Republic, 1–4 June.
Beltrán S., Maiza M., de la Sota A., Villanueva J. M. and Ayesa E. (2012). Advanced data
management for optimising the operation of a full-scale WWTP. Water Science and
Technology, 46(2), 314–320.
Bernard O., Polit M., Hadj-Sadok Z., Pengov M., Dochain D., Estaben M. and Labat P.
(2001). Advanced monitoring and control of anaerobic wastewater treatment plants:
software sensors and controllers for an anaerobic digester. Water Science and
Technology, 43(7), 175–182.
Bourgeois W., Burgess J. E. and Stuetz R. M. (2001). On-line monitoring of wastewater
quality: a review. Journal of Chemical Technology and Biotechnology, 76(4), 337–348.
Brandt M., Middleton R., Wheale G. and Schulting F. (2011). Energy efficiency in the water
industry, a global research project. Water Practice & Technology, 6(2).
Cord-Ruwisch R., Mercz T. I., Hoh C. Y. and Strong G. E. (1997). Dissolved hydrogen
concentration as an on-line control parameter for the automated operation and
optimization of anaerobic digesters. Biotechnology and Bioengineering, 56(6), 626–634.
Estaben M., Polit M. and Steyer J. P. (1997). Fuzzy control for an anaerobic digester. Control
Engineering Practice, 5(98), 1303–1310.
García C., Molina F., Roca E. and Lema J. M. (2007). Fuzzy-based control of an anaerobic
reactor treating wastewaters containing ethanol and carbohydrates. Industrial and
Engineering Chemistry Research, 46, 6707–6715.
Garrett M. T. (1998). Instrumentation, control and automation progress in the United States
in the last 24 years. Water Science and Technology, 37(12), 21–25.
Guay M., Dochain D. and Perrier M. (2005). Adaptive extremum-seeking control of
nonisothermal continuous stirred tank reactors. Chemical Engineering Science,
60(13), 3671–3680.
Haimi H., Mulas M., Corona F. and Vahala R. (2013). Data-derived soft-sensors for
biological wastewater treatment plants: an overview. Environmental Modelling &
Software, 47, 88–107.
Ingildsen P., Jeppsson U. and Olsson G. (2002a). Dissolved oxygen controller based on
on-line measurements of ammonium combining feed-forward and feedback. Water
Science Technology, 45(4–5), 453–460.
Ingildsen P., Lant P. and Olsson G. (2002b). Benchmarking plant operation and
instrumentation, control and automation in the wastewater industry. Water Supply,
2(4), 163–171.
Ingildsen P. and Wendelboe H. (2003). Improved nutrient removal using in situ continuous
on-line sensors with short response time. Water Science and Technology, 48(1), 95–102.
Irizar I., Beltrán S., Urchegui G., Izko G., Fernández O. and Maiza M. (2014). Lessons
learnt from the application of advanced controllers in the Mekolalde WWTP: good
simulation practices in control. Water Science and Technology, 69(6), 1289–1297.
ISO 15839 (2003). Water quality – on-line sensors/analysing equipment for water –
Specifications and performance tests. International Organization for Standardization.
Jeppsson U., Alex J., Pons M. N., Spanjers H. and Vanrolleghem P. A. (2002). Status and
future trends of ICA in wastewater treatment – a European perspective. Water Science
and Technology, 45(4–5), 485–494.
Kaelin D., Rieger L., Eugster J., Rottermann K., Banninger C., Siegrist H. (2008). Potential
of in-situ sensors with ion-selective electrodes for aeration control at wastewater
treatment plants. Water Science and Technology, 58(3), 629–637.
Krstic M. (2000). Performance improvement and limitations in extremum seeking control.
Systems and Control Letters, 39, 313–326.
Liu J., Olsson G. and Mattiasson B. (2004a). Monitoring and control of an anaerobic
up-flow fixed bed reactor for high loading rate operation and rejection of disturbances.
Biotechnology and Bioengineering, 87(1), 43–53.
Liu J., Olsson G. and Mattiasson B. (2004b). Towards an economically competitive
anaerobic degradation process – an ICA approach. In: Proceedings of the 10th IWA
Anaerobic Digestion Congress (AD10), Montreal, Canada, 29 Aug.–2 Sept.
Liu J., Olsson G. and Mattiasson B. (2004c). Control of an anaerobic reactor towards
maximum biogas production. Water Science and Technology, 50(11), 189–198.
Lynggaard-Jensen A. (1999). Trends in monitoring of waste water systems. Talanta, 50(4),
707–716.
Mailleret L., Bernard O. and Steyer J. P. (2004). Non-linear adaptive control for bioreactors
with unknown kinetics. Automatica, 40, 1379–1385.
Marcos N., Guay M. and Dochain D. (2004). Output feedback adaptive extremum seeking
control of a continuous stirred tank bioreactor with Monods kinetics. Journal of
Process Control, 14(7), 807–818.
Méndez-Acosta H., Steyer J. P., Femat R. and González-Álvarez V. (2007). Robust nonlinear
control of a pilot-scale anaerobic digester. Dynamics and Control of Chemical and
Biological Processes, LNCIS 361, 165–199.
Olsson G. (1987). Dynamic Modelling, Estimation and Control of Activated Sludge
Systems. In: Systems Analysis in Water Quality Management, M. B. Beck (ed.),
Pergamon Press, Oxford, UK.
Olsson G., Aspegren H. and Nielsen M. K. (1998). Operation and control of wastewater
treatment – a scandinavian perspective over 20 years. Water Science and Technology,
37(12), 1–13.
Olsson G. and Newell B. (1999). Wastewater Treatment Systems. Modelling, Diagnosis and
Control, IWA Publishing, London, UK.
Olsson G., Ingildsen P., Jeppsson U., Kim C. W., Lynggaard-Jensen A., Nielsen M. K.,
Rosen C., Spanjers H., Vanrolleghem P. A. and Yuan Z. (2004). Instrumentation,
Control and Automation – Hidden Technologies in Wastewater Treatment. Keynote
paper. In: Proceedings of the 2nd IWA Leading-Edge Water and Wastewater Treatment
Technologies (LET 2004), Prague, Czech Republic, 1–4 June.
Olsson G., Nielsen M. K., Yuan Z., Lynggaard-Jensen A. and Steyer J.-P. (2005).
Instrumentation, Control and Automation in Wastewater Systems. Scientific and
Technical Report No. 15, IWA Publishing, London, UK.
Olsson G. (2012). ICA and me – A subjective review. Water Research, 46(6), 1585–1624.
Olsson G., Carlsson B., Comas J., Copp J., Gernaey K. V., Ingildsen P., Jeppsson U.,
Kim C., Rieger L., Rodríguez-Roda I., Steyer J.-P., Takács I., Vanrolleghem P. A.,
Vargas A., Yuan Z. and Åmand L. (2013). Instrumentation, Control and Automation
in wastewater–from London 1973 to Narbonne 2013. In: Proceedings of the 11th
IWA Conference on Instrumentation Control and Automation (ICA2013), 18–20
September, Narbonne, France.
Orellana G., Cano-Raya C., López-Gejo J. and Santos A. R. (2011). Online monitoring
sensors. In: Treatise on Water Science, Volume 3, Aquatic Chemistry and Biology,
Peter Wilderer (ed.), Elsevier B. V., pp. 221–261.
Rieger L., Alex J., Winkler S., Boehler M., Thomann M. and Siegrist H. (2003). Progress in
sensor technology – progress in process control? Part I: sensor property investigation
and classification. Water Science and Technology, 47(2), 103–112.
Rieger L., Jones R., Dold P. and Bott C. (2012). Myths about ammonia feedforward aeration
control. In: Proceedings of the 85th Water Environment Federation’s Annual Technical
Exhibition and Conference (WEFTEC2012), New Orleans, USA, September 29 –
October 3.
Rodríguez J., Ruiz G., Molina F., Roca E., Lema J. M. (2006). A hydrogen-based variable-
gain controller for anaerobic digestion processes. Water Science and Technology,
54(2), 57–62.
Sedlak R. I. (1991). Phosphorus and Nitrogen Removal from Municipal Wastewater:
Principles and Practice. 2nd edn, CRC Press, New York, USA.
Simeonov I., Noykova N. and Stoyanov S. (2004). Modelling and extremum seeking
control of the anaerobic digestion. In: IFAC workshop DECOM-TT, Bansko, Bulgaria,
October 3–5.
Simeonov I., Noykova N. and Gyllenberg M. (2007). Identification and extremum seeking
control of the anaerobic digestion of organic wastes. Cybernetics and Information
Technologies, 7(2), 73–84.
Spanjers H., Vanrolleghem P. A., Olsson G. and Dold P. (1998). Respirometry in Control
of Activated Sludge Processes. IAWQ Scientific and Technical Reports, London,
UK.
Spanjers H. and van Lier J. B. (2006). Instrumentation in anaerobic treatment – Research
and practice. Water Science and Technology, 53(4–5), 63–76.
Standard methods for the examination of water and wastewater (2012). 22nd ed, American
Public Health Association/American Water Works Association/Water Environment
Federation, Washington DC, USA.
Steyer J. P., Buffiere P., Rolland D. and Molleta R. (1999). Advanced control of anaerobic
digestion processes through disturbances monitoring. Water Research, 33(9),
2059–2068.
Suescun J., Ostolaza X., García-Sanz M. and Ayesa E. (2001). Real-time control strategies
for predenitrification–nitrification activated sludge plants: biodegradation control.
Water Science Technology, 43(1), 209–216.
11.1 Introduction
There are different problems of using fossil fuels, which meet about 80% of the
world energy demand. One is that they are limited in amount and sooner will
be depleted and the other is that fossil fuels are causing serious environmental
problems (air pollution, ozone layer depletion and global warming). Therefore,
together with strong improvements in energy conservation and efficiency, new
technologies are needed to gradually replace fossil sources by renewable ones.
Based on that, in the last years, the interest in fuel cells has increased dramatically,
due to their high efficiencies, low or zero emissions of pollutants, simplicity and
absence of moving parts. Particularly, the interest on microbial fuel cells (MFCs)
has been growing, because they are able to simultaneously treat and produce
electricity directly from the wastes that our society produces and can degrade toxic
compounds and pollutants (phenol, sulfates and chromium). The MFCs operate
at ambient temperatures, are fueled by organic matter which is neither toxic as
methanol nor explosive as hydrogen and can contribute to optimize the overall
efficiency of wastewater treatment facilities.
The problem with MFCs is that they are technically still very far from attaining
acceptable levels of power output, since the performance of this type of fuel cells is
affected by limitations based on irreversible reactions and processes occurring both
on the anode and cathode side. Electricity generation in a MFC is accomplished by
potential is always lower than its maximum value due to three irreversible loss
types: the activation, the Ohmic and the concentration losses (Figure 11.2).
The activation loss is dominant at low current densities and is due to the
activation energy that must be overcome by the reacting species. Phenomena
involving adsorption or desorption of reactant species, transfer of electrons and the
nature of the electrode surface contribute to the activation polarization. In MFCs
where microbes do not readily release electrons directly on the anode electrode
surface, activation polarization gets lower when mediators are added to the anode
compartment. In the middle of the operating range, the predominant loss is the
Ohmic loss and is due to ionic and electronic conduction. This loss can be reduced
by shortening the distance between the two electrodes and by increasing the ionic
conductivity of the electrolytes. At very high current densities, the major loss is
the concentration loss and is due to the inability to maintain the initial substrate
concentration in the bulk fluid, mainly due to mass transport limitations. As can be
seen in Figure 11.2, the polarization curve of a MFC indicates the various loss types
and their extent over the current density range, pointing to possible measures to
minimize them and approach the maximum potential. These measures may include
selection of new microbes, mediators, substrates, modifications in the MFC design
(short spacing between electrodes) and configurations (improvement in electrode
structure, new catalyst and electrolytes). Besides the three major losses described
that lead to a decrease in power output, on the anode compartment of a MFC
methanogens compete with electrochemically active microorganisms to convert
organic material in methane, reducing the electricity generation process. Therefore
it is extremely important to study the effect of the different operating conditions on
methane production in order to avoid that and increase the power production.
In the earliest MFC concept, electrical energy was produced from living
cultures of Escherichia coli and Saccharomyces spp. by using platinum electrodes,
although, the power output was very low. This concept only generated much
interest when it was discovered that current density and power output could be
greatly enhanced by the addition of electron mediators (Ieropoulos et al. 2005;
Du et al. 2007). Good mediators should be able to cross the cell membrane easily,
possess a high electrode reaction rate, should be non-biodegradable and non-toxic
to microbes and have low cost. Typical synthetic exogenous mediators include dyes
and metallorganics (neutral red (NR), methylene blue (MB), Fe(III)EDTA), but
their high toxicity and cost, instability and low efficiency limit their application
in MFCs (Ieropoulos et al. 2005; Du et al. 2007). These facts made scientists
look more closely at the microbiological features in the anode and to use bacteria
that could transfer electrons directly to the electrode and increase the Coulombic
efficiency, the so-called electricigens (Geobacter and Rhodoferax spp.) (Lovley
& Nevin, 2008). The use of these species is of interest because closely related
microorganisms have been found on the anodes of fuel cells harvesting electricity
from aquatic sediments and complex wastes (Holmes et al. 2004; Jung & Regan,
2007). Furthermore, these species have the ability to completely oxidize organic
substrates to carbon dioxide with an anode serving as the electron acceptor.
Also these types of bacteria develop biofilms on the MFC electrodes, allowing
considerable conversion capacity, and thus a potentially efficient microbial system
to enhance the electricity (Rittmann, 2008; Lovley, 2008). Despite the efforts made
in order to select the best bacterial consortia to achieve power densities needed
for MFC applications, the performances of this type of fuel cells still remain
lower than the ideal one. For a MFC it is also possible to enhance performance by
optimization of the operating conditions, such as electrode materials, fuel type,
proton transfer material, substrate concentration and feed rate, pH, temperature
and hydrodynamic stress.
11.4 Scale Up
Most MFCs studies were conducted at lab scales, however in order to make MFCs
suitable for real applications, such as wastewater treatment plants, it is critical
to achieve high power densities at a large scale. Also, for scaling up a MFC it
is mandatory to develop low-cost, simple construction and easy to maintenance
systems that can generate high power outputs. Many efforts have been made in
order to achieve these goals (Tender et al. 2008; Ieropoulos et al. 2008; Dewan
et al. 2008; Clauwaert et al. 2008; Dekker et al. 2009; Fan et al. 2010; Logan,
2010; Liu et al. 2008; Zhang et al. 2013; Zhu et al. 2013; Cheng & Logan, 2011;
Cheng et al. 2014; Zhuang & Zhou, 2009; Liu et al. 2008; Zhuang et al. 2009,
2010b, 2012a, 2012b; Ieropoulos et al. 2010a, Kim et al. 2012; Gurung et al. 2012;
Zuo et al. 2007, 2008; Lefebvre et al. 2009; Heijnea et al. 2011), but until now, no
consensus was found. The scale-up process can be accomplished by connecting
together single small units or by increasing the volume of a single unit (Tender
et al. 2008; Ieropoulos et al. 2008; Dewan et al. 2008; Clauwaert et al. 2008;
Dekker et al. 2009; Fan et al. 2010; Logan, 2010; Liu et al. 2008; Zhang et al.
2013; Zhu et al. 2013; Cheng & Logan, 2011; Cheng et al. 2014; Zhuang & Zhou,
2009; Liu et al. 2008; Zhuang et al. 2009, 2010b, 2012a, 2012b; Ieropoulos et al.
2010a, Kim et al. 2012; Gurung et al. 2012; Zuo et al. 2007, 2008; Lefebvre et al.
2009; Heijnea et al. 2011).
The first demonstration of using a microbial fuel cell as an alternative and viable
power supply for a low power consuming application was reported by Tender et al.
(2008). The specific application reported was a meteorological buoy that measures
different parameters such as air temperature, pressure and relative humidity, and
water temperature. The MFC employed in this demonstration was the benthic
microbial fuel cell (BMFC), which operates at the bottom of marine environments.
It is maintenance free, does not deplete and is sufficiently powerful to operate a
wide range of low power marine deployed scientific instruments normally powered
by batteries.
Despite this first attempt, bringing the MFC technology out of laboratory
appears as a challenge in the MFC developments, due to the high power densities
and low cost materials required. Also, the reactor configuration, the operation at
large scale, the electrode performance and the longevity are key factors in MFC
scale-up. In the last years, some efforts have been made in order to scale-up MFCs
(Tender et al. 2008; Ieropoulos et al. 2008; Dewan et al. 2008; Clauwaert et al.
2008; Dekker et al. 2009; Fan et al. 2010; Logan, 2010; Liu et al. 2008; Zhang
et al. 2013; Zhu et al. 2013; Cheng & Logan, 2011; Cheng et al. 2014; Zhuang &
Zhou, 2009; Liu et al. 2008; Zhuang et al. 2009, 2010b, 2012a, 2012b; Ieropoulos
et al. 2010a, Kim et al. 2012; Gurung et al. 2012; Zuo et al. 2007, 2008; Lefebvre
et al. 2009; Heijnea et al. 2011), but there are still many challenges that must be
overcome before that.
One of the challenges of scaling up MFCs is to maintain the power outputs at
levels needed for real applications, since previous studies concerning this issue
revealed that the power density decrease in the scale-up process (Ieropoulos et al.
2008; Dewan et al. 2008; Clauwaert et al. 2008; Dekker et al. 2009; Fan et al.
2010). As mentioned before, one possible way of scale-up MFCs is to increase its
volume. However, large MFCs units can alter electrode spacing which will affect the
power density through changes in the internal resistance and non-uniform current
distribution (Logan, 2010; Liu et al. 2008; Zhang et al. 2013; Zhu et al. 2013; Cheng
& Logan, 2011; Cheng et al. 2014). Questions such as whether the energy generated
by MFCs increases linearly or not with their size and how power is related to the
surface area of electrodes and electrode spacing should be therefore answered.
Dewan et al. (2008) studied the relation between the limiting electrode surface
area (anode compartment) and the cell power output and found that the maximum
power density is not directly proportional to the surface area of the anode, but
is proportional to the logarithm of the surface area. Cheng and Logan (2011)
studied the effect of the anode and cathode electrode surface area on the MFC
performance and found that doubling the anode and cathode size can increase the
power output to 12% and 62% respectively. Based on these findings, the three-
dimensional electrodes having the largest surface area for the same reactor volume
(compared to the other alternatives, such as carbon paper or carbon cloth), are
expected to be the best option for full-scale MFC (Cheng & Logan, 2011).
The electrode spacing is another important issue on MFC scale-up, since it
has significant influence on the cell internal resistance. Liu et al. (2008) studied
the effect of the electrode spacing on MFC performance. The electrode spacing
was found to be a key factor affecting the specific area, internal resistance and
power density. Also, in MFCs scale-up the large electrode array will affect the
biofilm growth on the electrode surface and will lead to a non-uniform internal
distribution in the cell. This will cause low energy generation, substrate utilization
and electrode utilization and less biomass production (Zhang et al. 2013; Zhu et al.
2013). Therefore, to maintain the power density during scale-up, the larger reactor
architecture must maintain or even reduce the electrode spacing by applying a
parallel electrode orientation (Liu et al. 2008; Zhang et al. 2013).
In the MFC scale-up process, as the reactor dimensions become larger, the
distance between the points where the electrons are generated and the leading-
out terminal where current flows out increases and consequently the ohmic losses
increase. Cheng et al. (2014) studied the impact of the anode ohmic resistance
on large-scale MFCs by changing the leading-out configurations and found
that a major part of the power loss from small to large scale was due to a bad
leading-out terminal. They found that an effective way to reduce the power loss
is changing the connecting configuration and the anode material since it reduces
the anode resistance, which is one of the limitations for high performances in
large-scale MFCs.
To avoid some of the problems of increasing the volume of a single MFC
unit, some effort has been puton the other commonly accepted configuration for
scale-up; the stack configuration. This consists of connecting single small units
together in series or parallel (Ieropoulos et al. 2008; Dekker et al. 2009; Zhuang
& Zhou, 2009; Liu et al. 2008; Zhuang et al. 2012a, 2012b; Ieropoulos et al.
2010a, Kim et al. 2012; Gurung et al. 2012). Ieropoulos et al. (2008) compared the
performance of three different sizes of microbial fuel cell (MFC) (large, medium
and small), when operated under continuous flow conditions using acetate as the
fuel substrate, by means of polarization curves. The authors, also, investigated the
connection of multiple small-scale MFCs, in series, parallel and series–parallel
configurations. Among the three combinations tested, the series–parallel proved to
be the more efficient one, stepping up both the voltage and current of the system.
The results from this study suggested that MFC scale-up may be better achieved
by connecting multiple small-sized units together rather than increasing the size of
an individual unit. When the single units are connected in stacks, voltage, current
or both can be increased depending on the stack size and configuration. This is
very useful since the power output needed can be met by adjusting the size of
the stack through the removal or addition of single units. It should be mentioned,
that when cells are connected in series the voltage of the MFC stack is the sum of
the individual cell unit voltages. When the stack presents a parallel configuration
the stack current is the sum of the individual currents (Ieropoulos et al. 2008,
2010a). However, connecting multiple units together in series and/or parallel
present some problems, such as voltage reversal, voltage losses and unpredictable
operation (Zhuang & Zhou, 2009; Liu et al. 2008). Moreover, the connection of
multiple MFCs may be complicated if the units are running under continuous
flow conditions, which involves electrical and hydraulic connections. Connecting
multiple MFCs in continuous flow requires the units to be joined to an inflow
and outflow stream (Ieropoulos et al. 2008). However, a study regarding this issue
shows that there is a voltage loss when the cells are hydraulically and electrically
connected (Zhuang & Zhou, 2009). To overcome that, different studies regarding
the stack configuration have been performed (Liu et al. 2008; Zhuang et al. 2012a,
2012b). Another problem of the stack configuration is the fact that the materials
used for a single operating cell may be not the best choice for a stack ofcells. In
their work, Ieropoulos et al. (2010a) found that the PEM that allows the highest
energy output level in lab-scale MFCs was not the best option for stacks.
One major challenge to MFCs became suitable for real applications, is developing
low-cost, simple constructions and easy to maintenance systems that can generate
high power outputs. The noble metals used as catalysts (like Platinum) are very
expensive and therefore unsuitable for large-scale applications. Using different
cathode catalyst materials or other electron acceptors instead of oxygen to increase
current densities is desirable (Zuo et al. 2007, 2008; Lefebvre et al. 2009; Heijnea
et al. 2011; Zhuang et al. 2009, 2010b).
11.5 Operational Conditions
Operating conditions like pH, organic load, feed rate, shear stress and temperature
are key parameters for MFC optimization. The comprehension of their
interdependence is also important.
11.5.1 Effect of pH
The pH is crucial to the MFCs power output. Anodic pH microenvironment
influences substrate metabolic activity and in turn affects the electron and proton
generation mechanism. Generally, bacteria require a pH close to neutral for their
optimal growth and respond to the changes of internal and external pH by adjusting
their activity. Depending on the bacteria and growth conditions, variations in pH
can cause modifications in several primary physiological parameters, such as, ion
concentration, membrane potential, proton-motive force and biofilm formation
(Zhang et al. 2011; Yaun et al. 2011). Most MFCs are operated at neutral pH in
order to optimize bacterial growth conditions (Biffinger et al. 2008; Jadhav &
Ghangrekar, 2009; Behera and Ghangrekar, 2009; He et al. 2008; Puig et al.
2010, Li et al. 2013b; Guerrini et al. 2013; Vologni et al. 2013). However, the
low concentration of protons at this pH makes the internal resistance of the cell
relatively high compared to chemical fuel cells that use acidic electrolytes.
The anode reactions produce protons and the cathode reaction consumes
protons. Accumulation of protons due to slow and incomplete proton diffusion
and migration through the membrane will cause a pH decrease in the anode. This
will lead to a decrease in the bacterial activity and the electron transfer in the
anode compartment. On the other hand, the continuous proton consumption by
the oxygen reduction reaction results in a pH increase in the cathode compartment
which, according to the Nernst equation, results in a decrease in current generation
(the oxygen reduction reaction rate decreases with an increase of pH) (Biffinger
et al. 2008; Jadhav & Ghangrekar, 2009; Behera & Ghangrekar, 2009; Erable et al.
2009; Zhuang et al. 2010a). The difference between the pH values in the anode
and cathode compartments causes a pH gradient, and consequently, a decrease in
the voltage efficiency and power generation especially at high current densities.
Therefore, neutral pH at the anode side and lower pH values at the cathode side
are desired. This can be achieved in a traditional dual-chamber MFC since two
different pH conditions can be maintained to optimize, respectively, the anodic
and cathodic reactions (Zhang et al. 2011; Yaun et al. 2011; Erable et al. 2009;
Zhuang et al. 2010a; Nimje et al. 2011). However, this is impossible in the case
of air-cathode MFCs, because only one electrolyte is present and the pH of this
electrolyte affects the reactions in both compartments (He et al. 2008; Raghavulu
et al. 2009). Since the air-cathode MFC configuration is more advantageous due to
higher power outputs and simplified reactor configuration, the research is focused
on maintaining a favorable pH value for both anode and cathode. According to He
et al. (2008), the air-cathode MFC can tolerate an electrolyte pH as high as 10 with
the optimal values ranging between pH 8 and 10. The anodic bacterial activity is
optimal at a neutral pH, while the cathodic reaction was improved at a higher pH.
Moreover, it was shown that the polarization resistance of the cathode was the
dominant factor limiting power output. Contrarily, in a continuous flow air-cathode
MFC, the higher power output was observed at acidophilic conditions (pH 6.3)
(Martin et al. 2010). Raghavulu et al. (2009), also, observed higher current densities
at acidophilic conditions (pH 6) when compared to neutral (pH 7) and alkaline (pH
8). The results showed that a better substrate degradation was achieved at neutral
conditions. The second best was for alkaline conditions and last one for acidophilic
conditions. Due to the poor cathodic oxygen reduction and the negative buildup of
a pH gradient between anode and cathode, Cheng et al. (2010) proposed a solution
to overcome these limitations. By inverting the polarity of the MFC successively
in the same half-cell they could neutralize the pH effects. They found that a mixed
culture forming an acidophilic biofilm can also catalyze the cathodic reaction of
oxygen in a single bioelectrochemical system. However, further research is needed
to explore the application of such bidirectional microbial catalytic properties for
sustainable MFC processes.
be taken when using these highly saline solutions at the anode side, since the
anodic microbial communities can be affected (Lefebvre et al. 2012).
11.5.2 Effect of temperature
In conventional fuel cells, most of the experimental studies are performed at
high temperatures to favor the electrochemical kinetics in the anode and cathode,
increase the open-circuit voltage, reduce the activation over voltage (according to
the Arrhenius relation) and, thus, increase the performance. Moreover, an increase
in temperature leads to an increase in solution conductivity with a consequent
decrease in the ohmic resistance. However, the high cell temperature decreases
the membrane stability and the oxygen partial pressure. Both positive and negative
effects of temperature are similar to all types of fuel cells. In the case of biological
fuel cells, such as MFCs, the high temperature levels of the conventional fuel
cells are detrimental to the microorganism. In MFCs, temperature is one of the
most important parameters that affects directly the growth and the metabolism of
microbial populations, affecting not only the intracellular biochemical processes
but also, the extracellular chemical and biochemical ones (Jadhav et al. 2009;
Martin et al. 2010; Min et al. 2008; Guerrero et al. 2010; Campo et al. 2013;
Michie et al. 2011; Liu et al. 2012; Li et al. 2013a; Wei et al. 2013). An increase in
temperature, leads to an increase of the intracellular biochemical reaction rate, an
increase on the microorganisms metabolism rate, which results in an increase of
the microorganisms growth rate and an increase inpower outputs. But in the long
run, when the microorganisms are under high temperatures, the other important
compounds of the cell such as proteins, nucleic acids or other temperature sensitive
parameters, may suffer an irreversible damage, which will lead to a decline in cell
function. Therefore, the microorganisms growth rate and concentration decrease
and consequently the power output decreases. The growth and reproduction rates
of microorganism will be fastest and the cell function will not be affected only if
they live in the optimum temperature (Min et al. 2008; Wei et al. 2013).
Generally, different types of microorganism grow optimally at different
temperature ranges. The microorganisms used in MFCs are categorized in
mesophilic (32–42°C) and thermophilic (48–55°C), while at the transitory range
of 40–45°C, both mesophilic and thermophilic microorganisms function under
suboptimal conditions. The methanogens are more sensitive to rapid changes of
temperature. To some extent, this is an exciting field for research since the results
would have potential application in areas with large temperature variations. In the
case of MFCs, it was found that the more advantageous operating temperature
range is between 30°C and 40°C, leading to higher power outputs and microbial
activity (Jadhav et al. 2009; Martin et al. 2010; Min et al. 2008; Guerrero et al.
2010; Campo et al. 2013; Michie et al. 2011; Liu et al. 2012; Li et al. 2013a; Wei
et al. 2013). However, the startup time for a MFC operating at higher temperatures
(above 40°C) is longer than for low temperatures (below 40°C), despite the fact
11.5.3 Organic load
In the anodic chamber of a MFC, microorganisms produce electrons and protons
due to the oxidation of organic compounds. Therefore, the organic load presented
in the effluent will affect both the performance and the microbial community of a
MFC. Some studies have been conducted in order to study the effect of the organic
load on the MFC performance through the parameter, organic loading rate (OLR),
which is related to the capacity of conversion the organic substrates per reactor
volume (Jadhav et al. 2009; Behera et al. 2009; Martin et al. 2010; Guerrero et al.
2010; Campo et al. 2013; Aelterman et al. 2008; Mohan et al. 2009; Lorenzo et al.
2010; Reddy et al. 2010; Juang et al. 2011; Velvizhi et al. 2012; Ozkaya et al. 2013;
Jia et al. 2013). These studies allowed to conclude that there is an optimum range of
values for OLR which has a direct influence on the power output, COD (chemical
oxygen demand) removal and coloumbic efficiency (CE). Generally, increasing
the OLR, leads to an increase in the cell voltage, power output, as well as, the
COD removal rate. This is due to the fact that under higher organic loads, more
substrate is available to sustain the metabolic activity and more organic matter is
used for power generation. Moreover, increasing the OLR, the ionic strength of the
anodic solution is increased and the activity and concentration of the anodophilic
microorganisms are also favored (Martin et al. 2010; Velvizhi et al. 2012). However,
higher organic loading rate lead to a decrease on power generation, even though
the increase on the substrate degradation occurs (Jadhav et al. 2009; Martin et al.
2010; Campo et al. 2013; Aelterman et al. 2008; Mohan et al. 2009; Lorenzo et al.
2010; Reddy et al. 2010; Juang et al. 2011; Velvizhi et al. 2012; Jia et al. 2013).
This can be explained by the fact that a further increase on the organic load leads
to an excessive nutrient supply. This additional supply will be metabolized by the
non-electricity generation microorganisms and/or will hindered the performance
of the biocatalyst, generating less current. The increase in the substrate removal
rate observed at high loading rates may be attributed to the direct anodic oxidation
(DAO) mechanism. This helps the further oxidation of the substrate leading to
enhanced removal rates (Mohan et al. 2009). Despite the fact that an increase on
the organic load leads to an increase on the power output and COD removal, the
Coulombic efficiency decreases (Jadhav et al. 2009; Martin et al. 2010; Campo
et al. 2013; Aelterman et al. 2008; Mohan et al. 2009; Lorenzo et al. 2010; Reddy
et al. 2010; Juanget al. 2011; Velvizhi et al. 2012; Ozkaya et al. 2013). This means
that most of the electrons produced from the oxidation of organic compounds are
diverted to non-electricity generating processes, such as methane production. High
organic loading rates cause saturated conditions which will lead to competition
between microorganisms involved in electricity production and other processes
(such as methane production), leading to a greater organic matter removal that may
not be directly related to current generation. Therefore, the Coulombic efficiency
decreases, despite the fact that the substrate degradation increases.
Studies regarding the effect of the OLR on methane production, show that the
volumetric methane production rate increases with an increase of the organic
loading rate (Martin et al. 2010). Therefore to avoid that, low organic loading rates
should be used (Martin et al. 2010; Lorenzo et al. 2010; Velvizhi et al. 2012),
but working in such conditions will lead to lower power outputs, higher internal
resistances and lower COD removal rates. So, it is necessary to find a way to
simultaneously increase the power and Coulombic efficiency and decrease the
methane production. Studies regarding that suggested that an optimized OLR can
be used to avoid in some extent the methane production and to shift the electricity-
to-methane production rate towards the electricity production (Martin et al. 2010;
Lorenzo et al. 2010; Velvizhi et al. 2012).
11.6 Modelling Studies
Fuel cell modelling has received much attention over the last decade in an attempt
to understand the phenomena occurring within the cell. Different modeling
approaches have been developed and have led to analytical, semi-empirical and
mechanistic models (Oliveira et al. 2007). In Figure 11.3 the models developed
for MFCs are categorized according to the features studied, such as dimension,
mass transport, pH effects, multispecies, biofilm, biochemical, electrochemical
and dynamic (Zeng et al. 2010; Picioreanu et al. 2007, 2008, 2010a, 2010b;
Pinto et al. 2010, 2012; Oliveira et al. 2013, Zhang et al. 2014; Wen et al. 2009;
Marcus et al. 2007; Merkey & Chopp, 2012).
11.7 Economic Evaluation
The world demand for adequate sanitation and access to portable water leads to
the need to treat the wastewaters. As example in Unites States (US) 4% to 5% of
the energy produced is used in water infrastructures, such as water collection,
treatment and distribution. The costs of maintaining such infrastructures are
significant and is expected that will increase over the next years to maintain and
improve this infrastructures (Logan, 2008). Wastewaters contain energy, in the
form of organic matters, that we expend energy to remove rather than trying to
recover it. At a conventional wastewater treatment plant it was estimated that there
was 9.3 times as much energy in the wastewater than used to treat it (Logan, 2008).
Industrial, agricultural and domestic wastewaters are estimated to contain a total
of 17 Gigawatts, which is, for example, the same amount of energy that is used
for the whole water structure in US. Therefore, if this energy can be recovered a
water treatment plant can be self-sufficient. The most common technology that
is used to extract this energy on a commercial scale is the anaerobic digestion.
However, anaerobic digestion requires meso-to thermophilic temperatures to
achieve sufficient turnover and limited methane solubility. The major operating
costs for wastewater treatment are water aeration, sludge treatment and wastewater
pumping. The aerobic treatment consumes large amounts of electrical energy
for aeration accounting for half of the operation costs at a typical wastewater
treatment plant.Also, the sludge treatment cost to wastewater treatment plants can
reach up to 500€/ton dry matter. Therefore, eliminating or reducing this two costs
can save appreciable amount of energy and can decrease the economic balance
of the process. Based on that, the MFCs appear as an interesting technology for
the production of energy from wastewaters. MFCs have many advantages over
the conventional technologies used for generating energy from organic matter,
such as: direct conversion of organic matter into electricity, which allows higher
conversion efficiencies; can operate at ambient or low temperatures reducing
the heat costs; do not require gas treatment since the gas produced in MFCs
is carbon dioxide which as no useful energy content; do not need energy input
for aeration since the cathode can be passively aerated; can be used in remote
locations (locations lacking electrical infrastructures); can use a diversity of fuels
allowing to satisfy the energy requirements for each application; and can reduce
considerably the solids production at a wastewater treatment plant, reducing the
operating costs for solids handling. However, the MFC technology has to compete
with the well establish anaerobic digestion technology since both can use the same
biomass to produce energy and have advantages and disadvantages. To implement
this kind of system, the cost which is one of the MFC challenges, needs to be
faced and exceeded. In order to become an advantageous technology, the cost
benefits of the MFC system (energy production and wastewater treatment) should
be higher than the total costs (implementation an operational cost). To achieve
this, the benefits should be maximized and the costs minimized. In general, high
conversion rates, whose evaluation is based on the power output, are a major
condition for low costs. However, this rate is limited by the energy losses in the
systems, already explained in the fundamentals and challenges section. Therefore,
it is extremely important to reduce these losses and as presented above different
studies have been done in order to achieve this. Moreover, it is fundamental to
identify the major costs related to the MFCs technology and find ways to reduce
them (Sleutels et al. 2012; Fornero et al. 2010b; Rozendal et al. 2008; Pham et al.
2006; Rabaey & Verstraete, 2005).
In the work developed by Rozendal et al. (2008) an overview of the estimated
implementation costs of a MFC based on the materials available commercially
and commonly used on these systems, as well as, a comparison of estimated costs
among the different wastewater treatments available is presented. They compared
the costs of MFCs technology with the anaerobic digestion (AD) treatment. The
authors concluded that with the materials available, the MFC capital costs are
higher than those of conventional treatments, 8€/kg COD for MFC and 0.01€/
kg COD for AD. Among the different costs associated to the MFC technology,
they showed that the major percentage for the total cost is the cathode side
material (47%) due to the need of using platinum as cathode electrode to promote
the oxygen reduction reaction, with an estimate cost of 500€/m2. The second
major contribution for the overall cost (38%) is related to the membrane used to
separate both compartments with an estimated cost of 400€/m2. Therefore, MFC
systems can only became economically interesting if these larger capital costs are
reduced, based on changing the materials commonly used for less expensive ones.
The authors estimated a total costs of 0.4€/kg COD, which was acceptable and
near the costs of the AD. Also, it should be emphasized that 1 kg of COD can
be converted to 1kWh using anaerobic digestion while the same amount of COD
can be converted in 4 kWh when a MFC is used (Pham et al. 2006; Rabaey &
Verstraete, 2005). This shows a higher energy conversion rate when MFC are used
instead of the AD. Therefore, the slightly higher costs of the MFC technology may
be compensated by its higher conversion rate. The MFCs have, also, other benefits
over the AD such as operation on a small-scale, with low COD concentrations
and temperatures. Moreover, economic value can be added to this technology,
if besides the COD removal, the system is used for the production of valuable
chemical products instead of electricity. The production of hydrogen gas, caustic
or hydrogen peroxide that can be directly used in the treatment process, or the
removal of persistent and toxic wastewater compounds may provide an additional
cost benefit for the wastewater treatment plants when a MFC is used.
11.8 Summary
The fundamentals and challenges on MFCs have been summarized and the recent
modeling and experimental studies have been reviewed. An economic outlook of
this technology was also presented. As was mentioned for practical applications of
this technology cost effectiveness is essential. Therefore to a MFC be cost effective
the value of the products and of wastewater treatment (revenues) need to be higher
than the capital and operational costs. The production of the valuable chemicals
with MFCs as well as the reduction of the material costs are expected to offset the
higher capital investments of this technology. Apart from the economic aspects,
the MFC are a sustainable technology that can be adaptable to a wide variety of
application. They produce electrical energy from different array of electron donors
with different concentrations at low and moderate temperatures and no other
existing technology can achieve that.
Many studies have been focused on analyzing and improving single parts
of MFCs and different materials, designs, microbes, mediators and operating
conditions have been suggested, which manage to increase the power outputs.
However, power generation have not yet reached the levels needed for commercial
use. Also, the best operating and design conditions for one type of MFC may
be not necessary the same to another type. Therefore, when developing a MFC,
the materials and conditions need to be carefully considered and chosen in order
to achieve the best performance which means the highest power output. Since
significant challenges still exist before a MFC can be ready to commercialization,
further studies on more cost-efficient materials and optimization of configurations
are needed in order incorporate large-scale MFCs in conventional wastewater
treatment systems.
The development of mathematical models is essential to a better understanding
and prediction of the main processes occurring in a MFC. These models can be
used to predict the effect of various operating and structural parameters on cell
performance and are a fundamental tool for the design and optimization of fuel
cell systems. In spite of the modelling work on MFCs developed in the past few
years, new models are still needed to describe the main biological, mass transfer,
energy and charge effects in the anode and cathode compartment of a MFC, as
well as the biofilm formation.
The developments both on experimental and modeling issues have led to an
increase on the MFC power output at lab-scale. Therefore, the scaling-up and
durability need to emerge as future development areas in the MFC research.
11.9 Acknowledgements
V. B. Oliveira acknowledges the post-doctoral fellowship (SFRH/BDP/91993/2012)
supported by the Portuguese “Fundação para a Ciência e Tecnologia” (FCT),
POPH/QREN and European Social Fund (ESF). POCI (FEDER) also supported
this work via CEFT.
11.10 References
Aaron D., Tsouris C., Hamilton C. Y. and Borole A. P. (2010). Energies, 3, 592–606.
Aelterman P., Versichele M., Marzorati M., Boon N. and Verstraete W. (2008). Bioresource
Technology, 99, 8895–8902.
Ahn Y. and Logan B. E. (2013). Bioresource Technology, 132, 436–439.
Behera M. and Ghangrekar M. M. (2009). Bioresource Technology, 100, 5114–5121.
Biffinger J. C., Pietron J., Bretschger O., Nadeau L. J., Johnson G. R., Williams C. C., Nealson
K. H. and Ringeisen B. R. (2008). Biosensors and Bioelectronics, 24, 900–905.
Campo A. G., Lobato J., Cañizares P., Rodrigo M. A. and Morales F. J. F. (2013). Applied
Energy, 101, 213–217.
Cheng K. Y., Ho G. and Ruwisch R. C. (2010). Environmental Science & Technology, 44,
518–525.
Cheng S. and Logan B. E. (2011). Bioresource Technology, 102, 4468–4473.
Cheng S., Ye Y., Ding W. and Pan B. (2014). Journal of Power Sources, 248, 931–938.
Clauwaert P., Aelterman P., Pham T. H., Schamphelaire L. De, Carballa M., Rabaey K. and
Verstraete W. (2008). Applied Microbiology and Biotechnology, 79, 901–913.
Dekker A., Heijne A. T., Saakes M., Hamelers H. V. M. and Buisman C. J. N. (2009).
Environmental Science & Technology, 43, 9038–9042.
Dewan A., Beyenal H. and Lewandowski Z. (2008). Environmental Science & Technology,
42, 7643–7648.
Du Z., Li H. and Gu T. (2007). Biotechnology Advances, 25, 464–82.
Erable B., Etcheverry L. and Bergel A. (2009). Electrochemistry Communications, 11,
619–622.
Fan Y., Han S. K. and Liu H. (2012). Energy & Environmental Science, 5, 8273–8280.
Fan Y., Hu H. and Liu H. (2007a). Journal of Power Sources, 171, 348–354.
Fan Y., Hu H. Q. and Liu H. (2007b). Environmental Science & Technology, 41, 8154–8158.
Fornero J. J., Rosenbaum M., Cotta M. A. and Angenet L. T. (2010a). Environmental
Science & Technology, 44, 2728–2734.
Fornero J. J., Rosenbaum M. and Angenet L. T. (2010b). Electroanalysis, 22, 832–843.
Guerrero A. L., Scott K., Head I. M., Mateo F., Ginesta A. and Godinez C. (2010). Fuel,
89, 3985–3994.
Guerrini E., Cristiani P. and Trasatti S. P. M. (2013). International Journal of Hydrogen
Energy, 38, 345–353.
Gurung A. and Oh S. E. (2012). Energy Sources, Part A 34, 1591–1598.
He Z., Huang Y., Manohar A. K. and Mansfeld F. (2008). Bioelectrochemistry, 74, 78–82.
Heijnea A. T., Liu F., van Rijnsoever L. S., Saakes M., Hamelers H. V. M. and Buisman C.
J. N. (2011). Journal of Power Sources, 196, 7572–7577.
Holmes D. E., Bond D. R., O’Neil R. A., Reimers C. E., Tender L. R. and Lovley D. (2004).
Microbial Ecology, 48, 178–190.
Ieropoulos I., Greenman J. and Melhuish C. (2008). International Journal of Energy
Research, 32, 1228–1240.
Ieropoulos I., Greenman J. and Melhuish C. (2010a). Bioelectrochemistry, 78, 44–50.
Ieropoulos I., Winfield J. and Greenman J. (2010b). Bioresource Technology, 101,
3520–3525.
Ieropoulos I. A., Greenman J., Melhuish C. and Hart J. (2005). Enzyme and Microbial
Technology, 37, 238–245.
Jadhav G. S. and Ghangrekar M. M. (2009). Bioresource Technology, 100, 717–723.
Jia J., Tang Y., Liu B., Wu D., Ren N. and Xing D. (2013). Bioresource Technology, 144,
94–99.
Juang D. F., Yang P. C., Chou H. Y. and Chiu L. J. (2011). Biotechnology Letters, 33,
2147–2160.
Juang D. F., Yang P. C. and Kuo T. H. (2012). International Journal of Environmental
Science and Technology, 9, 267–280.
Jung S. and Regan J. M. (2007). Applied Microbiology and Biotechnology, 77, 393–402.
Kim D., An J., Kim B., Jang J. K., Kim B. H. and Chang I. S. (2012). ChemSusChem, 5,
1086–1091.
Lefebvre O., Ooi W. K., Tang Z., Abdullah-Al-Mamun Md., Chua D. H. C. and Ng H. Y.
(2009). Bioresource Technology, 100, 4907–4910.
Lefebvre O., Tan Z., Kharkwal S. and Ng H. Y. (2012). Bioresource Technology, 112,
336–340.
Li L. H., Sun Y. M., Yuan Z. H., Kong X. Y. and Li Y. (2013a). Environmental Technology,
34, 1929–1934.
Li X. M., Cheng K. Y. and Wong J. W. C. (2013b). Bioresource Technology, 149, 452–458.
Liu H., Cheng S., Huang L. and Logan B. E. (2008). Journal of Power Sources, 179,
274–279.
Liu L., Tsyganova O., Lee D. J., Suc A., Chang J. S., Wang A. and Ren N. (2012).
International Journal of Hydrogen Energy, 37, 15792–15800.
Liu Z., Liu J., Zhang S. and Su Z. (2008). Biotechnology Letters, 30, 1017–1023.
Logan B. E. (2008). Microbial Fuel Cells. Wiley Bicentennial, United States of America.
Logan B. E. (2010). Applied Microbiology and Biotechnology, 85, 1665–1671.
Lorenzo M. D., Scott K., Curtis T. P. and Head I. M. (2010). Chemical Engineering Journal,
156, 40–48.
Lovley D. R. (2008). Geobiology, 6, 225–231.
Lovley D. R. and Nevin K. P. (2008). Electricity production with electricigens. In: Bioenergy,
J. D. Wall, C. S. Harwood and A. Demain (eds), Chap. 23, ASM Press, Washington, DC,
pp. 295–306.
Marcus A. K., Torres C. I. and Rittman B. E. (2007). Biotechnology and Bioengineering,
98, 1171–1182.
Martin E., Savadogo O., Guiot S. R. and Tartakovsky B. (2010). Biochemical Engineering
Journal, 51, 132–139.
Merkey B. V. and Chopp D. L. (2012). Bulletin of Mathematical Biology, 74, 834–857.
Michie I. S., Kim J. R., Dinsdale R. M., Guwy A. J. and Premier G. C. (2011). Environmental
Science & Technology, 4, 1011–1019.
Min B., Román O. B. and Angelidaki I. (2008). Biotechnology Letters, 30, 1213–1218.
Mohan S. V., Raghavulu S. V., Peri D. and Sarma P. N. (2009). Biosensors and Bioelectronics,
24, 2021–2027.
Nam J. Y., Kim H. W., Lim K. H., Shin H. S. and Logan B. E. (2010). Biosensors and
Bioelectronics, 25, 1155–1159.
Nimje V. R., Chen C. Y., Chen C. C., Tsai J. Y., Chen H. R., Huang Y. M., Jean J. S.,
Chang Y. F. and Shih R. C. (2011). International Journal of Hydrogen Energy, 36,
11093–11101.
Oliveira V. B., Falcão D. S., Rangel C. M. and Pinto A. M. F. R. (2007). International
Journal of Hydrogen Energy, 32, 415–424.
Oliveira V. B., Simões M., Melo L. F. and Pinto A. M. F. R. (2013). Energy, 61, 463–471.
Ozkaya B., Cetinkaya A. Y., Cakmakci M., Karadag D. and Sahinkaya E. (2010). Bioprocess
and Biosystems Engineering, 36, 399–405.
Pham H. T., Rabaey K., Aelterman P., Clauwaert P., Schamphelaire L. D., Boon N. and
Verstraete W. (2006). Engineering in Life Sciences, 6, 285–292.
Pham H. T., Boon N., Aelterman P., Clauwaert P., Schamphelaire L. D., Oostveldt P., Verbeken
K., Rabaey K. and Verstraete W. (2008). Microbial Biotechnology, 6, 487–496.
Picioreanu C., Head I. M., Katuri K. P., van Loosdrecht M. C. M. and Scott K. (2007).
Water Research, 41, 2921–2940.
Picioreanu C., Head I. M., Katuri K. P., van Loosdrecht M. C. M. and Scott K. (2008).
Water Science & Technology, 57, 965–971.
Picioreanu C., Head I. M., Katuri K. P., van Loosdrecht M. C. M. and Scott K. (2010a).
Journal of Applied Electrochemistry, 40, 151–162.
Picioreanu C., van Loosdrecht M. C. M., Curtis T. P. and Scott K. (2010b).
Bioelectrochemistry, 78, 8–24.
Pinto R. P., Srinivasan B., Manuel M. F. and Tartakovsky B. (2010). Bioresource Technology,
101, 5256–5265.
Pinto R. P., Tartakovsky B. and Srinivasan B. (2012). Journal of Process Control, 22,
1079–1086.
Potter M. C. (1911). Proceedings of the Royal Society of London. Series B, 84(571), 260–276.
Puig S., Serra M., Coma M., Cabré M., Balaguer M. D. and Colprim J. (2010). Bioresource
Technology, 101, 9594–9599.
Qiang L., Yuan L. J. and Ding Q. (2011). HuanjingKexue/Environmental Science, 32,
1524–1528.
Rabaey K. and Verstraete W. (2005). Trends in Biotechnology, 23, 290–298.
Raghavulu S. V., Mohan S. V., Reddy M. V., Mohanakrishna G. and Sarma P. N. (2009).
International Journal of Hydrogen Energy, 34, 7547–7554.
Reddy M. V., Srikanth S., Mohan S. V. and Sarma P. N. (2010). Bioelectrochemistry, 77,
125–132.
Rittmann B. E. (2008). Biotechnology and Bioengineering, 100, 203–212.
Rochex A., Godon J.-J., Bernet N. and Escudie R. (2008). Water Research, 42, 4915–4922.
Rozendal R. A., Mahelers H. V. M., Rabaey K., Keller J. and Buisman C. J. N. (2008).
Trends in Biotechnology, 26, 450–459.
Sleutels T. H. J. A., Heijne A. T., Buisman C. J. N. and Hamelers H. V. M. (2012).
ChemSusChem, 5, 1012–1019.
Tender L. M., Gray S. A., Groveman E., Lowy D. A., Kauffman P., Melhado J., Tyce R. C.,
Flynn D., Petrecca R. and Dobarro J. (2008). Journal of Power Sources, 179, 571–575.
Torres C. I., Lee H. S. and Rittmann B. E. (2008). Environmental Science & Technology,
42, 8773–8777.
Velvizhi G. and Mohan S. V. (2012). International Journal of Hydrogen Energy, 37,
5969–5978.
Vologni V., Kakarla R., Angelidaki I. and Min B. (2013). Bioprocess and Biosystems
Engineering, 36, 635–642.
Wei L., Han H. and Shen J. (2013). International Journal of Hydrogen Energy, 38,
11110–11116.
Wen Q., Wu Y., Cao D., Zhao L. and Sun Q. (2009). Bioresource Technology, 100,
4171–4175.
Yong L., Renduo Z., Guangli L., Jie L., Bangyu Q., Mingchen L. and Shanshan C. (2011).
Bioresource Technology, 102, 3827–3832.
You S. J., Zhao Q. L., Jiang J. Q. and Zhang J. N. (2006). Chemical and Biochemical
Engineering Quarterly, 20, 407–412.
Yuan Y., Zhao B., Zhou S., Zhong S. and Zhuang L. (2011). Bioresource Technology, 102,
6887–6891.
Zeng Y., Choo Y. F., Kim B. H. and Wu P. (2010). Journal of Power Sources, 195, 79–89.
Zhang L. and Deshusses M. (2014). Environmental Technology, 35, 1064–1076.
Zhang L., Li C., Ding L., Xu K. and Ren H. (2011). Journal of Chemical Technology &
Biotechnology, 86, 1226–1232.
Zhang L., Li J., Zhu X., Ye D. and Liao Q. (2013). Chemical Engineering Journal, 223,
623–631.
Zhongjian L., Xingwang Z. and Lecheng L. (2008). Process Biochemistry, 43, 1352–1358.
Zhu X., Zhang L., Li J., Liao Q. and Ye D. (2013). International Journal of Hydrogen
Energy, 38, 15716–15722.
Zhuang L., Zhoua S., Wang Y., Liu C. and Geng S. (2009). Biosensors and Bioelectronics,
24, 3652–3656.
Zhuang L., Zhou S., Li Y. and Yuan Y. (2010a). Bioresource Technology, 101, 3514–3519.
Zhuang L., Feng C., Zhou S., Li Y. and Wang Y. (2010b). Process Biochemistry, 45,
929–934.
Zhuang L., Yuan Y., Wang Y. and Zhou S. (2012a). Bioresource Technology, 123, 406–412.
Zhuang L., Zheng Y., Zhou S., Yuan Y., Yuan H. and Chen Y. (2012b). Bioresource
Technology, 106, 82–88.
Zhuang L. and Zhou S. (2009). Electrochemistry Communications, 11, 937–940.
Zuo Y., Cheng S., Call D. and Logan B. (2007). Environmental Science & Technology, 41,
3347–3353.
Zuo Y., Cheng S. and Logan B. (2008). Environmental Science & Technology, 42,
6967–6972.
12.1 Introduction
Energy saving is a widespread goal in any country but dealing with problems of
consumption rationalization is complicated. In fact even though it is clear that in
each energy sector, saving should be obtained by multilevel strategies, the general
trend is to pursue the goal without considering integrated approaches, sometimes
even adopting conflicting actions.
Integrated water cycle services play an important role in energy consumption –
about 6.3 billion kWh per year in Italy, in 2009–2011 (Terna, 2012) – and the
contribution of the WWTP (Waste Water Treatment Plant) – about 3 billion kWh
per year in Italy, in the same period – can be really significant.
Therefore as the energy consumption for wastewater treatment plants is a
significant fraction of total energy consumption – about 0.9% in Italy in 2009–
2011 (DPS, 2010; Co.N.Vi.RI, 2011) – and above all it represents a significant part
of the WWTP managements costs, applying energy saving strategies to wastewater
treatments is important; but it must be done by considering the overall factors and
avoiding negative impacts on depuration process.
Often this objective is only dealt with in terms of improvement of sludge
treatment, however other important aspects of the WWTPs – such as the lack of a
sufficient flexibility in the treatments, the oversizing and inflexibility of important
sections as the aeration, the pumping and mixing systems, as well as the adoption
of operative procedures not well adjusted on process real needs – can undoubtedly
In most reported cases, the efficacy of the treatment was the primary goal of
the interventions and once the conditions able to guarantee the process reliability
were identified and set out, the energy saving goals were pursued. In other cases
instead the goal was the efficiency and the energy savings but when no the whole
wastewater treatment has been assessed, the effectiveness of some steps was
sometimes compromised. In Table 12.2 the typical energy consumptions of the
plants before the interventions are reported.
all the utilities in terms of energy consumption and operation hours. Each section
was analysed by means of field surveys and spreadsheets, specifically designed to
compare actual and designed conditions and as a result the actions were conceived
only for pump stations, aeration systems and mechanical devices.
In all cases the interventions were developed on 3 subsequent levels: a first
level of immediate correction through management and maintenance procedures;
a second level with the implementation and minimal structural upgrades;
and finally a third long-term level, not dealt with in this chapter, concerning
the up-grade of entire treatment sections and/or the restoring of sewage systems.
In Table 12.3 the main critical situations founded in all the WWTPs are listed.
Figure 12.1 Incoming wastewater load and oxygen concentration in oxidation tank
of WWTP4 with different blower combinations and power consumption.
Another crucial point was to improve the efficiency of the air distribution network
and diffusers, introducing some periodical maintenance procedures such as: the
removing of the material interfering with the diffusers on the bottom of the aeration
tanks; the cleaning of the air supply networks by means of an online washing system;
the replacing of the damaged or obstructed membranes of the diffusers.
All the different interventions – enabling lower working pressures of membranes
and networks – led to a general improvement in the oxygen transfer efficiency.
Lastly the aeration devices were analyzed in terms of the installed powers and
energy yields in relation to the requirements, to identify those systems no longer
able to guarantee the required performance with suitable power consumption and
renew them. In this way, for example at the WWTP1, two lobe blowers were replaced
by a single turbo-blower equipped with inverter, this resulted in 200,000 kWh/
year estimated savings; a similar intervention has been scheduled for 2014 at the
WWTP4 – a turbo-blower of 75 kW in replacement of a 90 kW lobe blower- with
estimated savings of 66.000 kWh/year. In some other plants moreover old air
systems were replaced by fine bubbles diffusing devices.
A good choice, as it was for the WWTP1, could be to install pressure control
devices in the air networks, with the purpose of monitoring the diffusers fouling
trend and correct it at the right time.
Table 12.6 refers to some examples of the interventions and the relative energy
savings obtained.
e.i.: equivalent inhabitants; kWh/e.i.f. and kWh/e.i.b.: calculated on the basis of flows
and BOD5 values respectively; WWTP: wastewater treatment plant; (*): estimated values;
H.S.: high season (summer).
45 N H 4 o ut N - N O 3 o ut m o dif ic a t io n
40
ne w a e ra t io n
35 s ys t e m
30
mg/L
25
20
15
10
0
10/01/12
10/04/12
10/07/12
10/10/12
11/01/12
11/04/12
11/07/12
11/10/12
12/01/12
12/04/12
12/07/12
12/10/12
Figure 12.2 Concentration of NH4+ and NO3-N in the effluent of WWTP5 in
2010–2012 (Ragazzo et al. 2013).
30
25 N-NO2+N-NO3
Concentration
20
(mg/L)
15 Line DON
Line O Before
10
Line O After
5
0
0 15 30 45 60 90 120 150 180 210
Time (min)
Figure 12.3 Respirometric tests (NUR) of sludge of WWTP6 (Ragazzo et al. 2013).
problems in municipal wastewater treatment. They are certainly a good choice for
WWTPs having space problems and reuse purposes, but extending their application
anywhere is not a good solution: this in fact could entail energy consumption up
to 10 times higher as compared to conventional systems. This is evident from the
management experiences of advanced MBR technologies documented by Brepols
and Schäfer in (2010) where, for a large scale WWTP (80,000 e.i.) characterized
by high flow variations, energy consumptions were from 1.7 up to 9.3 times higher
than those registered in conventional WWTPs (30,000–160,000 e.i.) working in
the same conditions.
An example of a wrong application of this technology is represented by the case
where a cross flow tubular membrane system, suitable to treat homogeneous and
constant concentrated wastewaters, was adopted at a municipal wastewater plant –
WWTP7 3000 e.i.– receiving sewage from a combined system. In this plant, the
energy consumptions ranged between 4 and 6.4 kWh/m3 and, even though many
efforts were done to reduce the wastings – the optimization of washing procedures,
the addition of a pre-filtration stage and the redesigning of the pumping system –
consumptions are still from 3 up to 15 times higher than those of similar
conventional activated sludge WWTP.
Another system clearly in contrast with energy saving goals is the UV
disinfection technology, often proposed as the most credible alternative to chemical
disinfection in wastewater treatment.
However considering energy consumptions issues only, at doses expected to
be effective (40–80 mJ/cm2), the only UV radiation entails an additional impact
between 0.05 and 0.15 kWh/m³ and this could be even higher according to some
literature data stating the actual UV doses reducing pathogens are higher.
In any case, in contrast with the reasons supporting the adoption of such
technology, the chemical dosage is always required either to pursue the bacterial
reductions for reuse purposes and/or to support the UV inefficiency to reduce
pathogens or to maintain the cleaning of the mandatory filtration steps.
Moreover considering that available chemical alternatives are supposed to
be safe for the environment (also chlorine could be used without a significant
by-products production if dosed at a specific ratio with ammonium nitrogen),
adopting UV technology that implies such high energy consumptions, often does
not make sense.
12.4 Conclusions
The energy saving in relation to wastewater treatment is an important goal, but
stable results can be achieved only using integrated approaches and acting at
different levels: on the design criteria; on technology choice; on plant monitoring
and management practices.
Even though the energy saving should start from restoring the sewage systems
itself, the fact that the WWTPs are rigidly conceived on the basis of the only
designing reference data and are provided with inflexible structures and machines,
implies significant energy wasting. Therefore a fundamental step to pursue energy
saving goals should be the cooperation among the different subjects – plant
managers, designers and so on – to meet the actual treatment needs, realizing the
most appropriate and flexible solutions.
This work, reporting the energy savings achieved in northern Italy through some
general interventions on the management and controls as well as through minimal
up-grades on machines and structures, shows how the multilevel approach could
lead to significant changes in energy consumption (Table 12.7).
Table 12.7 Energy savings for each WWTP, deriving from all interventions
2010–2013.
12.5 References
Brepols C. and Schäfer H. (2010). Operational Cost. In: Operating Large Scale Membrane
Bioreactors for Municipal Wastewater Treatment, C. Brepols (ed.), IWA Publishing,
Chapter 7, London, UK.
Co.N.Vi.R.I. (National Committee for surveillance of water resources) (2011). Rapporto
sullo stato dei servizi idrici (Report on the state of water services). Annual Report to
the Parliament, Roma.
DPS (Development Department-Ministry of Economic Development) (2010). Obiettivi di
servizio – Servizio idrico integrato- S.11 (Service objectives – Water Service – S11). http://
www.dps.tesoro.it/obiettivi_servizio/servizio_idrico.asp (accessed 13 December 2013).
The Netherlands
3 Regional Water Authority Aa en Maas, ‘s-Hertogenbosch, Noord
The Netherlands
6 Regional Water Authority de Dommel, Boxtel, Noord Brabant,
The Netherlands
7Regional Water Authority Vallei en Veluwe, Apeldoorn, Gelderland,
The Netherlands
Abstract
The energy factory is a sewage treatment plant (STP) that produces all the required
energy itself (Energy factory). The concept has given a strong boost to sustainable
innovation within the regional water authorities. This concept supports the philosophy
behind the ‘Wastewater management roadmap towards 2030’, which gives a new
perspective on wastewater as a source of raw materials, instead of a waste product.
13.1 Energy Factory
13.1.1 The concept
Wastewater contains potentially 8 times the energy needed to purify it (Energy
Factory). This means that an STP can generate energy, at least in theory. From
this aspect, an energy-producing STP can considered to be an energy factory.
This involves two major aspects: minimizing energy demand and maximizing
energy production.
There are two pathways to minimizing energy demand: (1) Use of energy-
efficient equipment (high-efficiency motors, plate aerators, etc.); (2) use of process
13.1.2 The history
In light of the increasing worldwide demand for raw materials and other issues,
the international Global Water Research Coalition (GWRC) has called for new
concepts and visions for the future of the STP. The Dutch contribution to this
discussion was the NEWater concept (NEWater). Nutrients, Energy and Water are
the major products addressed in this concept.
In early 2008, there was very little experience in the Netherlands on sludge
pretreatment for increasing biogas yield and/or faster sludge decomposition. WBL
(the regional water Board implementing body for Limburg) ascertained early
on that there were real energy gains to be obtained from secondary sludge, but
the available hydrolysis and other techniques were only economically viable for
very large installations. In 2008, together with Dutch firm Sustec, they began
investigating the potential of treating the sludge in a continuous process, ultimately
applying the process at pilot scale in 2009. The results were encouraging enough
for WBL to move to full scale application of thermal pressure hydrolysis (TPH)
at the STP in Venlo. The first TPH in the Netherlands was completed in 2012.
In the sequel, two more TPH pilots were constructed in 2010, one at the STP
in Amersfoort and another at the STP Hengelo. Both were successful, and they
ultimately led multiple regional water authorities to adopt the TPH technology.
The regional water authority Reest & Wieden applied a different technology, a
two-stage digestion process (thermophilic followed by mesophilic).
The development of the energy factory concept is the result of a unique approach.
Four regional water authorities (Aa & Maas, Hoogheemraadschap Hollands
Noorderkwartier, Veluwe and Rivierenland) developed the energy factory concept
as a response to a competitive project called by the Dutch Water Authorities.
Together these four authorities investigated the potential to convert an STP into an
energy factory. This joint approach generated so much enthusiasm that it did not
remain limited to these four pioneers; it very quickly became apparent that over
half of Dutch regional water authorities were interested in joining this network
organisation. By becoming a member, every water authority undertook the task to
transform its own STP into a concrete business case, setting out first the economic
target, and then actually implementing the conversion. The projects described in
the second part of this chapter are the tangible results.
The water authorities have come to agreement with the national government on
several issues.
– Multiyear Arrangement (MYA): in the period of 2005 to 2020, the
annual energy efficiency increases with 2% so in 2020 the specific energy
consumption is 30% less compared to 2005.
– Climate Agreement/Energy Agreement: in 2020, 40% of the energy
consumed by regional water authorities should be self-generated.
When national and international policies get connected on climate and energy,
the political support on a concept promoting these goals is increasing. This gave
an extra boost to this concept. Likewise, the national government is enthusiastic
about the water authorities’ approach. This led to the 2011 ‘Green Deal’ between
the Dutch Water Authorities and the ministries of Economic Affairs, Agriculture
& Innovation and Environment & Infrastructure. In this deal, apart from making
commitments to eliminate a wide range of impediments, the government also
allocated a research budget of half a million euro. This research budget has been
used to make further technological development in the field of gasification and
supercritical gasification possible.
13.1.4 Economic aspects
The treatment of wastewater costs energy and generates sludge alongside purified
wastewater. In the Netherlands, sewage sludge can no longer be used in agriculture,
due to the high content of heavy metals such as copper and zinc (Disposal of
sewage sludge by destination). As a result, most sludge is incinerated in mono-
incinerators. Generally speaking, the costs of sludge incineration are considerable
(€40–€60 per tonne of sludge cake). Converting this sludge into biogas at this
point serves two purposes: (1) The extra biogas can be converted into electricity,
reducing the amount of electricity that needs to be purchased elsewhere. (2)
It reduces the amount of sludge remaining, which reduces the costs of sludge
4The STP Apeldoorn is being implemented in phases. This refers to the last section (see also case
description).
processing. In addition, it has been shown that some measures have a positive
effect on sludge dewatering, meaning that less water need be transported and
evaporated. This combination makes the conversion of STPs into energy factories
financially attractive as well. The Investments for techniques like TPH are high
and are only attractive for larger STP. Therefore in practise they are mostly
combined with centralised sludge treatment. The same applies for the recovery
of phosphate. Struvite deposits can lead to extra maintenance costs, especially
when a higher concentration of dry matter is treated in the digestion process.
By removing the phosphate in a controlled manner (in the form of struvite), this
product can be sold separately, and it also means reduced maintenance costs and
less sludge.
Table 13.1 Impact of the energy factory on the energy production in future.
13.2 Cases
In this section different cases are described. A general description of the plants
is given in Table 13.2 and Table 13.3 gives an overview of the measures taken to
transform the STP into an energy factory.
by marinhoeng@hotmail.com
Table 13.2 Description of the WWTP before transformation to an energy factory.
Plant Echten Plant Apeldoorn Plant Plant Tilburg Plant Venlo Plant Den Bosch Plant
Olburgen Amersfoort
Design capacity 169.000 317.000 137.500 350.000 300.000 310.000 307.000
in p.e. based on
150 g TOD (total
oxygen demand)
Waterline
Primary treatment Screens (2 × 3 mm), Screens, primary Screens, Screens, grit Screens, grit Screens, primary Screens, primary
grit- and sand clarifiers, grit- and primary and sand and sand clarifiers, grit- and clarifiers, grit-
removal sand removal clarifiers, removal, removal sand removal and sand removal
grit- and sand primary
removal clarifiers
by marinhoeng@hotmail.com
thickener thickener thickener
and After
transformation
Gravity
followed by
belt thickener
Secondary Sludge Gravity thickener Belt thickener pre Gravity Belt thickener Gravity Stirred up with the Belt thickener
followed by Belt thickening existing thickener thickener primary sludge and (4,55) and
thickener (new) BT, after buffer end (2 × 350 m3) treated with gravity after Wasstrip
thickening with ZBP thickeners thickening by
to 12% ZBP to 13%
Sludge digestion None/1 mesophilic 41 mesophilic tanks 2 mesophilic 2 mesophilic None/2 2 mesophilic tanks 3 mesophilic
Before/after & 1 thermophilic parallel (before and tanks (before tanks mesophilic (before and after tanks (before
tank after) and after) 3 mesophilic tanks and after)
tanks
Dewatering Plate and frame filter 3 centrifuges 2 centrifuges 2 centrifuges 2 centrifuges 1 centrifuge 2 centrifuges
press
Using the biogas CHP CHP (future CNG Heating purpose CHP (future – CHP (future biogas CHP
and LNG), heating at potato CNG?) and LNG)
households processing
industry
Energy 3,5 MWh/year 9,5 MWh/year 2,4 MWh/year 7,7 MWh/year 10 MWh/year 5,3 MWh/year 6,7 MWh/year
consumption before Incl bellen
transformation to
an Energy factory
Energy 0 12,3 MWh/year 1,0 MWh 5,0 MWh/year 0 MWh 4,5 MWh/year 3,1 MWh/year
production before 22.4 GJ/year Heat
transformation to
an Energy factory
(Continued)
259
Table 13.2 Description of the WWTP before transformation to an energy factory (Continued). 260
by marinhoeng@hotmail.com
Plant Echten Plant Apeldoorn Plant Plant Tilburg Plant Venlo Plant Den Bosch Plant
Olburgen Amersfoort
Energy 4,5 MWh/year 10 MWh/year 1,8 MWh 7–8 MWh/year 7,5 MWh/year 7,1 MWh/year
consumption after (mainly because of
transformation to the double sludge
an Energy Factory amount to be treated)
Energy 4,7 MWh/year 18.7 MWh/year Extra biogas 2–3 MWh/year 0 MWh 8.2 MWh/year
production after 30 GJ/year Heat 300.000 m3/
transformation to year (replacing
an Energy Factory natural gas)
Energy 420.000 euro −160.000 €/year 290.00 euro/ 2.688.000 euro/ 1.000.000 100.000 euro/year 370.000 €/year
costs before year year euro/year
transformation to
an Energy Factory
Energy costs after 60.000 euro −520.000 €/year 150.000 euro/ 700.00 – 750.000 – − 140.000 €/year
transformation to year 800.000 euro/ 1.200.000 = 450.000
an Energy Factory year euro benefit
Sludge disposal 2.700.000 euro/year 5.900.000 euro/year Equal 6.775.000 euro/ 1.530.000 785.000 euro/year 2.900.000 euro/
costs before 2.200.000 euro/year 3.400.000 euro/year year euro/year 1.410.000 euro/year year
and after 765.000 euro/ 2.400.000 euro/
transformation year year
Biogas production 0 6.000.000 Nm3/year 700.000 Nm3/ 2.480.000 Nm3/ 0 m3/year 2.0000.000 NM3/ 2.200.000 Nm3/
before and after 2.000.000 Nm3/year 9.000.000 Nm3/year year year 2.240.000 Nm3/ year year
transformation 1.000.000 Nm3/ 9.850.000 Nm3/ year 4.0000.000 NM3/ 4.200.000 Nm3/
year year year year
Investment costs 12.500.000 euro 10.000.000 euro 900.000 euro 6.000.000 euro 13.000.000 euro 10.500.000 euro
1
2 mesophilic tanks for the digestion of fluid waste.
Table 13.3 Overview of the measures taken to transform the STP into an energy factory.
by marinhoeng@hotmail.com
STP Aeration Digestion Return on Capacity of Nutrient Biogas Remarks
measures measures investment sludge processing reclamation production
time installation (Nm3/
(tonnes DS/year) years)
& reduction of slib
quantity (%)
Tilburg Anammox TPH <10 years 25,900 Phosphate 8,000,000 In the future, potential
45% DS 438 tonnes partnership with fruit/
59% ODS struvite veg waste and manure
digestion
‘s-Hertogenbosh Reference Depending on +/− 8 years 19,900 Phosphate. 4,000,000 Supply of bio-LNG to
design based tender. Reference 29% Is a the municipal cleaning
on a Nereda design based desirable, service, and supply of
system for the on thermophilic but not an biogas to industry
extra aeration digestion obligation
Amersfoort DEMON, TPH hydrolysis of 6.7 years 11,555 Phosphate 4,200,000 In the future, potential
Table 13.4 Comparison in costs (euro) between de different alternatives for the
STP ‘s-Hertogenbosch.
Table 13.5 Comparison in costs (euro) between fuels for the municipality of
‘s-Hertogenbosch.
Figure 13.1 Overview STP Echten of the waterboard Reest & Wieden.
Supplying the industrial customer directly means that the gas does not have to be
transported to the natural gas network, so it does not need to be refined to natural
gas quality. Aviko uses the natural gas for heating purposes, which means the
energy content of the biogas is utilized virtually completely. Waterstromen, which
at present also has a surplus of residual heat, supplies heat to the STP to heat the
digestion process. This heat, which is released from the generation of power in the
CHP, was up to now dissipated as waste heat. The earn-back period of the first four
measures is 1 to 2 years. The earn-back period of the last measure, replacement of
the bubble aerators, is longer (15 years). On balance, the earn-back period is 7 years.
Additionally, all the concentrated waste water from the digestion is now routed
from the STP to the struvite reactor, so this phosphate is also reclaimed, creating
a ‘phosphate factory.’ The partnership that has now been initiated is seen as a
first step toward further cooperation. Further advantages of synergy that could be
obtained from scaling up biogas production from the digestion of waste, cuttings,
and so on are now being considered.
Tendering
Because the water authorities’ expertise in the area of optimisation of biogas
production is limited, the choice was made to adjust the tendering process
accordingly. The black box approach was used, in which the request is defined in
functional terms and it is then up to the market to find the best solution. This puts
the responsibility and process guarantee completely in the hands of the market.
The market is also responsible for adequate reduction of the extra nitrogen load
released with the digestate.
Figure 13.2 gives an overview of the STP Tilburg before the transformation to
an Energyfactory.
6 Normally, this figure is 14,000,000 kWh, but due to maintenance this was less in 2013.
Table 13.6 Guarantees that have been received for Amersfoort STP.
The drying of the sludge with residual heat at STP Amersfoort has proven to
be impossible within the project period. The choice has now been made for a pilot
at STP Ede in the 2014–2016 period. The results of this pilot will be reworked for
STP Amersfoort in 2015.
13.3 Conclusion(S)
It has been shown that the innovative approach of the Dutch regional water
authorities has led to a broadly supported concept with a new perspective on
wastewater (Waste as a Resource for Energy and Raw materials). The results of
this new concept have been proven by applying it in practice directly and initially
on a small scale. The coordinated, sector-wide approach has further increased the
support base for it. Additionally, this has also given the industry the freedom to
put their innovative ideas into practice, with the regional water authority taking
on the role of launching customer. The projects were not limited to the field of
wastewater purification, but instead the STPs were considered as a component of
the environment. As a result broad-spectrum solutions with greater benefit to the
environment and financial profits arose. The more sustainable way need not always
to be the more expensive way.
13.4 References
Energy Factory – Water Boards Inside Out, February 2009, www.energiefabriek.com.
(accessed 18 February 2014).
Wastewater management roadmap towards 2030 – A sustainable approach to the collection
and treatment of wastewater in the Netherlands, November 2013, www.uvw.nl. (accessed
18 February 2014).
Verslag specialistendag voorbehandelingstechnieken 21 juni 2011 (STOWA, Energiefabriek).
(Report Specialits day pre-treatment techniques 21 juni 2011). www.stowa.nl. (accessed
18 February 2014).
Afzet van zuiveringsslib naar bestemming, 1981–2011. (Disposal of sewage sludge by
destination 1981 – 2011), CBS, 2013 http://www.compendiumvoordeleefomgeving.nl/
indicatoren/nl0154-Afzet-van-zuiveringsslib-naar-bestemming.html?i=1–4. (accessed
18 February 2014).
Nereda® praktijk onderzoek 2010–2012 (Nereda® practical research 2010–2012), STOWA,
2013, Report 29.
NEWater, Op weg naar de RWZI 2030 (The STP of the future 2030), STOWA, 2010,
Report 11.
14.1 Introduction
The cleaning performance of sewage treatment plants (STPs) is typically
expressed by removal rates (e.g., of nitrogen (N) or phosphorus (P)). There are
also mandatory emission limits, such as a N-removal rate in excess of 70% at
temperatures >12°C for all STPs greater than 5,000 population equivalents (PE)
as well as an NH +4 concentration of less than 5 mgN/L (Phillippitsch & Grath,
2006), (Thaler, 2009). The advantage of using the N-removal rate as a measure
of cleaning performance is that it can be easily calculated. The N-compounds
are quantified by the N-content of the influent and effluent. However, in
this way, N is not differentiated in its various chemical species (e.g., NH +4 ,
NO3− , N2 etc.); thus, NH +4 and NO3− have equal weighting factors in the total
N-content, although the environmental impact of these two forms of N are
different. The Austrian Water Act and Life Cycle Impact Assessment (LCIA)
assign different emission limits and impact factors to individual N-compounds
(BGBl II Nr, 98/2010), (BGBl II Nr, 99/2010), (BGBl II Nr, 461/2010), (Guinée
et al. 2002)). For example, NH +4 has a higher eco-toxicological impact on water
bodies than NO3− and therefore a lower emission limit. Other studies have
concluded that the role and fate of aquatic Norg is still unknown, which implies
that Norg in water bodies cannot be regarded as NH +4 (Westgate & Park, 2010).
N2O plays a major role in global warming (Kampschreur et al. 2009). The
omission of gaseous N-compounds is a drawback in the use of the N-removal
Step 2: normalization of M i
the mass flow mim = (Eq. 1)
∑∑ i
X
m
im
where
X im = M i * c im (Eq. 2)
Step 3: calculation of the cim − cim,geog
diluting masses for each ′ = mi *
mim * 100 + mi (Eq. 3)
cim,geog
N-compound
and
cim * cim,geog
′ =
cim 0.01 (Eq. 4)
cim − 0.99 * cim,geog *
∑∑m
Step 4: computation of the
H (mim ′ )= −
′ , cim ′ * log2 (cim
* cim ′ ) (Eq. 5)
statistical entropy im
i m
M i is the measured mass flow in kg per day. The index i refers to the wastewater,
effluent, off-gas, sludge, and the environmental compartments atmosphere
and hydrosphere. The variable cim (in kgN/kg) corresponds to the measured
concentration of a N-compound (m) in the particular mass flow i. The background
concentration of a compound m in an environmental compartment i is indicated
by cim,geog. The mass flows are normalized according to Equation 14.1. The
denominator of Equation 1 equals the total flow of N through the STP such that
the masses mi are related to one mass-unit of N (e.g., kg in the effluent per kg of
the total N throughput). Normalization is required to make processes of different
size comparable. In the next step, the diluting masses m’im and the corresponding
concentration terms c’im are calculated according to Equation 3 and Equation 4,
respectively. The mass-function calculates how much water or air is needed to
dilute the emitted concentration to its corresponding background concentration.
The dimensionless mass-function for the N in the sludge is computed according
to Equation 1. The statistical entropy H is then calculated according to Equation
5 for every N-compound in the effluent, the atmosphere and the sludge. Thus, the
statistical entropy is a function of the mass flows, such as the wastewater, effluent,
air and sludge, the emission concentrations and the corresponding background
concentrations of all N-compounds. More diluted mass flows and larger differences
between the emitted and the background concentrations result in increased
dilution in the environment and, consequently, in higher entropy values. Given
the assumption that dilution should be avoided whenever possible for sustainable
resource management, low entropy values are desired. The entropy values of all
emitted N-compounds are added to obtain the total statistical entropy of N after
WWT (HafterWWT). This value is compared with the statistical entropy determined
for a hypothetical scenario in which the untreated wastewater is directly discharged
into the receiving waters (HnoWWT). The benefit of a STP for N-treatment is then
expressed as the reduction in the statistical entropy (ΔH) relative to the direct
discharge of wastewater into receiving waters (see Figure 14.1b). A higher ΔH
value indicates a more favorable performance of a particular STP because it results
in less dissipation of N-compounds into the environment. The main advantage of
eSEA for the assessment of the N-performance of STPs is that all N-compounds,
as well as the flows of wastewater, effluent and sludge, can be considered both
qualitatively and quantitatively. The environmental impact is reflected in the
quantification of the dilution. Thus, a comprehensive evaluation of the N-removal
performance of a STP can be provided by this analysis. The disadvantages of eSEA
are the relatively large data requirement, the inability to perform the calculation
with a single equation, in contrast to the estimation of the N-removal rate, and the
lack of other studies that can be used to compare the results.
STP N°
1 2 3 4 5
0
entropy ∆H and N -removal rate
Reduction in statistical
60
70
[%]
80
90 eSEA
N-removal rate
100
All of the STPs report N-removal rates of ca. 75%. Consequently, these plants
would be considered to exhibit equally good cleaning performances. The eSEA
results, however, reveal that there are differences in the N-removal performance of
the individual STPs. STP N°5 achieves the highest reduction in statistical entropy
(ΔH = 85%) and is thus the most favorable one. STP N°4 is the least favorable,
with a ΔH of 73%. STP N°1 achieves a ΔH of 74%, STP N°2 has a ΔH of 77%, and
STP N°3 attains a ΔH of 83%. In Table 14.2 the proportion of each N-compound
in the total N amount and its contribution to the statistical entropy after WWT
(HafterWWT), which has a direct influence on the reduction in the statistical entropy
ΔH, are summarized.
STP N°4 emits more N in the form of Norg in the effluent to the river (15%) than
does STP N°5 (5%). Norg represents a composition of various organic compounds
that are naturally present in scarce quantities in rivers with a good ecological
status. Norg is therefore the main contributor to the statistical entropy after WWT
by STP N°4. STP N°5 achieves a higher denitrification rate because most of the
N is emitted as N2 (74%). By contrast, only 56% of the N leaves STP N°4 as N2.
The emission of N2 into the atmosphere does not generate entropy because of the
high concentration of N2 that is already present in the atmosphere (75% mass
fraction). Similar considerations can be made for the other STPs. These examples
demonstrate that the N-removal performance of a STP can appear quite different
if the different N-compounds, their distribution in the individual mass flows and
their dilution in the environment are considered.
long-term water quality. However, a very high concentration of O2,diss. can lead to
undesirable effects due to the incomplete denitrification caused by the recirculation
of O2 from the aerobic to the anoxic reactor thus inhibiting the denitrification
process (Flores-Alsina et al. 2011). To compare STPs of different sizes, the values
of both the energy-consumption and the costs are divided by the individual PE.
The energy-efficiency is defined as the energy consumption that is required for
every PE to achieve a reduction in statistical entropy. The cost-effectiveness is
calculated as the ΔH per PE-specific costs, which defines the cost-effectiveness
of the N-treatment as the reduction in the statistical entropy that is achieved for
every EUR and PE. In Table 14.3 the N-removal performance of the 5 analyzed
Austrian STPs according to eSEA and the energy-efficiency and cost-effectiveness
are presented. The best performances are indicated in bold.
According to the eSEA results, STPs N°3 and N°5 exhibit the best N-removal
performance. However, STPs N°2 and N°3 exhibit the most energy-efficient
N-treatment, and STP N°1 achieves the highest cost-effectiveness. The results
indicate that an energy-efficient N-removal performance does not necessarily
imply cost-effectiveness. Because all five STPs comply with Austrian mandatory
emission standards, it is reasonable to nominate the most energy-efficient or the
most cost-effective STP as the best practice STP, which in this case would be
STP N°1, N°2 or N°3. A different approach would be to propose the best practice
STP as the plant with good results in all 3 categories, which, according to the
analysis presented in this work, would be STP N°3. In Austrian benchmarking,
cost-effectiveness plays the decisive role, which is reasonable because the costs
for energy consumption are included and the relationship between the energy
consumption and the energy costs is usually proportional (Lindtner, et al. 2002),
(Lindtner, 2009). The low cost-effectiveness of STP N°3 compared to STP N°1, for
example, originates from higher expenses for labor and external services.
Table 14.4 Overview over the number of STPs among the different size groups.
Size group/PE
10,000 20,000 30,000 40,000 50,000 60,000 70,000 80,000 90,000 >100,000
– – – – – – – – –
20,000 30,000 40,000 50,000 60,000 70,000 80,000 90,000 100,000
Number of STPs among the different size groups
10 10 8 10 1 4 0 2 1 10
STPs < 50,000 PE STPs > 50,000 PE
Total number of small and large STPs
38 18
0.50
0.45
Energy consump on per reduc on
0.40
in sta s cal entropy and PE
0.35
[kWh/(ΔH%*PE)
0.30
0.25
0.20
0.15
0.10
STPs > 50,000 PE
0.05
STPs < 50,000 PE
0.00
90
80
Reducon in stascal entrpoy per
costs of the mechanical biological
70
treatment [ΔH%/( €*PE)
60
50
40
30
20
STPs > 50,000 PE
10
STPs < 50,000 PE
0
14.4 Conclusions
The use of eSEA offers a more comprehensive assessment of the N-removal
performance of STPs than the N-removal rate because it considers different
N-compounds, including gaseous emissions; the distribution of the N in the
wastewater, effluent, and sludge; and the dilution of the emissions in the
environment. The application of eSEA rewards STPs that transform and transfer
N-compounds from the wastewater into harmless (or less harmful) species, such
as N2 or NO3− , instead of into NH +4 or Norg, which would be discharged into water
bodies. The eSEA results can be related to the energy-consumption and costs of the
N-treatment. Thus, the evaluation can be extended to economic factors. The results
of the analysis of 5 Austrian STPs demonstrate that an energy-efficient plant is not
necessarily cost-effective. The N-removal performances, the energy-efficiencies
and cost-effectiveness of 56 different size STPs are compared, revealing that
individual, small STPs (10,000–50,000 PE) are able to compete with the larger
plants (50,000–950,000 PE). These results can contribute to the discussion about
the advantages and disadvantages of different size STPs offering a new perspective
on the efficiency of small, decentralized STPs.
14.5 References
Agis H. (2002). Energieoptimierung von Kläranlagen (Energy optimization of waste
water treatment plants). Wiener Mitteilungen, Band 176, pp. 133–177.
BGBl II Nr, 9., 99/2010. Verordnung des Bundesministers für Land- und Forstwirtschaft,
Umwelt und Wasserwirtschaft über die Festlegung des ökologischen Zustandes für
Oberflächengewässer Teil II (Quality Objective Ordinance – Ecological Status of
Surface Waters), Vienna, Austria; Bundesministerium für Land- und Forstwirtschaft,
Umwelt und Wasserwirtschaft (Austrian Ministry of Agriculture, Forestry,
Environment and Water Resources).
BGBl II Nr, 4., 461/2010. Verordnung des Bundesministers für Land- und Forstwirtschaft,
Umwelt und Wasserwirtschaft über die Festlegung des chemischen Zustandes für
Oberflächengewässer Teil II (Quality Objective Ordinance – Chemical Status of
Surface Waters), Vienna, Austria; Bundesministerium für Land- und Forstwirtschaft,
Umwelt und Wasserwirtschaft (Austrian Ministry of Agriculture, Forestry,
Environment and Water Resources).
Flores-Alsina X., Corominas L., Snip L. and Vanrolleghem P. (2011). Including greenhouse
gas emissions during benchmarking of wastewater treatment plant control strategies.
Water Research, 45, 4700–4710.
Guinée J. B., Gorrée M., Heijungs R., Huppes G., Kleijn R., Koning A. de, Oers L. van,
Wegener Sleeswijk A., Suh S., Udo de Haes H. A., Bruijn H. de, Duin R. van and
Huijbregts M. A. J. (2002). Handbook on life cycle assessment. Operational guide
to the ISO Standards. I: LCA in perspective. IIa: Guide. IIb: Opeartional annex. III:
Scientific background. Kluwer Academic Publishers, Dordrecht, The Netherlands.
Hernandez-Sancho F. and Sala-Garrido R. (2009). Technical efficiency and cost analysis in
wastewater treatment processes: a DEA approach. Desalination, 249, 230–234.
Kampschreur M. J., Temmink H., Kleerebezem R., Jetten M. S. M. and van Loosdrecht M.
C. M. (2009). Nitrous oxide emission during wastewater treatment. Water Research,
43, 4093–4103.
Kaufman S., Kwon E., Krishnan N., Castaldi M. and Themelis N. (2008). Use of Statistical
Entropy and Life Cycle Analysis to Evalaute Global Warming Potential of Waste
Management Systems. Proceedings of NAWTEC16, 16th Annual North American
Waste-to-Energy Conference May 19–21–2008, Philadelphia, Pennsylvania, USA.
Lindtner S. (2009). Benchmarking: Grundlagen/Praxiserfahrungen (Benchmarking:
Basics and experiences), k2W Enginnering Kaltesklareswasser, Vienna, Austria.
Lindtner S. (2010). Benchmarking für Kläranlagenfür das Geschäftsjahr 2009
(Benchmarking of WWTPs for 2009), k2W Engineering Kaltesklareswasser, Vienna,
Austria.
Lindtner S., Nowak O. and Kroiss H. (2002). Benchmarking für Abwasserreinigungsanlagen
(Benchmarking of WWTPs). Institute for Water Quality, Resource and Waste
Management, Vienna University of Technology, Vienna, Austria.
Lindtner S., Schaar H. and Kroiss H. (2008). Benchmarking of large municipal waste water
treatment plants >100,000 PE in Austria. Water Science and Technology, 57(19),
1487–1493.
Kroiss H. and Svardal K. (2009). Energiebedarf von Abwasserreinigungsanlagen (Energy
consumption of wastewater treatment plants). Institute for Water Quality, Resource
and Waste Management, Vienna University of Technology, Vienna, Austria.
Kroiss H., Paraviccini V. and Svardal K. (2007). Abschätzung der Veränderung der
klimarelevanten Gase CO2, CH4 und N2O am Standort der Regionalkläranlage Linz-
Astendurch den Betrieb der Schlammbelüftungsanlage (Estimation of the climate
relevant gases CO2, CH4, and N2O from an Austrian WWTP), Institute for Water
Quality, Resource and Waste Management, Vienna University of Technology, Vienna,
Austria.
Nowak O. (2002). Energie-Benchmarking von Kläranlagen – Überlegungen aus
abwassertechnischer Sicht (Energy-benchmarking of WWTPs). Institute for Water
Quality, Resource and Waste Management, Vienna University of Technology, Vienna,
Austria.
Phillippitsch R. and Grath J. (2006). Erhebung der Wassergüte in Österreich – Jahresbericht
2006 A/3 Fliessgewässer (Evaluation of the water quality 2006), Vienna, Austria:
Bundesministerium für Land- und Forstwirtschaft, Umwelt und Wasserwirtschaft
(Ministry for Land, Forestry, Environment and Water Resources).
Rechberger H. (2001a). An entropy based method to evaluate hazardous inorganic
substance balances of waste treatment systems. Waste Management and Research,
19, 186–192.
Rechberger H. (2001b). The use of the statistical entropy to evaluate the utilisation of
incinerator ashes for the production of cement. Waste Management and Research, 19,
262–268.
Rechberger H. and Brunner P. H. (2002). A new, entropy based method to support waste
and resource management decisions. Environmental Science and Technology, 34(4),
809–816.
Rechberger H. (2012). Waste-to-Energy (WTE): Decreasing the Entropy of Solid Wastes
and Increasing Metal Recovery. In: Encyclopedia of Sustainability Science and
Technology, Robert A. Meyers (ed.), Springer, New York.
15.1 Introduction
Sewage sludge management is a well-known key point in the operation of biological
WasteWater Treatment Plants (WWTPs): in effect, sludge treatment and disposal
often account for one half of the plant’s operating cost (Neyens et al. 2004;
Saveyn et al. 2008; Ruiz-Hernando et al. 2010; Ozdemir & Yenigun, 2013), so
that wastewater treatment processes may convert a water pollution control problem
into a solid waste disposal problem (Weemaes & Verstraete, 1998). Several EU
research funding programs have been issued in this field during the last decade:
among the latest, ROUTES (Novel processing routes for effective sewage sludge
management) and END-O-SLUDG (Wastewater transformed for good) projects
can be mentioned within the Seventh Framework Programme.
Moreover, sludge production has been unceasingly increasing, as a consequence
of both the growing quantity of collected and treated wastewaters, and the
application of progressively stricter standards for WWTPs effluent quality (Neyens
et al. 2004; Kouloumbos et al. 2008). In EU, the implementation of the European
Urban Waste Water Treatment Directive (UWWTD – 91/271/EEC) led to a strong
rise in sludge generation, up to 50% (Fytili & Zabaniotou, 2008): a total yearly
amount around 10 million tons of dry solids (referred to 2007 for EU27 + Norway)
has been reached (EEA European Environment Agency; Eurostat), and the
projection for the year 2020 is over 13 million (Kelessidis & Stasinakis, 2012).
Similarly, in China approximately 3 million tons of dry solids were generated
during the year 2007 (Wang et al. 2010).
by marinhoeng@hotmail.com
Table 15.1 Items used for the calculation of the total treatment cost for sludge dewatering (as Euro per ton of total dry solids of
fed sludge).
by marinhoeng@hotmail.com
hman: daily man hours spent by the operators for machine run [%]
Sludge disposal: cost due to the final disposal of S.D. =u.c.sludge disposal ⋅ SL
dewatered sludge SL: mass flow of dewatered sludge [tdewat sludge/t TS fed sludge]
Supernatant Oxygen demand: cost due to O2 =[0.5 ⋅ α ⋅ BOD + γ ⋅ (Ntot – Nass)] ⋅ u.c.O2
recirculation oxygen consumption during α: specific oxygen consumption for BOD oxidation [kg O2/kg BOD]
(treated with a the biological purification of BOD: BOD load in the supernatant [kg BOD/t TS fed sludge]
conventional supernatant recirculated in the γ : specific oxygen consumption for ammonia oxidation [kg O2/kg N]
biological wastewater treatment line (the Ntot: nitrogen load in the supernatant [kgN/t TS fed sludge]
process) hydrolyzed fraction, assumed Nass: nitrogen load assimilated by the activated sludge
a half of the total BOD) [kgN/tTS fed sludge]
u.c.O2: cost for the supply of 1 kg O2 [€/kg]
Supernatant sludge S.S.T.D. BOD ⋅ Yobs [TS]supern ⋅ SUP
treatment and disposal: = 0 .5 ⋅ +
TSout TSout ⋅ u.c.sludge disposal + 0.5⋅
Reprinted from Journal of Environmental Management, 132, Bertanza, G., Papa, M., Canato, M., Collivignarelli, M. and Pedrazzani, R., How can
sludge dewatering devices be assessed? Development of a new DSS and its application to real case studies, 86–92, (2014), with permission
287
from Elsevier.
288 Sewage Treatment Plants
15.2.2 Data processing
Experimental data processing is aimed at:
15.3.1 Technical issues
Technical performances are graphically summarized in Figure 15.2 as the
mean ± standard error, chosen in lieu of standard deviation to evaluate data
variability, being aware of the modest number of samples collected during each
test (n = 4). In summary:
25%
20%
15% 24%
[%]
23%
10%
22%
21%
5% I I* II II* III IV IV* V VI
0%
I I* II II* III IV IV* V VI
4,000
COD concentration in supernatant [mg/L]
3,000
2,000
1,000
0
I I* II II* III IV IV* V VI
test
Centrifuge 1 Centrifuge 2 Centrifuge 3
Figure 15.2 Dewatering efficiency of tested devices: total dry solids concentration
in dewatered sludge (up) and COD concentration in supernatant (down). Source:
Reprinted from Journal of Environmental Management, 132, Bertanza, G., Papa,
M., Canato, M., Collivignarelli, M. and Pedrazzani, R., How can sludge dewatering
devices be assessed? Development of a new DSS and its application to real case
studies, 86–92, (2014), with permission from Elsevier.
15.3.2 Economic issues
The total treatment cost was calculated under the following assumptions (i.e., the
numerical value of coefficients appearing in Table 15.1):
• α = 0.5 kg O2/kg BOD (explanation in Table 15.1);
• γ = 4.5 kg O2/kg N;
• hman = 1 h/8 h;
• I = 105,000 €, 125,000 € and 150,000 € for centrifuges #1, #2 and #3,
respectively; capital costs do not include equipment other than the device itself
(e.g., accessory apparatuses, the hosting building, . . .), because they were already
available at WWTP, and, anyway, they are exactly alike for all the machines;
• n = 10 y;
• r = 8%;
• u.c.cond = ranging from 1.5 to 1.7 €/kg;
• u.c.E.E. = 0.10 €/kWh;
• u.c.H2O = 1 €/m3;
• u.c.O2 = 2.63 €cent/kg;
• u.c.sludge disposal = 100 €/tdewatered sludge;
• y.c.labor = 35,000 €/y;
• Yobs = 0.5 kg TS/kg BOD.
Figure 15.3 Economic assessment of tested devices: (a) weight of different cost
items (the meaning of the labels is reported in Table 15.1); (b) cost as a function
of dewatering efficiency; (c) overall comparison among tested devices. Source:
Reprinted from Journal of Environmental Management, 132, Bertanza, G., Papa,
M., Canato, M., Collivignarelli, M. and Pedrazzani, R., How can sludge dewatering
devices be assessed? Development of a new DSS and its application to real case
studies, 86–92, (2014), with permission from Elsevier.
15.4 Conclusions
This chapter presented a practical tool to assess and rank sludge dewatering
devices. First of all, for the experimental procedure and the data processing, a well-
established and standardized methodology, which can be easily followed whenever
a device has to be evaluated, has been implemented. As model outputs, the tool
provides both a technical (dewatering efficiency + process parameters) and an
economic (treatment cost, split up for item) overview. Thanks to a comprehensive,
objective and detailed representation of the costs, it is possible to drive the decision
making process through an economic-criterion (as usual for sectors with limited
economic resources, such WWTPs); anyway, the final decision on the most suitable
device can be also supported by the other technical parameters analyzed by the
DSS, the importance of which depends on the case-specific features.
In order to validate the model, it was applied to a real case study, where three
industrial size devices (mobile centrifuges) were assessed and compared with
the dewatering system installed in the WWTP. The latter was proven to be not
convenient, with treatment costs higher (>10%) with respect to the tested centrifuges.
Among them, the cheapest one was identified (#2), and the role of ‘most suitable’
was assigned to it, also weighing the technical factors (mainly machine flexibility).
Moreover, as a general outcome, a conflict between the technical and the economic
‘optimum’ was highlighted: the minimum treatment cost may not correspond to
the maximum achievable sludge TS concentration.
In conclusion, the plant manager entrusted to the model can identify the ‘best
technology’ for this unit of treatment.
15.5 References
Aziz A. A. A., de Kretser R. G., Dixon D. R. and Scales P. J. (2000). The characterisation
of slurry dewatering. Water Science and Technology, 41(8), 9–16.
Bertanza G., Papa M., Canato M., Collivignarelli M. and Pedrazzani R. (2014). How can
sludge dewatering devices be assessed? Development of a new DSS and its application
to real case studies. Journal of Environmental Management, 137, 86–92.
Boran J., Houdkova L. and Elsaesser T. (2010). Processing of sewage sludge: dependence
of sludge dewatering efficiency on amount of flocculant. Resources Conservation and
Recycling, 54(5), 278–282.
Chen C., Zhang P., Zeng G., Deng J., Zhou Y. and Lu H. (2010). Sewage sludge conditioning
with coal fly ash modified by sulfuric acid. Chemical Engineering Journal, 158(3),
616–622.
Dursun D., Turkmen M., Abu-Orf M. and Dentel S. K. (2006). Enhanced sludge
conditioning by enzyme pre-treatment: comparison of laboratory and pilot scale
dewatering results. Water Science and Technology, 54(5), 33–41.
Emir E. and Erdincler A. (2006). The role of compactibility in liquid-solid separation of
wastewater sludges. Water Science and Technology, 53(7), 121–126.
Fytili D. and Zabaniotou A. (2008). Utilization of sludge in EU application of old and new
methods – a review. Renewable & Sustainable Energy Reviews, 12(1), 116–140.
Gratziou M., Ekonomou S. and Tsalkatidou M. (2005). Cost analysis and evaluation of
urban sewage processing units. Water Science and Technology: Water Supply 5(6),
155–162.
Hong J., Hong J., Otaki M. and Jolliet O. (2009). Environmental and economic life
cycle assessment for sewage sludge treatment processes in Japan. Waste Management,
29(2), 696–703.
http://epp.eurostat.ec.europa.eu/ (accessed 18 November 2014).
http://www.eea.europa.eu/data-and-maps/data/waterbase-uwwtd-urban-waste-water-
treatment-directive-1 (accessed 18 November 2014).
http://www.end-o-sludg.eu/ (accessed 18 November 2014).
http://www.eu-routes.org/ (accessed 18 November 2014).
Kelessidis A. and Stasinakis A. (2012). Comparative study of the methods used for treatment
and final disposal of sewage sludge in European countries. Waste Management, 32(6),
1186–1195.
Kouloumbos V. N., Schaffer A. and Corvini P. F. X. (2008). The role of sludge conditioning
and dewatering in the fate of nonylphenol in sludge-amended soils. Water Science and
Technology, 57(3), 329–335.
Mamais D., Tzimas A., Efthimiadou A., Kissandrakis J. and Andreadakis A. (2009).
Evaluation of different sludge mechanical dewatering technologies. Journal Residuals
Science & Technology, 6(1), 27–34.
Mulder J. W., van Loosdrecht M. C. M., Hellinga C. and van Kempen R. (2001). Full-scale
application of the sharon process for treatment of rejection water of digested sludge
dewatering. Water Science and Technology, 43(11), 127–134.
Neyens E., Baeyens J., Dewil R. and De Heyder B. (2004). Advanced sludge treatment
affects extracellular polymeric substances to improve activated sludge dewatering.
Journal of Hazardous Materials, 106(2–3), 83–92.
Ozdemir B. and Yenigun O. (2013). A pilot scale study on high biomass systems: energy and
cost analysis of sludge production. Journal of Membrane Science, 428(1), 589–597.
Peng G., Ye F. and Li Y. (2011). Comparative investigation of parameters for determining
the dewaterability of activated sludge. Water Environment Research, 83(7), 667–671.
Qi Y., Thapa K. B. and Hoadley A. F. A. (2011). Application of filtration aids for improving
sludge dewatering properties – a review. Chemical Engineering Journal, 171(2), 373–384.
Ruiz-Hernando M., Labanda J. and Llorens J. (2010). Effect of ultrasonic waves on the
rheological features of secondary sludge. Biochemical Engineering Journal, 52(1),
131–136.
Saveyn H., Curvers D., Thas O. and Van der Meeren P. (2008). Optimization of sewage
sludge conditioning and pressure dewatering by statistical modelling. Water Research,
42(4–5), 1061–1074.
Scholz M. (2005). Review of recent trends in capillary suction time (CST) dewaterability
testing research. Industrial & Engineering Chemistry Research, 44(22), 8157–8163.
Uggetti E., Ferrer I., Molist J. and García J. (2011). Technical, economic and environmental
assessment of sludge treatment wetlands. Water Research, 45(2), 573–582.
Volcke E. I. P., Gernaey K. V., Vrecko D., Jeppsson U., van Loosdrecht M. C. M. and
Vanrolleghem P. A. (2006). Plant-wide (BSM2) evaluation of reject water treatment
with a Sharon-Anammox process. Water Science and Technology, 54(8), 93–100.
Wang W., Luo Y. and Qiao W. (2010). Possible solutions for sludge dewatering in China.
Frontiers of Environmental Science & Engineering in China, 4(1), 102–107.
Weemaes M. P. J. and Verstraete W. H. (1998). Evaluation of current wet sludge disintegration
techniques. Journal of Chemical Technology and Biotechnology, 73(2), 83–92.
Yang Y., Zhao Y. Q., Babatunde A. O. and Kearney P. (2007). Co-conditioning of the
anaerobic digested sludge of a municipal wastewater treatment plant with alum sludge:
benefit of phosphorus reduction in reject water. Water Environment Research, 79(13),
2468–2476.
Yu G. H., He P. J., Shao L. M. and He P. P. (2008). Stratification structure of sludge flocs
with implications to dewaterability. Environmental Science & Technology, 42(21),
7944–7949.
16.1 Introduction
Enhanced nutrient removal in municipal wastewater treatment plants (WWTPs) can
be partly and efficiently carried out by treating the ammonium and phosphorus-rich
reject water (other terms for reject water are ‘return liquor’, ‘digester supernatant’
or ‘sludge digester liquid’) produced from the dewatering of anaerobic digested
sludge in order to meet more stringent effluent standards. In conventional plants
the nitrogen flow from the reject water constitutes 10–30% of the total N-load
(Cervantes, 2009; Gustavsson et al. 2011). As far as phosphorus is concerned, the
concentration in reject water can be up to 130 mg L −1 (Oleszkiewicz & Barnard,
2006; Pitman, 1999; Ivanov et al. 2008). High P concentrations may be reached
when anaerobic co-digestion of sewage sludge and organic waste are applied
(Malamis et al. 2014; Battistoni et al. 2005). Thus, the reject water is returned to
the activated sludge process and accounts for 10 to 50% of the nutrients in the main
stream of municipal WWTPs.
In addition, innovative schemes aiming at energy neutral-positive municipal
WWTPs consider the anaerobic digestion as the core process for biogas recovery
from sewage sludge. As a consequence, the enhanced nutrient removal from the
digester supernatant is proposed to take place separately from the main stream and
becomes a significant stage within the new-conceived WWTPs.
16.1.1 Removal or recovery?
Separate treatment of the digester supernatant requires a minimum extension
because of the high temperature of the supernatant (25–35°C), potentially leading
to short sludge retention times (SRTs) and high reaction rates. Furthermore, a
high ammonium concentration and a low COD:N ratio favor high autotrophic
ammonium reduction rates. The alkalinity content is often around 1.1 mol HCO3
per mol NH +4 -N, while 1.98 mol HCO3 per mol NH +4 -N is required for complete
nitrification, so extra alkalinity is required if complete ammonium oxidation
is needed. Among the different treatment technologies for digester supernatant
(cfr Chapter 5), the innovative biological processes have been proved to be the
most economically sustainable in terms of nitrogen removal. Compared to the
physicochemical processes, these processes do not allow nitrogen recovery.
However, the sustainability of the currently available and future nitrogen removal
systems has been investigated by several authors (Mulder, 2003; STOWA, 2012).
Ammonium can be stripped from ammonium rich side streams (e.g., rejection
water) by means of air stripping. This is a well-known technique. In order to strip
ammonium a high pH is required (pH 10 to 12). Usually NaOH or Ca(OH) are
added as alkali to realize this pH increase. In the air stripping process the rejection
water is led through a stripping column in reverse flow through an air stream. The
ammonia is transferred to the air stream which is led to an absorber. The adsorbed
substance contains acid (H2SO4 or HNO3) in which the ammonia dissolves and
ammonium salts are formed. The ammonium salts are drained from the absorber
while the ammonia free air can be recycled to the stripper (STOWA, 2012).
In general, the energy demand of such a reference stripping varies from 100 to
150 MJ/kg N (aeration, heat, chemicals as well as their cost) and is significantly
higher than the energy demand of the nitrogen producing Haber-Bosch process
combined with Anammox (total 60 MJ/kg N).
This shows that nitrogen recovery is more expensive (1.9–3.2 €/kg N) than
nitrogen removal using Anammox (0.8 €/kg N) because of the higher energy
utilization, as well as the price and quantity of the chemicals required (NaOH or
CaO and H2SO).
In contrast to nitrogen, phosphorus is a limited resource which must be recovered
and reused. It is estimated that the remaining accessible reserves of phosphate rock
will run out in 50 years, if the growth of demand for fertilizers remains at 3% per
year (Gilbert, 2009; Elser & Bennet, 2011). Reducing usage will help the reserves
last longer, but the biggest gains will probably be derived from the recovery of
phosphates, both from wastewaters and livestock waste (Gilbert, 2009).
Struvite (MgNH4PO4 ⋅ 6H2O) is generally considered as the optimal phosphate
mineral for recovery as it contains 51.8% of P2O5 (based on MgNH4 PO2) and
et al. 2012) as well as from low strength effluents such as domestic wastewater
(Blackburne et al. 2008). Nitritation/denitritation has been examined for various
environmental and operating conditions, including low and high DO concentrations
in the reactor (Pollice et al. 2002; Blackburne et al. 2008; Guo et al. 2009a), high
temperature (Hellinga et al. 1998), high salinity (Ye et al. 2009), different FA and
free nitrous acid (FNA) concentrations (Park et al. 2010), different solids retention
times (SRTs) (Pollice et al. 2002).
SCNR can be conveniently coupled with suitable bioprocesses for phosphorus
removal through its accumulation in biomass and can be a sustainable option
resulting in the production of high added value products. In fact, the mechanism
of phosphorus uptake can be realized under anoxic conditions by denitrifying
phosphorus accumulating organisms (PAOs) that can utilize nitrate or nitrite
as electron acceptors (Kishida et al. 2006; Carvalho et al. 2007). Denitrifying
PAOs require less carbon source compared to aerobic PAOs (Li et al. 2011; Peng
et al. 2011). The rate of phosphate uptake can be higher in the presence of nitrite
compared to nitrate (Lee et al. 2001).
to the year 2001 and to design guidelines and technical data given by the authors
van Dongen et al. (2001).
Table 16.1 Full scale cost of Sharon-AnAmmOx process (van Dongen et al. 2001).
On the other hand, Siegrist et al. (2012) reported the cost of ammonia stripping
versus Nitritation/Anammox in SBR pointing out the 50% cost savings using the
biological processes (Table 16.2).
NH3-Stripping Nitritation/Anammox
WWTP Opfikon WWTP St. Gallen-Au
19.2 t NH4 –Nelim/year 46.6 t NH4 –Nelim/year
Operating costs (chemicals, 2.50 0.60
energy, sludge disposal)
Maintenance costs (spare parts) 1.50 0.20
Personnel costs (25–30% of site) 1.50 0.70
Proceeds of sale of fertilizer 0.60 –
Capital cost 3.50 2.70
Net cost 8.40 4.20
Volcke et al. (2007) reported economic evaluations on the basis of the operating
cost index (OCI) for a simulated WWTP treating 21100 m3/day, which resulted in
reject water of 172 m3/day (Table 16.3).
As opposed to nitrogen, phosphorus is a non-renewable resource. There is a
wide range of technologies to remove and recover phosphorus from wastewater,
acquisitions (of SC Process units of 200–400 k€), the actual costs of which may be less
than the current estimates used in these calculations, significantly increasing profits
returned: (a) Growth in net asset value from zero to $3.8 million in 5 years; (b) High
returns on equity of 44% in Year 5; (c) Profits returned in Year 2, profits consistently
increasing to $1.5 million in Year 5; (d) Most of the equity is held in physical assets.
However, besides the reasons concerning the agronomic properties and the
cost stated above (Hao et al. 2013), struvite may be not easily marketable due
to legislation constraints in some EU countries which can influence a lot the
economical sustainability of the scheme anammox + struvite.
16.4 S.C.E.N.A. System
16.4.1 Pilot-scale results
16.4.1.1 S.C.E.N.A. system integrated in co-digestion of WAS and
OFMSW for bio-hythane production
The first pilot scale S.C.E.N.A. (Short-Cut Enhanced Nutrients Abatement) system
was applied and validated for the treatment of the supernatant of anaerobic co-digestion
of sewage sludge and organic fraction of the municipal solid waste (OFMSW) (Fatone
et al. 2011) within the pilot hall of the Treviso (northern Italy) municipal treatment
plant. The authors discussed the start-up strategy and carbon source to enhance the
short-cut nitrogen removal and via-nitrite enhanced biological phosphorus uptake
from anaerobic supernatant (Frison et al. 2012, 2013). The first integrated scheme
of (1) two-phase anaerobic digestion for the bio-hythane production (Cavinato et al.
2013) and (2) via-nitrite biological nutrients removal was proposed by Frison et al.
(2013) (Figure 16.1, adapted from Malamis et al. 2013).
In this first scheme the hydrolysis reactor (dark fermentation) of the two-
phase anaerobic digestion is used to provide short chain volatile fatty acids to
the anoxic phase in the short-cut sequencing batch reactor (scSBR), which is
treating the anaerobic supernatant. Starting from the conventional activated
sludge inoculum, the start-up of the scSBR is carried out in two periods (Frison
et al. 2012) and the stable via-nitrite route is achieved in 15–30 days. Then,
nitritation-denitritation and significant phosphorus luxury anoxic uptake was
observed. However, the dark acid fermentation did not optimize the contents of
propionic and butyric acid that can enhance the via-nitrite (anoxic) biological
phosphorus uptake.
On the basis of pilot scale trials, Frison et al. (2013) calculated the specific
costs of via-nitrite nitrogen removal (Table 16.4). It was found that using the
OFMSW fermentation liquid instead of methanol, the overall specific cost for
nitrogen removal in the nitritation-denitritation decreased by 22%. In addition,
the enhanced phosphorus biological removal was an important added value of
the scSBR. Moreover, the added value of the contemporary via-nitrite anoxic
phosphorus uptake was not considered, thus underestimating the advantages of the
S.C.E.N.A. process with comparison to the complete autotrophic nitrogen removal
which must be followed by struvite recovery to achieve the same nutrients removal
from anaerobic supernatant.
(Tong et al. 2007; Zheng et al. 2010). Higher phosphorus removal efficiency was
achieved with the use of SCVFA derived from WAS compared to acetate (Tong et al.
2007). The authors explained that the presence of propionate was probably the reason
for better phosphorus removal, while the higher nitrogen removal efficiency might
be due to the better use of exogenous denitrification pathway for nitrogen removal.
In the sludge fermentation process, the pH plays an important role on the hydrolysis
of sludge and the production of SCVFAs from excess sludge in fermentation. Under
alkaline conditions, the yield of SCVFAs can be significantly enhanced (Yuan &
Weng, 2006). Recent studies have demonstrated enhanced SCVFAs production
and inhibition of methanogenic activity (resulting in less SCVFAs consumption)
under alkaline conditions (Wu et al. 2010). NaOH and Ca(OH)2 are widely used
for alkaline sludge treatment; the type of chemical that is used impacts on waste
activated sludge (WAS) hydrolysis, acidification and dewatering ability (Kim et al.
2003). Thus, the type of reagent that is used for the pH adjustment in alkaline
fermentation influences the effectiveness of the process. The optimum pH range
9–11 was reached using NaOH and Ca(OH)2 (Su et al. 2013). This technique is
not economically and environmentally sustainable and enhances the salinity of the
carbon source, thus decreasing the rates of nitritation/nitrification. Furthermore, the
sludge dewatering characteristics and the separation of the produced fermentation
liquid from sludge can be adversely affected from the use of NaOH (Su et al. 2013;
Longo et al. 2014). Recent studies have shown that the use of WAS fermented liquid
as carbon source results in the reduction of nitrous oxide (N2O) and nitric oxide
(NO) production during the via nitrite processes (Zhu & Chen, 2011).
Therefore, the initial S.C.E.N.A. process was upgraded for application in a
conventional municipal wastewater treatment plant and applied in the Carbonera
WWTP.
The Carbonera (Veneto Region – Italy) WWTP plant has actual treatment
potential of approximately 40,000 PE. The full scale wastewater treatment line
is composed of the following operation units: screening and degritting, primary
sedimentation, activated sludge process (+chemical P precipitation) and secondary
clarifier; anaerobic digestion of the sewage sludge and dewatering (by centrifuge).
Currently, an S.C.E.N.A. pilot system is operating for the separate treatment of
part of the reject water of the plant. The pilot unit consists of three main subunits:
the alkaline fermentation unit, the membrane unit for the solid/liquid separation
of the fermentation effluent and the sequencing batch reactor (SBR) for the via
nitrite nutrient removal. The pilot scale fermentation unit (reaction volume 0.5 m3)
receives mixed (primary and secondary sewage sludge) from the full scale WWTP
plant. The sewage sludge is fermented to produce an effluent that is rich is SCVFA.
An ultrafiltration (UF) membrane filtration skid is employed for the solid/liquid
separation of the fermentation effluent (MO P13U 1 m, Berghof, Germany),
while alternative less energy-intensive microscreens are under investigation and
validation at pilot scale. Fermented sludge is first screened through 50 mm to
prevent clogging of the membrane modules. The sludge fermentation liquid is then
directed to the short-cut via nitrite SBR process (3 m3) that treats separately the
anaerobic supernatant, removing N and P via nitrite.
According to the pilot-scale results, the best parameters for the alkaline
fermentation of sewage sludge, the solid-liquid separation and the via-nitrite
nutrients removal were found out (Longo et al. 2015). The production of SCVFA
by alkaline fermentation has proved to be highly dependent on pH and temperature.
The use of wollastonite was tested in order to avoid the addition of chemicals in the
alkaline fermentation process. The fermentation liquid consisted mainly of acetic,
propionic and butyric acid (37, 34 and 15% respectively). Under the presence of
acids, the following silicate reaction can occur
Table 16.7 Estimation of annual cost for reject water treatment in Carbonera WWTP.
C.A.S.P.* S.C.E.N.A.
ELECTICAL ENERGY €/year 28,000 13,300
SLUDGE DISPOSAL €/year 26,800 19,300
PolyAlluminiumChloride €/year 4100 –
WOLLASTONITE €/year – 800
PERSONNEL ANNUAL COST €/year 700 1700
*Current conventional activated sludge process.
As far as the environmental impact is considered, the S.C.E.N.A. system allows the
biological uptake of the phosphorus in a form that can be recovered after composting
of the S.C.E.N.A. sludge. The impact of the S.C.E.N.A. system was also preliminary
evaluated in terms of quality of the secondary effluent as shown in Table 16.8.
Table 16.8 Quality of the secondary effluent in current and simulated future
scenario.
16.5 Conclusions
Sludge reject water is a nutrient-rich flux which should be properly treated managed
for the technical, economical and environmental optimization of the nitrogen
removal and phosphorus recovery in WWTPs. The completely autotrophic
nitrogen removal is the most attractive biological process for the treatment of
sludge reject waters in municipal WWTPs with several full scale applications. This
solution cannot enhance the phosphorus bioaccumulation and should be followed
by struvite crystallization for sustainable phosphorus recovery.
On the other hand, efforts for the technical, economical and environmental
sustainability of wastewater treatment plants should also address the novel
denitrifying biological phosphorus removal via nitrite which offers the potential to
integrate phosphorus and nitrogen removal in a robust single bioreactor in which
ammonium is oxidized to nitrite under aerobic conditions, while under anoxic
conditions the denitrification via nitrite and enhanced biological phosphorus uptake
occur simultaneously by the denitrifying phosphorus accumulating organisms.
Thus, the phosphorus could be recovered via the composted sludge.
The S.C.E.N.A. system achieved these objectives in the wastewater treatment
plant of Carbonera (Veneto region, northern Italy). Here the real anaerobic
supernatant was treated in a nitritation-denitritation short-cut SBR, where the best
available carbon source to enhance the phosphorus bioaccumulation was in-situ
recovered from the sewage sludge.
The OPEX estimation proved the economic viability of the system which led
the water utility Alto Trevigiano Servizi srl to apply the S.C.E.N.A. system by
retrofitting an existing tank. This will also minimize the CAPEX and demonstrate
how this system can be applied to integrate existing WWTPs.
16.6 References
Anthonisen A. C., Loehr R. C., Prakasam T. B. and Srinath E. G. (1976). Inhibition of
nitrification by ammonia and nitrous acid. Water Pollution Control Federation, 48,
835–852.
Battistoni P., Paci B., Fatone F. and Pavan P. (2005). Phosphorus removal from supernatants
at low concentration using packed and fluidized-bed reactors. Industrial & Engineering
Chemistry Research, 44, 6701–7.
Beier M. and Schneider Y. (2008). Abschlussbericht Entwickelung von Bilanzmodellen
für die Prozesse Deammonifikation und Nitritation zur Abbildung gross-and
halb-technischer Anlagen (Final report development of balance models for
deammonification and nitritation processes to illustrate full and half technical scale
installations). Leibniz University Hannover, Hannover: 39.
Blackburne R., Zhiguo Y. and Keller J. (2008). Partial nitrification to nitrite using low dissolved
oxygen concentration as the main selection factor. Biodegradation, 19(2), 303–312.
Carvalho C., Lemos P. C., Oehmen A. and Reis M. A. M. (2007). Denitrifying phosphorus
removal: linking the process the process performance with the microbial community
structure. Water Research, 41, 4383–4396.
Cavinato C., Bolzonella D., Pavan P., Fatone F. and Cecchi F. (2013). Mesophilic and
thermophilic anaerobic co-digestion of waste activated sludge and source sorted
biowaste in pilot-and full-scale reactors. Renewable Energy, 55, 260–265.
Cervantes F. J. (2009). Environmental Technologies to Treat Nitrogen Pollution. IWA
Publishing, London.
Dockhorn T. (2009). About the Economy of Phosphorus Recovery. International Conference
on Nutrient Recovery from Wastewater Streams.
Elser J. and Bennett E. (2011). Phosphorus cycle: a broken biogeochemical cycle. Nature,
478, 29–31.
Fatone F., Dante M., Nota E., Di Fabio S., Frison N., Pavan P. and Fatone F. (2011). Biological
short-cut nitrogen removal from anaerobic digestate in a demonstration sequencing
batch reactor. Chemical Engineering Transactions, 24, 1135–1140.
Frison N., Lampis S., Bolzonella D., Pavan P. and Fatone F. (2012). Two-stage start-up
to achieve the stable via-nitrite pathway in a demonstration SBR for anaerobic
codigestate treatment. Industrial & Engineering Chemistry Research, 51(47),
15423–15430.
Frison N., Di Fabio S., Cavinato C., Pavan P. and Fatone F. (2013). Best available carbon
sources to enhance the via-nitrite biological nutrients removal from supernatants of
anaerobic co-digestion. Chemical Engineering Journal, 215, 15–22.
Fux C., Velten S., Carozzi V., Solley D. and Keller J. (2006). Efficient and stable nitritation
and denitritation of ammonium-rich sludge dewatering liquor using an SBR with
continuous loading. Water Research, 40(14), 2765–2775.
Ganigué R., López H., Balaguer M. D. and Colprim J. (2007). Partial ammonium oxidation
to nitrite of high ammonium content urban landfill leachates. Water Research, 41(15),
3317–3326.
Ganigué R., Volcke E. I., Puig S., Balaguer M. D. and Colprim J. (2012). “Impact of influent
characteristics on a partial nitritation SBR treating high nitrogen loaded wastewater.
Bioresource Technology, 111, 62–69.
Gilbert N. (2009). The disappearing nutrient. Nature, 461(8), 716–718.
Gu S. B., Wang S. Y., Yang Q., Yang P. and Peng Y. Z. (2012). Start up partial nitrification at
low temperature with a real-time control strategy based on blower frequency and pH.
Bioresource Technology, 112, 34–41.
Guo J. H., Peng Y. Z., Wang S. Y., Zheng Y. N., Huang H. J. and Ge S. J. (2009). Effective
and robust partial nitrification to nitrite by real-time aeration duration control in an
SBR treating domestic wastewater. Process Biochemistry, 44, 979–985.
Gustavsson D. J. I. (2010). Biological sludge liquor treatment at municipal wastewater
treatment plants – a review, Vatten, 66, 179–192.
Gustavsson D. J. I. and Jansen J. la Cour (2011). Dynamics of nitrogen oxides emission
from a full-scale sludge liquor treatment plant with nitritation. Water Science and
Technology, 63(12), 2838–2845.
Hao X., Wang C., van Loosdrecht M. C. M. and Hu Y. (2013). Looking Beyond Struvite for
P-Recovery. Environmental Science & Technology, 47, 4965–4966.
Hellinga C., Schellen A. A. J. C., 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, 135–142.
Ivanov V., Kuang S., Stabnikov V. and Guo C. (2009). The removal of phosphorus from
reject water in a municipal wastewater treatment plant using iron ore. Journal of
Chemical Technology and Biotechnology, 84, 78–82.
Jardin N., Thöle D. and Wett B. (2006). Treatment of sludge return liquors: experiences
from the operation of full-scale plants. In: Proceedings of the Water Environment
Federation, WEFTEC 2006, Dallas, 21–25 October, 5237–5255.
Joss A., Salzgeber D., Eugster J., König R., Rottermann K., Burger S., Fabijan P., Leumann
S., Mohn J. and Siegrist H. (2009). Full-scale nitrogen removal from digester
liquid with partial nitritation and anammox in one SBR. Environmental Science &
Technology, 43(14), 5301–5306.
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.
Kishida N., Kim J., Tsuneda S. and Sudo R. (2006). Anaerobic/oxic/anoxic granular sludge
process as an effective nutrient removal process utilizing denitrifying polyphosphate-
accumulating organisms. Water Research, 40(12), 2303–2310.
Lee D. S., Che O. J. and Jong M. P. (2001). Biological nitrogen removal with enhanced
phosphate uptake in a sequencing batch reactor using single sludge system. Water
Research, 35(16), 3968–3976.
Li X., Chen H., Hu L. F., Yu L., Chen Y. G. and Gu G. W. (2011). Pilot-scale waste activated
sludge alkaline fermentation, fermentation liquid separation, and application of
fermentation liquid to improve biological nutrient removal. Environmental Science &
Technology, 45, 1834–1839.
Longo S., Katsou E., Malamis S., Frison N., Renzi D. and Fatone F. (2015). Recovery
of volatile fatty acids from fermentation of sewage sludge in municipal wastewater
treatment plants. Bioresource Technology, 175, 436–444.
Malamis S., Katsou E., Di Fabio S., Bolzonella D. and Fatone F. (2013). Biological
nutrients removal from the supernatant originating from the anaerobic digestion
of the organic fraction of municipal solid waste. CRC Critical Reviews in
Biotechnology, 34(3), 244–257.
Morse G. K., Brett S. W., Guy J. A. and Lester J. N. (1998). Review: Phosphorus removal
and recovery technologies. Science of The Total Environment, 212, 69–81.
Mulder A. (2003). The quest for sustainable nitrogen removal technologies. Water Science
and Technology, 48(1), 67–75.
Münch, Elisabeth V. and Keith Barr. (2001). Controlled struvite crystallisation for removing
phosphorus from anaerobic digester side streams. Water Research, 35(1), 151–159.
Oleszkiewicz J. A. and Barnard J. L. (2006). Nutrient removal technology in North America
and the European Union: a review. Water Quality Research Journal of Canada, 41,
449–462.
Peng Y. and Zhu G. (2006). Biological nitrogen removal with nitrification and denitrification
via nitrite pathway. Applied Microbiology and Biotechnology, 73, 15–26.
Peng Y.-Z., Wu C.-Y., Wang R.-D. and Li X.-L. (2011). Denitrifying phosphorus removal
with nitrite by a real-time step feed sequencing batch reactor. Journal of Chemical
Technology and Biotechnology, 86, 541–546.
Pitman A. R. (1999). Management of biological nutrient removal plant sludges – change the
paradigms? Water Research, 33, 1141–1146.
Pollice A., Laera G., Saturno D. and Giordano C. (2008). Effects of sludge retention time
on the performance of a membrane bioreactor treating municipal sewage. Journal of
Membrane Science, 317(1), 65–70.
Seongjun P., Bae W. and Rittmann B. E. (2010). Multi-species nitrifying biofilm model
(MSNBM) including free ammonia and free nitrous acid inhibition and oxygen
limitation. Biotechnology and Bioengineering, 105(6), 1115–1130.
STOWA (2012). Explorative research on innovative nitrogen recovery, Rapport 51.
Su G., Huo M., Yuan Z., Wang S. and Peng Y. (2013). Hydrolysis, acidification and
dewaterability of waste activated sludge under alkaline conditions: combined effects
of NaOH and Ca(OH) 2. Bioresource Technology, 136, 237–243.
Taruya T., Ueno Y. and Fujii M. (2000). Development of phosphorus resource recycling
process from sewage, Poster Paper, First World Water Congress of the IWA, Paris,
3–7 July.
Tong J. and Chen Y. (2007). Enhanced biological phosphorus removal driven by short-chain
fatty acids produced from waste activated sludge alkaline fermentation. Environmental
Science & Technology, 41, 7126–7130.
van Dongen U., Jetten M. S. M. and van Loosdrecht M. C. M. (2001). The Sharon –
Anammox process for treatment of ammonium rich wastewater. Water Science and
Technology, 44(1), 153–160.
Volcke E. I. P., van Loosdrecht M. C. M. and Vanrolleghem P. A. (2007). Interaction
between control and design of a SHARON reactor: economic considerations in a
plant-wide (BSM2) context. Water Science and Technology, 56(9), 117–125.
Wett B. (2007). Development and implementation of a robust deammonification process.
Water Science and Technology, 56(7), 81–88.
Wu H., Gao J., Yang D., Zhou Q. and Liu W. (2010). Alkaline fermentation of primary
sludge for short-chain fatty acids accumulation and mechanism. Chemical Engineering
Journal, 160, 1–7.
Ye L., Tang B., Zhao K. F., Pijuan M. and Peng Y. Z. (2009). Nitrogen removal via nitrite
in domestic wastewater treatment using combined salt inhibition and on-line process
control. Water Science and Technology, 60(6), 1633–1639.
Yuan H., Chen Y., Zhang H., Jiang S., Zhou Q. and Gu G. (2006). Improved bioproduction
of short-chain fatty acids (scfas) from excess sludge under alkaline conditions.
Environmental Science & Technology, 40, 2025–2029.
Zhang S., Peng Y., Wang S., Zheng S. and Gou S. (2007). Organic matter and concentrated
nitrogen removal by shortcut nitrification and denitrification from mature municipal
landfill leachate. Journal of Environmental Sciences, 19, 647–651.
Zheng X., Chen Y. and Liu C. (2010). Waste activated sludge alkaline fermentation
liquid as carbon source for biological nutrients removal in anaerobic followed by
alternating aerobic–anoxic sequneching batch reactors. Chinese Journal of Chemical
Engineering, 18, 478–485.
Zhu X. Y. and Chen Y.G. (2011). Reduction of N2O and NO generation in anaerobic-aerobic
biological wastewater treatment process by using sludge alkaline fermentation liquid.
Environmental Science & Technology, 45, 2137–2145.
17.1 Introduction
Energy is one of the three highest costs for wastewater treatment facilities besides
personnel and sludge disposal (Lazarova et al. 2012; Zessner et al. 2010). Most of
the middle and large-scale wastewater treatment plants (WWTPs) are designed to
leverage some kind of activated sludge process; however, such processes are highly
demanding in terms of energy, since more than 50% of the total energy consumed
in WWTPs is normally used for aeration and mixing of the activated sludge tank
(Lazarova et al. 2012). Since the aeration systems are fundamental to support
the biological activity in activated sludge and, at the same time, the most energy
demanding process, much research and technological innovation has been funded
to improve their efficiency (Amand & Carlsson, 2012; Jeppsson et al. 2013).
WWTPs are designed to manage peak flowrates or at least a large fraction
of the worst-case scenario flow rates (Tchobanoglous et al. 2003). This leads
to the design of oversized tanks if compared to the volumes actually required
to treat the average flowrate, so WWTPs usually work at low loading rates.
The design of oversized tanks (or ‘safety margins’) together with large plant
dynamics could cause large energy waste if plants are not properly managed
(Olsson & Newell, 1999).
Nowadays, an efficient management of any WWTP cannot ignore the support
offered through automation and information technology, in order to reach the
effluent quality standards required by legislation and process sustainability and
to limit the energy costs. To this aim, optimizing oxygen addition through the
aeration systems by using control logic or automatic policies is fundamental;
however, the choice of the most effective control logic or policy strictly depends on
the configuration of the evaluated plant.
This chapter provides an overview of a pilot-scale study carried out to evaluate
the potential energy saving for aeration in wastewater treatment by using sequencing
batch reactors.
Germany and, up to 1999, 138 SBR plants for domestic sewage were built, while
about 50 SBRs were installed for industrial applications. The size of most of the
plants is less than 5,000 p.e., but the number of plants with a capacity of more than
10,000 p.e. is increasing. Among these, the largest SBR plant is located in the City
of Neubranbenburg and serves 140,000 p.e. Nowadays, SBR appears to be the most
appropriate treatment for wastewater management in sensitive coastal tourist areas,
where a good flexibility to seasonal fluctuations in wastewater quantity and quality
is very important (Tasli et al. 2001).
A SBR and a conventional (i.e., continuous-flow) activated sludge (CAS) plant can
be compared. Both the systems are designed to treat easily a wide range of highly
variable influent conditions. When the hydraulic loading increases, the hydraulic
retention time (HRT) in the continuous flow automatically decreases. In SBRs, the
same relationship is applied simply by changing the cycle time, adjusting the time set
for the idle phase. Indeed, this phase thus serves to buffer peak influent loads while
the react phase is kept in the required range (Wilderer et al. 2001). Most WWTPs
include a CAS process, where wastewater is continuously fed to the plants. In these
plants, aeration systems management is normally performed by one or more cascaded
controllers belonging to the relatively simple PID family (Wahab et al. 2009) or
by more advanced systems such fuzzy controllers (Baroni et al. 2006), which are
designed to continuously optimise the aeration of the aerobic tank (Olsson et al. 2005).
Biological nitrogen removal from wastewater is usually achieved via nitrification
and denitrification processes. During the nitrification process, ammonia is
oxidized to nitrate, which is then reduced to nitrogen gas using organic matter as
electron donor during the denitrification process. Therefore, while conventional
activated sludge plants require two tanks operating under different conditions (i.e.,
aerobic and anoxic), in SBRs nitrogen removal is accomplished in the same tank
by simply managing the switch on and off of the blowers. However, if the process
is not monitored, the length of the different phases is usually fixed considering
the time required to process the maximum load. Assuming that the plant works
properly, this conservative choice guarantees that the concentration of pollutants
in the effluent will generally be low, but, at the same time, it causes a significant
waste of energy. In order to save energy and increase the overall performance of
the treatment system, instead, the duration of the phases should not be set to a
fixed-time control strategy based on a worst-case, but, on the contrary, should be
managed according to the actual duration of the reactions, which, in turn depend
on wastewater load; in particular, aeration should only be maintained on as long as
the time necessary to complete nitrification.
Shaw et al. 2009). Moreover, SBR management should also be automated in order
to allow the plant to work continuously and autonomously (Marsili-Libelli et al.
2008; Spagni et al. 2008). Recently, the automation of SBR processes has been
performed using ‘intelligent control systems’, which rely on AI-based tools, such as
neural networks (Aguado et al. 2009; Luccarini et al. 2002; Luccarini et al. 2010;
Sottara et al. 2007) or fuzzy logic controllers (Marsili-Libelli, 2006; Ruano et al.
2010). These tools have been employed to analyse time series acquired by probes
installed in SBR tanks, trying to detect the end of the process reactions, in order to
optimize the length of the phases.
Recently, much interest has also been shown in the development of remote
control infrastructures, which should include a data acquisition system, a data
storage facility to store the data, a remote control channel to issue commands to
the actuators such as pumps and blowers, and a user interface. Such infrastructures
are desirable since plant management includes activities, such as the diagnosis
of malfunctioning or the regulation of control parameters, where the experience
of expert operators is fundamental. Moreover, Dürrenmatt and Gujer (2012)
investigating on the applicability of various data-driven modelling techniques to
support WWTP management, concluded that a high degree of expert knowledge was
available for long-term operation. All this available knowledge could potentially
be transferred to automated Environmental Decision Support Systems (EDSS) and
integrated with remote control infrastructures for plant running optimisation.
This work shows the potential advantages of the application of control systems
to wastewater treatment. Specifically, this study compares the amount of energy
effectively consumed with that theoretically required by aeration of a pilot-scale
SBR fed with real municipal wastewater. The former was estimated as the total
consumed energy by the blower considering a fixed length of the oxidation phase
according to the worst-case scenario. The latter, instead, was estimated as the
required energy considering a variable length of the aerobic phase, according
to the real duration of the nitrification process. The analysis has been executed
off-line, using the data collected in the course of seven days (28 complete cycles)
of uninterrupted running of the pilot plant. We focused, especially, on the
monitoring of the state of the nitrification process, manageable and controllable
by a knowledge-based EDSS. This infrastructure, presented as case-study able to
control the pilot-scale SBR plant in Sottara et al. (2014), allowed the application
of control policies, recognizing some anomalous operating conditions in order to
ensure a safe operation of the plant.
designed for 2,000 p.e. The raw wastewater was collected after an initial screening
treatment.
The reactor had a working volume of 500 L and was fed with 150 L of screened
municipal wastewater 4 times per day. The plant has been equipped with a
mechanical mixer, a blower connected to a membrane diffuser set to provide a
dissolved oxygen (DO) concentration during the nitrification of approximately
1 mgO2/L, two peristaltic pumps for influent loading and effluent discharge (flow
rate = 6 L/min) and a pump for sludge wastage (flow rate = 1 L/min). Besides, it had
been equipped with a digital modular multi-parameter system for the measurement
and the on-line acquisition of pH, oxidation-reduction potential (ORP), DO and
temperature. All signals were acquired in current (4–20 mA) as analog inputs by
a multi-function data acquisition device (National Instruments 6052E), while the
electrical components were actuated by a PLC (Omron Sysmac CJ series). The
equipment was located on a car trailer.
Screened
wastewater
17.2.2 Process monitoring
Preliminary experiments were performed without the use of EDSS, hardcoding a
fixed-phase duration policy, performing four 6-hour cycles per day. During the fill
phase, 150 L of sewage was loaded in the reactor. The anoxic phase lasted 90 min
while the aerobic phase was ensured by the activation of a blower for 3 hours. A
30-minute settling phase was necessary before discharging the effluent during the
draw phase. Sludge waste took place at the end of the aerobic phase to manage the
solids retention time (SRT) at approximately 20 d. The plant was operated under
similar conditions for a 6 month period, during which the influent, the effluent and
the content of the tank were sampled approximately three times per week. Weekly
track studies were also carried out to monitor the trend of the biological processes
within a complete cycle.
The most relevant values of wastewater composition, measured during the
experimental period, were: total COD of 250–400 mg/L, soluble COD of 200–
350 mg/L, BOD5/COD (0.4–0.7), pH of 8.0–8.2, NH4+ -N of 45–75 mgN/L, and
total Kjeldahl nitrogen (TKN) of 55–85 mgN/L. During the whole experimental
runs, the plant usually showed good removal efficiency, always higher than 80%,
both for ammonia and organic compounds. Concentrations of ammonia in the
effluent of a few mgN/L were occasionally observed, having been caused by
unusually high nitrogen loads occurring sporadically. Moreover, the fairly low
COD/TKN ratio and the variable characteristics of the treated wastewater caused
some variability of the nitrogen-oxidized forms in the effluent. The plant also
showed good settling characteristics and the total suspended solids (TSS) content
in the effluent usually remained below 50 mg/L. Figure 17.2 shows an example of
continuous on-line acquisition of pH, ORP and DO signals, which display both
high variability of the influent and the repetitiveness of the processes. Figure
17.3 confirms that it is possible to observe a relationship between the indirect but
easily measurable signals (such as pH, ORP and DO) and the biological processes
in SBR plants (Peng et al. 2004; Spagni et al. 2001, 2007). Figure 17.3 shows an
example of a cycle, where denitrification and nitrification processes take place as
expected with the relative pH, ORP and DO trends.
8.2
8.0
7.8
pH
7.6
7.4
200
100
0
mV
-100
-200
-300
6
mgO2/L
0
0 10 20 30 40
Time (h)
12
10
NH4+
mgN/L
8 NO3-
6 NO2-
0
8.0 200 6
pH
7.9 ORP
100 5
DO
7.8
0 4
mgO2/L
mV
7.7
pH
-100 3
7.6
-200 2
7.5
7.4 -300 1
7.3 -400 0
0 50 100 150 200 250
Time (min)
Figure 17.3 Example of a SBR cycle: nitrogen forms above, pH, ORP and DO below.
Source: (Luccarini et al. 2010).
17.2.3 EDSS architecture
An effective optimization of the treatment process requires constant monitoring,
in order to adapt the configuration of the plant actuators to the actual state of the
process. Instrumentation, control and automation (ICA) is a reasonably cost-effective
solution to enable the continuous management of one or more plants, even when
the scale or the location of the plant would not justify the employment of human
personnel (Olsson & Newell, 1999). The control policies, however, are context-
dependent and involve some non-trivial decisions. The decision to switch from one
operational phase to a different one – typically the one that follows in the operating
cycle – is generally based on a trade-off between the attempt to ensure the correct
completion of the current process phase and the desire to minimize its duration.
However, an optimal choice depends on the complete knowledge of the state of the
process, which to this date is not (economically) feasible. Since only an indirect
estimate is usually available, any decision should take into account the inherent
uncertainty about the real operating conditions. More generally, the automated
controller should also be able to detect, handle and possibly recover from
failures both in the process and the plant instrumentation, so that their impact
is minimized. This complex scenario suggests the adoption of an automated
Environmental Decision Support System (EDSS) (Poch et al. 2004): the
EDSS will constantly monitor the plant’s state, estimate the current operating
conditions and apply the appropriate policies to optimize the cost/effectiveness
ratio of the treatment process, actively involving the operators only when the
circumstances mandate it.
The EDSS we propose is based on the architecture shown in Figure 17.4. It is
based on a hybrid Service Oriented (SOA) and Event-Driven (EDA) Architecture.
Rather than building a dedicated, monolithic software, we have partitioned the
common functionalities of the EDSS into self-contained modules with different
responsibilities, including (but not necessarily limited to):
• A data acquisition service, that collects and pre-processes the signals
acquired in real time from the probes installed on the plant;
• A control interface, which commands the plant actuators (pumps, blowers,
stirrers, etc.) in order to enforce the desired operational conditions, according
to the policy recommendations;
• A data store, where historical time series are persisted for short- and
long-term usage;
• A user interaction module, to enable (remote) communication between the
EDSS and the plant operators;
• One or more modules implementing the decision support and decision
making policies, as described later in this section;
• Modules dedicated to functionalities supporting the architecture itself,
such as enabling security and authentication, or a registry of the available
modules.
The modules are exposed as services: the core of the architecture is based
on an Enterprise Service Bus (ESB) that enables the communication between
servers and clients. The SOA, then, allows to implement more complex, higher
level functionalities by orchestration of the different services. SOAs, however, are
inherently request-driven: any process must be explicitly initiated by some client.
An EDSS with monitoring and control responsibilities should also be reactive, to
respond promptly to external stimuli such as the changes observed in the plant.
To combine the best of both approaches, (Luckham, 2007), we have created a
hybrid architecture where external events can trigger the execution of one or
more services. During their execution, services can generate additional events
which, in turn, might trigger additional computations, effectively creating virtual
Event Processing Networks (Luckham, 2001). The events themselves are filtered
and routed using a dynamic content-based routing service (Enterprise Patterns),
ensuring that events are delivered only to those subsystems for which they are
relevant. More details on the architecture can be found in (Sottara et al. 2009a).
The core modules of the EDSS implement the analysis, decision and control
logic: in their development, we have adopted a model-driven, knowledge-based
approach, based on hybrid artificial intelligence (AI) technologies. A combination of
predictive models such as neural networks, rules, ontologies and business processes
has been used to model and execute the management and optimization policies.
When compared to ‘ad-hoc’ software, such an approach is considered more robust
and flexible from several perspectives. As in a model-driven architecture, the
separation of the model from the runtime platform allows the two components
to be developed independently, to the point that the latter is usually considered
commodity, while the real value lies in the former. The more declarative nature of
the knowledge base facilitates the interaction with stakeholders and subject matter
experts and improves maintainability.
The complexity of the application justifies the hybrid combination of modelling
techniques. Quantitative approaches such as neural networks are more suitable for
signal analysis and prediction, while declarative, logic-oriented techniques are more
appropriate to define the business logic and the operational policies to manage the
plant. In our EDSS, we have deployed an ensemble of different predictors to estimate
the process state (and thus the progress in each phase) using the indirect signals
17.3 Results
The energy consumption, related to the oxygen used in the aerobic phase, was
investigated through an off-line analysis and the data collected by the data
acquisition system of the plant in May 2008 during seven continuous days.
The real duration of each nitrification phase (which determines the minimum,
necessary duration of the aerobic phase) was estimated considering ten minutes
after the identification of the ‘ammonia valley’ in the pH (point A in Figure 17.5),
the ‘ammonia break-point’ in the ORP (point B in Figure 17.5) or the ‘ammonia
knee’ in the DO time series (point C in Figure 17.5) (Al-Ghusain et al. 1994;
Spagni et al. 2001). Figure 17.5 presents an example where all these events
(ammonia valley, break-point and valley) are clearly visible; moreover, the oxygen
added during the aerobic phase which was not used for biological processes is
highlighted in the same figure (Figure 17.5). Figure 17.5 confirms a significant,
unnecessary energy consumption in the aerobic phase; in fact, we can easily
evaluate in the specific example that the effective duration of the nitrification
process is approximately 40% of the overall time the plant has been operating in
the aerobic phase. However, due to the variability of the influent wastewater, this
time changes continuously and, therefore, the relative length of the aerobic phase
should be adjusted accordingly on a case-by-case basis.
Using the signal profiles of the pilot-plant, the results were divided as follows:
17.3.1 Nitrification time
The effective duration of the nitrification process evaluated in two different days
of the first week of May, Friday and Sunday, is reported in Table 17.1. On Friday,
the required time to complete the nitrification in the cycles at 4:30, 16:30 and 22:30
was 30% of the fixed time, while at 10:30, the time was longer (55% of fixed time)
than others, which was expected since it corresponds to the maximum load of
the day. On Sunday, the required time to complete the nitrification process was
always longer than Friday, being 70% for the cycle at 10:30 and approximately
40% for the others. This longer time is related to life style of the habitants of
the specific residential area: in fact, the sampled wastewater was collected from a
sewer which serves a typical residential area, whose inhabitants move to the City of
Bologna for working activities during the week. It is noteworthy that the effective
nitrification time was, in average, approximately 40% of the total aerobic phase,
confirming that the length of that phase (chosen as a precautionary value) could be
significantly reduced when the plant is not fed with a peak load, and even more in
case of a diluted load (e.g., during rain or storm events). Even with the maximum
observed load during the weeks reported in this study (i.e., the cycle on Sunday at
10.30) the time required to complete the nitrification was approximately 150 min,
while the total aerobic phase duration was set at 195 min. However, as observed
during the 6 month study, several cycles presented aerobic phases actually lasting
even longer than the worst-case pre-fixed duration (195 min), confirming the need
of an automatic control system for the proper and reliable management of the SBR.
Table 17.2 Saveable oxygen for the aerobic phase (unit as percentage).
17.3.3 Cost analysis
The electricity consumption shows large variation according to the applied processes
and the size of the plant. The cost analysis, thus, is evaluated on the basis of a small
(about 2,000 p.e.) full scale municipal WWTP. The energy consumption can be
estimated for small plants to be between 30 and 50 kWh per p.e. per year (a) (Becker
& Hansen, 2013; Kampschreur & van Loosdrecht, 2012).
According to the data reported in Table 17.2, approximately 50–55% of energy
related to aeration could be saved. Therefore, for the hypothetical full-scale plant, a
saving of 30,000 to 50,000 kWh/a corresponding to approximately 4,500 to 7,500
E/a at an electricity cost for aeration of 0.15 E/kWh (assumed as an average value
in Italy) could be achieved. Similarly, assuming a wastewater production of 200 L/d
per p.e., a saving equivalent to 0.03–0.05 E/m3 could be achieved.
Reaching an equivalent energy savings with a CAS plant is feasible, but the
required equipment is higher and much more expansive. First, a PI controller to
maintain DO concentration to a fixed set-point is essential and, consequently, an
inverter to regulate the air flow insufflated by the compressor in the oxidation tank.
Second, since the oxygen consumption depends on many factors, the best value of
the set-point is variable, but it can be estimated from ammonium measurements.
Today, this is a proven technology and the energy savings by DO control can be
significant. In effect, controlling the DO concentration to a constant set-point can
save 30–50%, while to base the set-point on ammonium measurement can save
another 10–15% (Olsson, 2012). Nevertheless, the investment costs to equip a plant
with this technology are probably too high for small and medium plants, for which
the SBR solution appears to be more sustainable, at least for plants which serve up
to 20,000 p.e.
17.4 Conclusions
The study confirms that small residential communities produce wastewater with
highly variable concentrations and flow rates, which could greatly affect the
WWTP processes.
Moreover, a continuous monitoring of the processes seems to be essential to
achieve an optimal cost/performance ratio, especially for those cases where high
influent variability occurs.
For this kind of very variable sewers, SBRs may be an effective and economic
solution; SBRs also seem to facilitate the monitoring of the processes.
For discontinuous processes, the use of cheap and reliable sensors, such as pH,
ORP and DO allows to observe the organic matter and nitrogen removal processes.
In particular, this study demonstrates that the aerobic phase in SBRs can be easily
and effectively monitored and controlled by using proper tools of analysis and
management policies. In fact, the results show that an energy saving up to 55%
can be achieved on the aeration energy consumption, when compared to a static
management, based on safety parameters used at design time.
17.5 Acknowledgements
The authors thank HERA S.p.A for having partly supported the activities reported
in this chapter.
17.6 References
Aguado D., Ribes J., Montoya T., Ferrer J. and Seco A. (2009). A methodology for
sequencing batch reactor identification with artificial neural networks: A case study.
Computers & Chemical Engineering, 33, 465–472.
Al-Ghusain I. A., Huang J., Hao O. J. and Lim B. S. (1994). Using pH as a real-time control
parameter for wastewater treatment and sludge digestion processes. Water Science and
Technology, 30, 159–168.
Amand L. and Carlsson B. (2012). Optimal aeration control in a nitrifying activated sludge
process. Water Research, 46, 2101–2110.
Baroni P., Bertanza G., Collivignarelli C. and Zambarda V. (2006). Process improvement
and energy saving in a full scale wastewater treatment plant: air supply regulation by
a fuzzy logic system. Environmental Technology, 27, 733–46.
Becker M. and Hansen J. (2013). Is the energy-independence already state-of-the-art at
NW-European wastewater treatment plants? In: Proceedings of Asset Management
for Enhancing Energy Efficiency in Water and Wastewater Systems. Marbella, Spain,
24–26 April.
Degremont (2007). Water Treatment Handbook. 7th edn, Degremont, France.
Dürrenmatt D. J. and Gujer W. (2012). Data-driven modeling approaches to support
wastewater treatment plant operation. Environmental Modelling and Software, 30, 47–56.
Jeppsson U., Alex J., Batstone D. J., Benedetti L., Comas J., Copp J. B., Corominas L.,
Flores-Alsina X., Gernaey K. V., Nopens I., Pons M.-N., Rodríguez-Roda I., Rosen
C., Steyer J.-P., Vanrolleghem P. A., Volcke E. I. P. and Vrecko D. (2013). Benchmark
simulation models, quo vadis? Water Science and Technology, 68, 1–15.
Kampschreur M. J. and van Loosdrecht M. C. M. (2012). Finding the balance between
greenhouse gas emission and energy efficiency of wastewater treatment. In: Water-
energy Interactions in Water Reuse, V. Lazarova, K. H. Choo and P. Cornel (eds), IWA
Publishing, London.
Kirkwood S. (2004). Yannawa wastewater treatment plant (Bangkok, Thailand): design,
construction and operation. Water Science and Technology, 50(10), 221–228.
Lazarova V., Peregrina C. and Dauthuille P. (2012). Toward energy self-sufficiency of
wastewater treatment. In: Water-Energy Interactions in Water Reuse, Lazarova V.,
Choo K. H. and Cornel P. (eds), IWA Publishing, London.
Luccarini L., Porrà E., Spagni A., Ratini P., Grilli S., Longhi S. and Bortone G. (2002). Soft
sensors for control of nitrogen and phosphorus removal from wastewaters by neural
networks. Water Science and Technology, 45, 101–107.
Luccarini L., Bragadin G. L., Colombini G., Mancini M., Mello P., Montali M. and Sottara
D. (2010). Formal verification of wastewater treatment processes using events detected
from continuous signals by means of artificial neural networks. case study: SBR plant.
Environmental Modelling and Software, 25, 648–660.
Luckham D. The Power of Events – An Introduction to Complex Event Processing in
Distributed Enterprise Systems, Addison-Wesley Longman Publishing Co., Inc.
Boston, MA, USA © 2001, ISBN: 0-201-72789-7.
Luckham D. (2007). SOA, EDA, BPM and CEP are all Complementary. http://www.
complexevents.com/wp-content/uploads/2007/05/SOA_EDA_Part_1.pdf; http://com
plexevents.com/wp-content/uploads/2007/07/Soa_EDA_Part2.pdf
Marsili-Libelli S. (2006). Control of SBR switching by fuzzy pattern recognition. Water
Research, 40, 1095–1107.
Marsili-Libelli S., Spagni A. and Susini R. (2008). Intelligent monitoring system for long-
term control of sequencing batch reactors. Water Science and Technology, 57, 431–438.
Olsson G. (2012). Water and Energy. Threats and Opportunities. IWA Publishing, London.
Olsson G. and Newell B. (1999). Wastewater Treatment Systems: Modelling, Diagnosis and
Control. IWA Publishing, London.
Pat A. M., Vargas A. and Buitron G. (2011). Practical automatic control of a sequencing
batch reactor for toxic wastewater treatment. Water Science and Technology, 63,
782–788.
Peng Y. Z., Chen Y., Peng C. Y., Liu M., Wang S. Y., Song X. Q. and Cui Y. W. (2004).
Nitrite accumulation by aeration controlled in sequencing batch reactors treating
domestic wastewater. Water Science and Technology, 50, 35–43.
Poch M., Comas J., Rodriguez-Roda I., Sanchez-Marrè M. and Cortes U. 2004. Designing
and building real environmental decision support systems. Environmental Modelling
and Software, 19(9), 857–873.
Ruano M. V., Ribes J., Sin G., Seco A. and Ferrer J. (2010). A systematic approach for
fine-tuning of fuzzy controllers applied to WWTPs. Environmental Modelling and
Software, 25, 670–676.
Tasli R., Artan N. and Orhon D. (2001). Retrofitting SBR systems to nutrient removal in
sensitive tourist areas. Water Science and Technology, 44(1), 121–128.
Shaw A., Watt J., Fairey A. W. and Iler M. (2009). Intelligent sequencing batch reactor
control from theory, through modelling, to full-scale application. Water Science and
Technology, 59, 167–173.
Sottara D. (2010). Integration of Symbolic and Connectionist AI Techniques in the
Development of Decision Support Systems Applied to Biochemical Processes. PhD
dissertation, University of Bologna.
Sottara D., Luccarini L. and Mello P. (2007). AI techniques for waste water treatment plant
control case study: denitrification in a pilot-scale SBR. Lecture Notes in Computer
Science, 4692, 639–646.
Sottara D., Manservisi A., Mello P., Colombini G. and Luccarini L. (2009a). A CEP-
based SOA for the management of waste water treatment plants. IEEE Workshop on
Environmental, Energy, and Structural Monitoring Systems, 2009. EESMS 2009, doi:
10.1109/EESMS.2009.5341314.
Sottara D., Colombini G., Luccarini L. and Mello P. (2009b). A pool of experts to evaluate
the evolution of biological processes in SBR plants. Lecture Notes in Computer
Science (including subseries Lecture Notes in Artificial Intelligence and Lecture
Notes in Bioinformatics) 5572 LNAI, pp. 368–375.
Sottara D., Mello P., Morlini G., Malaguti F. and Luccarini L. (2012a). Modelling SBR
Cycle Management and Optimization Using Events and Workflows. Proceedings of
6th International Congress on Environmental Modelling and Software (iEMSs) 1–5
July, Leipzig, Germany.
Sottara D., Mello P., Bragaglia S., Pulcini D., Giunchi D. and Luccarini L. (2012b).
Ontologies, Rules, Workflow and Predictive Models: Knowledge Assets for an IEDSS.
Proceedings of 6th International Congress on Environmental Modelling and Software
(iEMSs), 1–5 July, Leipzig, Germany.
Sottara D., Bragaglia S., Pulcini D., Mello P. and Luccarini L. (2014). A hybrid, integrated
IEDDS for the Management of Sequencing Batch Reactors. In: Proceedings of the
7th International Congress on Environmental Modelling and Software, D. P. Ames,
N. W. T. Quinn and A. E. Rizzoli (eds), June 15–19, San Diego, California, USA.
ISBN: 978-88-9035-744-2.
Spagni A., Buday J., Ratini P. and Bortone G. (2001). Experimental considerations on
monitoring ORP, pH, conductivity and dissolved oxygen in nitrogen and phosphorus
biological removal processes. Water Science and Technology, 43, 197–204.
Spagni A., Lavagnolo M. C., Scarpa C., Vendrame P., Rizzo A. and Luccarini L. (2007).
Nitrogen removal optimization in a sequencing batch reactor treating sanitary landfill
leachate. J. Environ. Sci. Heal. A., 42, 757–765.
Spagni A., Marsili-Libelli S. and Lavagnolo M. C. (2008). Optimisation on sanitary landfill
leachate treatment in a sequencing batch reactor. Water Science and Technology, 58,
337–343.
Tchobanoglous G., Burton F. L. and Stensel H. D. (2003). Wastewater Engineering:
Treatment and Reuse. Metcalf and Eddy, Mc Grw Hill, Boston.
Wahab N. A., Katebi R. and Balderud J. (2009). Multivariable PID control design for
activated sludge process with nitrification and denitrification. Biochemical Engineering
Journal, 45, 239–248.
Wilderer P. A., Irvine R. L. and Goronszy M. C. (2001). Sequencing Batch Reactor
Technology. Scientific and Technical Reports N. 10. IWA Publishing, London.
Zessner M., Lampert C., Kroiss H. and Lindtner S (2010). Cost comparison of wastewater
treatment in Danubian countries. Water Science and Technology, 62, 223–230.
E-mail: rashid.elmi@vav.oslo.kommune.no
2Researcher, Department of Hydraulic and Environmental Engineering,
18.1 Introduction
Mitigation of global warming is a top priority in the energy and environmental
policies of the city of Oslo. The city of Oslo targets a 50% reduction in greenhouse
emissions by year-2030 compared to 1991 levels (Byrådsavdeling for klima og
samferdsel, Oslo Kommune (2013)). One of the sectors being focused on, is public
transport. Meeting this ambitious target the city’s administration has set for itself,
entails the quick adoption of innovative approaches to optimisation of energy
consumption as well as alternate modes of renewable energy generation. Oslo has
stepped up to this challenge by effecting a switch from diesel to renewable energy
in its public transport system, since the beginning of year-2010. A wonderful
synergistic relationship has been uncovered in the process between one public
service (transport) and another (sanitation). The Bekkelaget wastewater treatment
plant (WWTP) in Oslo – more about which will be discussed later in this chapter –
in addition to performing its primary function of treating wastewater and ensuring
that the nutrient loading of the receiving water body (the Oslo fjord) is curtailed,
has doubled up as a site for renewable energy production.
Anaerobic digestion of sewage sludge is the oldest and the most-prevalent method
for decomposing and stabilizing organic matter in an oxygen-free environment.
In most cases, the essential and often the only purpose of digesting sewage sludge
anaerobically has been (and still is) the partial sterilization and mass-reduction of the
sludge generated. Biogas is generated in the process. While it can be flared and released
into the atmosphere (as happens in many WWTPs around the world, where sludge is
anaerobically digested), it can also be valued for its energy-content. In the 21st century,
with climate change (read global warming) emerging as a key global concern, biogas
from sludge digesters ought to be looked upon as a source of renewable energy (Speece,
1996; Lemaand Omil, 2001; McCarty, 2001). Biogas in general – and sewage sludge
digester biogas in particular – has been a subject of interest for several researchers
in the fields of energy and water/wastewater engineering, with the water-energy-
carbon nexus gaining increasing importance as a subject of research and analysis
(Venkatesh & Dhakal, 2008; Venkatesh, 2012). Figure 18.1 below (Venkatesh &
Elmi, 2013) is a simplified schematic sketch which illustrates the various ways in
which energy can be recovered (and used) from biogas in WWTPs.
Figure 18.1
Options of utilization of biogas from sewage sludge digesters in
wastewater treatment plants, excluding the possible production of chemicals
(Venkatesh & Elmi, 2013). (The pre-treatment process for biogas cleaning and/or
upgrading have not been included).
also contain traces of hydrogen sulphide, water vapour, oxygen, and siloxanes. The
concentration of each compound in the biogas depends on the type of substrate
digested, and its composition. Upgrading the biogas implies increases the content
of methane and removing carbon dioxide and other impurities from it. A study
from Finland stated that biogas from sewage sludge digestion has the highest
methane content and is least contaminated with benzene, hydrogen sulphide and
nitrogen (Rasi et al. 2007). This Finnish study also concluded that if biogas has to
be utilized as a transportation fuel (in addition to being a possible source of heat
and electricity), preference should be given to sewage sludge or manure given that
both have lower content of nitrogen, halogenated and silicon compounds; and so the
biogas generated therefrom is easier to upgrade. Appels et al. (2008), had put the
annual potential of biogas production in Europe at over 200 billion cubic metres,
and had stated that upgraded biogas would be an important future contributor to
the energy supply of Europe. While noting that biogas-use in Combined Heat and
Power plants in WWTPs was quite well-entrenched, this paper also observed that
its use as transport fuel was becoming more and more common in Europe.
At the time of writing, technologies to upgrade biogas to natural-gas standards
recommended for vehicle fuels, are in vogue. As listed in Zhao et al. (2010),
the main technologies used are water and polyethylene scrubbing (costs put
at 0.13 Euro/Nm3), pressure swing adsorption using molecular sieves (0.4 Euro
per Nm3), chemical absorption (0.17 Euro per Nm3), bio-filter, cryogenic separation
(0.44 Euro per Nm3) and membrane separation (0.17 Euro per Nm3). Ryckebosch
et al. (2011) conclude in their paper that most of the choices are determined by the
presence or absence of suppliers for the technology in the country; Sweden prefers
water scrubbers, Germany has a penchant for pressure swing adsorption, while in
the Netherlands, both these, as well as membrane technology is in vogue. Kaparaju
(2011) has noted that in year-2011, the water scrubber technology was used in 48
biogas upgrading facilities in Europe, physical absorption using organic solvent in
10, chemical absorption using organic solvent in 31, membrane technology in 6,
cryogenic separation in 1 and pressure swing adsorption in 41. The total installed
upgrading capacity in 2011 was approximately 115,155 Nm3 per hour. As far as the
capital investment in biogas upgrading plants are concerned, Urban (2009) has put
it at 1 million Euros on average, for plants treating 500 Nm3 biogas per hour, and
close to 3 million Euros for plants treating 2000 Nm3 biogas per hour; implying
economies of scale.
Bekkelaget WWTP – the WWTP in Oslo, has been using the chemical absorption
technology for upgrading biogas to biomethane, since the beginning of 2010. This
waste water treatment plant was built in 2001 with two thermophilic anaerobic
digesters for sludge treatment. The two digesters, during the period 2001–2010,
produced approximately 20 GWh-equivalent of biogas, annually (with a methane
content of 60–65%). Till 2010, most of the biogas (16.5 GWh) was used onsite
to deliver heat to the sludge digester (12 GWh) and for sludge drying (4.5GWh).
The rest – equivalent to 3.5 GWh – was flared. In year-2007, on the grounds of
handling process. The influent from this stage flows into the denitrification zone in
the activated sludge process. The activated sludge system consists of two zones –
anoxic and aerated. Nitrogen removal occurs in two steps – through alternating
biological processes. Denitrification occurs in the anoxic phase, while nitrification
in the aerobic phase. In this approach, ammonia is oxidised to nitrate which is
then denitrified to N2 gas. Ferrous sulphate is added to the excess recycled
activated sludge for phosphorus removal (simultaneous precipitation). The aerated
wastewater enters the secondary settler (also called the clarifier in Figure 18.2)
wherein the activated sludge flocs are removed, resulting in the outflow of a clear
treated effluent. This secondary sludge – also called biological sludge – is pumped
to the centrifugal thickener. The dual media filter (prior to which polyaluminium
chloride is added) is used as the final polishing step before the final treated effluent
is discharged to the Oslofjord.
The volume of the primary sludge is reduced in a gravity belt thickener
and polymer is added to it for conditioning. The biological sludge undergoes a
thickening process in a centrifugal thickener as mentioned earlier. The thickened
sludge, prior to entry into the two thermophilic anaerobic digesters, needs to be
heated to the temperature of the digesters. Post-digestion, the sludge is dewatered
by centrifuges, before being despatched for use as fertiliser in agricultural farmland
close to Oslo. The other important output from the digesters is, of course, biogas.
At the upgrading plant (Figure 18.3), activated carbon is used to adsorb and
remove hydrogen sulphide, siloxanes and other undesirable constituents of the
biogas, which if present may corrode/abrade gas storage tanks, compressors and
the engines in which the biomethane will eventually be combusted (Ryckebosch
et al. 2011). The biogas upgrading facility at Bekkelaget, uses LP COOAB (Low
Pressure CO2 Absorption by an amine) technology. The amine solution used for
absorption of the carbon dioxide from the biogas, is regenerated by heating. The
heat energy required for the regeneration is provided by a pellet-fuelled boiler; the
heat requirement being 2.6 GWh annually.
Figure 18.3 The biogas upgrading plant at Bekkelaget, Oslo (picture courtesy:
Oslo Water and Wastewater Authority, Norway).
Figure 18.4 Biomethane value chain – from producer to end-user (the increasing
thickness of the green arrows indicates the price increment from left to right, along
the supply chain).
by marinhoeng@hotmail.com
Cost Elements In 2012 Comments
(million
NOK)
Annual capital expenses
Fixed annual 2.53 Linear depreciation assumed over the 15-year lifetime
depreciation, after
discounting to year-2009
Interest payments 0.93 Constant payment of 930,000 NOK annually (at 2.44%)
Operation and maintenance expenses
Electricity 0.599 Consumption of 1 GWh assumed to remain constant till 2024. The tariff rate in Norway
fluctuates a lot, and is highly dependent on a host of external factors – precipitation,
temperature and demand from other countries on the Nordic grid (Eika, 2013). Tariff rate in
nominal NOK is assumed to rise in tandem with the general inflation rate – 2.07% per annum.
by marinhoeng@hotmail.com
kg in 2013. The facility was equipped with activated carbon for operation in 2010. In 2011
however, 1000 kilograms were purchased.
Salaries to employees 0.39 This is assumed to increase from 2013 onwards, at a rate of 3% per year (above the
general inflation rate). In 2010 and 2011, the salaries were 0.21 and 0.35 million NOK
respectively.
General maintenance 0.126 Assumed to be 3.5% of the capital costs in 2010 (Starr, 2013); and thereafter rise
uniformly at a rate of 2% annually. As a plant gets older, maintenance expenses are
bound to increase. This also factors in the cost of phased efficiency improvement of the
pellets-combustor.
Income to upgrading facility
Sale of biomethane 6.57 The energy equivalent of the biomethane sold annually assumed to remain constant at
19.9 GWh from 2013 onwards. If capacity of the wastewater treatment plant increases
and excess biogas is generated, we assume that this would be flared. This is taken into
2010
2011
2012
2013
2014
2015
2016
2017
2018
2019
2020
2021
2022
2023
2024
Year
1.0E+07
7.5E+06
5.0E+06
Value in nominal NOK
2.5E+06
0.0E+00
-2.5E+06
-5.0E+06
-7.5E+06
Depreciaon Interest payments
Amine Acvated carbon
Pellets Electricity
Salaries General maintenance
Figure 18.5 Life-cycle costing of the biogas upgrading facility (from 2009–2024)
(the X-axis labels, which are the calendar years from 2009 to 2024 have been
omitted for the sake of clarity).
Figure 18.5 depicts the cash-flows over time. The X-axis labels have been
omitted for clarity. The capital investments are committed to year-2009, and the
income (positive values) and expenses (negative values) for years 2010 to 2024
follow thereafter. Using a discount rate of 3.1%, the cash flows are discounted back
to year-2009, in order to estimate the net present value (NPV). The NPV at this
discount rate is 13.63 million NOK. While 3.1% is the discount rate adopted by
Oslo VAV in its accounting, the authors have tested for a series of discount rates
ranging from 2% to 10%, in Figure 18.6. As long as the discount rate is less than
approximately 7.7%, the NPV would be positive and the investment economically
feasible.
A sensitivity analysis can be conducted by assuming different unit cost scenarios
for electricity, pellets, amine and activated carbon, various selling price scenarios
for biomethane. Assumptions made about salaries and maintenance expenses, as
Figure 18.6 Effect of different discount rates on the Net Present Value.
The prevalent confidentiality requirement dictated that the exact unit price
paid by the end-user RUTER AS (the last link on the right in Figure 18.4) for the
biomethane could not be disclosed to the authors. However, as per an analysis
done by Xynteo (Oslo, Norway), the price more than triples en route; and the
final selling price of biomethane is almost as much as that of the fossil-diesel
it replaces. Thus, RUTER AS does not really profit economically by switching
over from fossil-diesel to sludge-derived-biomethane; though environmentally,
it succeeds in ‘greening’ its operations. On the one hand, the carbon dioxide
emitted by the combustion of biomethane is biogenic, and on the other, emissions
of other pollutant gases (which would have occurred had diesel been used) are
reduced considerably.
18.5 Conclusion
Whether the investment in the facility is profitable or not would strongly depend
on the assumptions made about the cashflows and the discount rate chosen for the
analysis. As mentioned earlier, the authors have only investigated the effect of
different discount rates for a fixed set of assumptions about the costs and price.
For this given set of assumptions, at the discount rate adopted by Oslo VAV (3.1%),
the investment in the facility turns out to be profitable. At the end of its 15-year
lifetime, though economically, the capital value of the plant would have been
depreciated down to zero, there would be a ‘technical’ salvage value, which along
with the net benefits incurred (as indicated by the positive NPV) over the 2010–
2024 period, could be carried forth into the next spell of operation, with the aid of
a fresh infusion of capital, and possibly an expansion in capacity to meet both the
rising supply of biogas and a likely rise in demand for biomethane.
In addition, the biomethane supply provides the Oslo municipality and the
public transport company RUTER AS with a ‘green profile’ and improves their
status in the eyes of the inhabitants of Oslo city who avail of the sanitation and
transportation services. Environmentally as well, the biogas-to-biomethane-to-
transport-fuel project contributes to the truncation of the carbon footprint of Oslo
city, while expanding its ‘green-print’ a little (refer Venkatesh & Elmi (2013) for
the environmental assessment which is beyond the scope of this chapter).
Diesel consumption by RUTER AS is supplanted by biomethane sourced
from Bekkelaget’s sewage sludge digester biogas, and the bus company benefits
economically in the process (Figure 18.7). In a green economy, after all, it is a
zero-sum game within the energy sector. What is good for the goose (read renewable
energy sub-sector) is not good for the gander (read fossil energy sub-sector).
18.6 References
Appels L., Baeyens J., Degreve J. and Dewil R. (2008). Principles and potential of the
anaerobic digestion of waste-activated sludge. Progress in Energy and Combustion
Science, 34, 755–781.
Byrådsavdeling for klima og samferdsel, Oslo Kommune (Department of Transport
and Environmental Affairs, City of Oslo) (2013). Miljø og Klimarapport 2013
(Environment and Climate Report 2013). http://www.enoketaten.oslo.kommune.no/
oslo_kommunes_miljo_og_klimarapport_2012 (accessed 18 November 2014).
Eika T. (2013). Statistics Norway. Personal Communication with Rashid Elmi (one of the
authors) in November.
Inflation. eu. (2013). Available at www.inflation.eu/inflation-rates/norway/historic-inflation/
cpi-inflation-norway.aspx (accessed 30 October 2013).
Kaparaju P. (2011). 2011: D5.1 Evaluation of Potential Technologies and Operational Scales
Reflecting Market Needs for Low-cost Gas Upgrading Systems. Valorgas. Available as
http://www.valorgas.soton.ac.uk/Deliverables/111129_VALORGAS_241334_D5-1_
Final_version.pdf. (accessed 17 October 2013).
Lema J. M. and Omil E. (2001). Anaerobic treatment: a key technology for a sustainable
management of wastes in Europe. Water Science and Technology, 44, 133–140.
McCarty P. L. (2001). The development of anaerobic treatment and its future. Water Science
and Technology, 44, 149–156.
Rasi S., Veijanen A. and Rintala J. (2007). Trace compounds of biogas from different biogas
production plants. Energy, 32, 1375–1380.
Ryckebosch E., Drouillon M. and Vervaeren H. (2011). Techniques for transformation of
biogas to biomethane. Biomass and Bioenergy, 35(5), 1633–1645.
Speece R. E. (1996). Anaerobic Biotechnology for Industrial Waste Waters. Archae Press,
Nashville, Tennessee, USA.
Starr K. (2013). Environmental and Economic Assessment of Carbon Mineralisation
for Biogas Upgrading. Thesis submitted in September 2013, in Fulfillment of the
Requirements for the PhD Degree in Environmental Sciences and Technology.
Universit at Autonoma de Barcelona, Barcelona, Spain.
Urban W., Girod K. and Lohmann H. (2009). Technologienund Kosten der
BiogasaufbereitungundEinspeisung in dasErdgasnetz. Ergebnisse der Merkterhebung
2007–2008. 123s. Fraunhoefer-InstitutfuerUmwelt-, Sicherheits-und Energietechnick,
Oberhausen (German).
Venkatesh G. (2012). The water-energy-carbon nexus. Asian Water, 28(2), Published by
SHP Media Sdn. Bhd. Kuala Lumpur, Malaysia.
Venkatesh G. and Dhakal S. (2012). An international look at the water-energy nexus.
Journal of American Water Works Association, 104(5), 93–96.
Venkatesh G. and Elmi R. A. (2013). Economic-environmental analysis of handling biogas
from sewage sludge digesters in wastewater treatment plants for energy recovery: case
study of Bekkelaget wastewater treatment plant in Oslo (Norway). Energy, 58(10),
220–235.
Zhao Q., Leonhardt E., MacConnell C., Frear C. and Chen S. Purification technologies
for biogas generated by anaerobic digestion. Chapter 9, Climate Friendly Farming,
CSANR Research Report 2010–001. http://csanr.wsu.edu/publications/researchreports/
CFF%20Report/CSANR2010-001.Ch09.pdf (accessed 10 August 2012).
RS, Brazil
19.1 Introduction
The initial concern of mankind, with respect to hygiene conditions, was to
remove the wastewater from inhabited areas without worrying about the effect of
wastewater on the environment. The problem induced by the wastewater generation
was intensified when large cities started to develop. The first public sewer system
was built in Rome (Cosgrove, 1909). It was a huge channel called Cloaca Maxima
that had the purpose of transporting the waste water away from the people of the
city to the Tiber River. Few other cities had systems for removal of sewage before
the Middle Ages and even before the Industrial Revolution.
The gradual increase of population and population concentration in large cities
made the infrastructure for waste removal inadequate. At the same time, concerns
about the capacity of the receiving bodies to accept increasing amounts of sewage
without being affected grown (Science Channel, 2013). A method to reduce costs
in wastewater treatment and seek better overall performances is to reduce the one
of the highest costs: energy consumption (Marco et al. 2013).
The increase in demand for sewage treatment will naturally lead to the increase
of energy consumption. The increase in energy consumption will eventually
supersede the increase in power supply to consumers. A survey carried out by
Silveira et al. (2010) concluded that the energy consumption in sewage treatment
plants is in the order of 0.050% of total energy consumption in Brazil. This
percentage will be doubled in ten years and quadrupled in twenty years, according
to projections in non worst-case scenarios.
There is a gradual evolution in the methods of treatment, in terms of higher
performance and lower energy consumption. One very interesting aspect would be
to install and use autonomous systems for power generation, such as wind turbines
and photovoltaic modules supported by diesel generator sets, possibly connected
to the grid. This would allow the installation of treatment plants in places far from
urban centers and close to the point of disposing the treated sewage to water bodies,
without requiring the extension of distribution lines for electricity.
Renewable energy is more suitable for decentralized use, with the energy
converters located close to consumers and providing supplies at concentrations far
lower than those obtainable with non-renewables. Among the renewable resources
with technical and economic viability to meet the typical demands of sewage
treatment plants, are mainly micro hydro, solar photovoltaic and wind power. The
gases released in sewage treatment plants, when in sufficient quantity and adequate
heat capacity, can often be exploited to recover a part of the heat energy they carry.
The photovoltaic modules can be used to meet the demands of sewage treatment
plants in places where sunlight is sufficient to produce a reasonable annual amount
of energy. However, the costs of PV modules are still quite high and equivalent
to the annual energy cost of a typical treatment plant. But it is possible to design
systems that have the support of a diesel generator or that are connected to the
energy system and are able to meet this energy demand.
Photovoltaic systems provide a power supply that has its own characteristics
and availability of energy concentration. A greater penetration of photovoltaic
components in hybrid systems necessarily require a reduction in their costs. The
increase in the world production of photovoltaic modules and a greater number
of incentive programs for their installation, among some other factors, could
contribute to cost reduction.
This chapter presents the results of a study on up-to-date alternatives for energy
supply to a sewage treatment plant that is under construction. The plant is located
near the town center and therefore connected to the grid. Among the alternatives
considered, wind turbines and photovoltaic modules were included. Generation
systems thus obtained were analyzed with respect to the best composition of the
energy system with regard to the energy cost using the software Homer.
The treatment plant will serve the population of the main city of the municipality,
built 370 meters above the sea level at the highest part of the watershed of Lake
Guaiba, near the lake formed by the dam of Passo Real. This dam [which can be seen
in the image goo.gl/maps/aT3zB (Google Maps, 2014)] is the first in a sequence of
hydraulic structures along the river Jacuí built by the electric state utility company.
The municipality had a total population of 2137 inhabitants in 2000, with 607
residents in the urban core (190 homes) that will be served by the plant. Figure 19.1a
shows a satellite image of the city. Sewage is currently eliminated through individual
solutions. The treated effluent will be disposed in a river that appears in the upper left
corner of Figure 19.1a, according to the current environmental legislation.
Figure 19.1 (a) Satellite image and (b) map of the city of Alto Alegre. On the map,
the city appears divided into two basins, one to the north, in light gray, and the other
to the south, in gray.
The city was divided into two basins, because of the topography of the region.
The sewage from the southernmost part should be pumped to the north, where the
treatment plant is located. These two basins appear on the map of Figure 19.1b.
In Figure 19.1, the map in (b) is covering the region that appears in the image
of (a). On the map, the pumping station appears represented in the central region.
The map also shows the location of the treatment plant in the north, represented as
a rectangle, in a region that does not appear in (a).
The facility is under construction and it will consist of septic tanks followed
by an anaerobic filter. The total area occupied by the septic tanks will be 84 m2,
while the area devoted to the anaerobic filter is 108 m2. The septic tank should be
14 m long and 6 m wide, with a depth of 1.80 m. The filter should be 18 m long and
6 m wide, also with 1.80 m depth. Figure 19.2 shows a scheme of the plant, septic
tanks first and then the anaerobic filter. The effluent will be released respecting
environmental laws.
The main consumption of electrical energy in the treatment plant is due to the
pumping station and blowers used in the filter. The pumping station will consist
of two pumps, one consuming 1475 kWh/h and the other consuming 0737 kWh/h.
The blowers consume 15.77 kWh/h. The average consumption is 539.46 kWh
per month. There is also an estimated consumption of 40.00 kWh per month for
auxiliary loads such as lighting and security system of the components of the
treatment system.
USD$ 90,000, with replacement cost of USD$ 72,000, operating and maintenance
cost of USD$ 4500 per year and lifetime estimated at 12.5 years. The acquisition of
these turbines in pairs reduces costs for USD$ 162,000, USD$ 129,600 and USD$
8100 for each two turbines.
Figure 19.3 Incident solar radiation on a horizontal plane, obtained with software
Homer. In (a), monthly averages; and in (b) daily availability.
operates with 12 V, with nominal capacity of 200 Ah, equivalent to 0.66 kWh,
and a lifetime throughput of 256 kWh. The acquisition cost is USD$ 100, with
replacement cost of USD$ 90.
Figure 19.4 Synthetic series of wind speed at the site of the treatment plant, at a
height of 50 meters. In (a), monthly averages; and in (b) daily availability.
In this study, converters, which carry out the functions of both inverters and
rectifiers in a single component, were considered. The device can operate as
rectifier and inverter with 100% of total capacity, with performance of 85% as an
inverter and 90% as a rectifier. The lifetime is estimated at 12.5 years.
The energy provided by the network has a cost of $ 0.162 during off peak and
$ 0.80 at peak times. The sale occurs with values of $ 0.08 during off peak and $ 0.45
at peak times. The installed power is 22 kW, but the sale to the grid is limited to
10 kW, at no charge for the contracted power or operations d interconnection. The
peak time is from 19 hours to 22 hours.
Electric loads, all in AC, are divided into two sets. The equipment for sewage
treatment has an average consumption of 432 kWh per day and peak consumption
of 18 kW. This set of loads can not experience interruptions in the power
supply. Lighting loads and safety, among other auxiliary loads, have an average
consumption of 50 kWh per day and peak consumption of 4 kW.
Figure 19.5 shows a schematic of the system considered for power supply to
the treatment plant. This figure was extracted from Homer interface. It shows the
DC and AC buses, as well as wind turbine, diesel generator and the connection
to the grid on the left side. It shows the PV modules and battery bank on the right
side. Between the two buses, the two sets of consumer loads and the converter
are shown.
Figure 19.5 Wind PV diesel hybrid system, connected to the grid and storing
energy in batteries, supplying energy to the sewage treatment plant considered in
the study.
This figure was taken from Homer and shows yet another of its commands.
Figure 19.5 left, the commands for insertion of the availability of energy resources.
Figure 19.5 right, the parameters for the simulation related to economic data, with
the control system and greenhouse gas emissions.
U.S. Department of Energy, and is available for universal access in its version
2.68 beta. Homer simulates a system for power generation over the time period
considered in the project at intervals of 60 minutes, presenting the results for a
period of one year (Lilienthal, 2004; Lambert, 2005). Homer interface facilitates
the composition of the system being simulated and the use of input data for the
simulations.
The simulations were performed for an operation period of 25 years, with
12% annual interest and 6% internal rate of return. The different generators and
converters could operate simultaneously and in parallel and that can be adopted
generators with power less than the maximum value of consumer demand. The
generators will be triggered in order to maintain the batteries at a minimum of
80% of their maximum capacity and not just meet the demand.
The hybrid system designed to supply power to the sewage treatment plant
considered in this case study is shown in Figure 19.5. It is a wind PV diesel hybrid
system connected to the grid and storing energy in a battery bank. This system
is also simulated without PV modules, batteries and converter and without grid
connection, but it was always simulated with the maximum annual capacity
shortage equal to zero.
Simulations were performed for the following values for the optimization
variables: 0 kW, 8 kW, 12 kW, 16 kW and 20 kW for PV modules; 0, 2, 4 and 6
wind turbines; 0 kW, 5 kW, 10 kW and 15 kW for diesel generation set; 0, 8, 16,
24 and 32 batteries; 0 kW, 4 kW, 8 kW, 12 kW and 16 kW for the converter; 22 kW
for the purchase capacity and 10 kW for the sell capacity from the grid, when
considered.
Simulations were performed for the following values for the sensitivity inputs:
2 m/s, 4 m/s, 6 m/s, 8 m/s, 10 m/s and 12 m/s for the scaled annual average velocity
of the wind and US$ 0.80, US$ 0.95, US$ 1.10, US$ 1.25 and US$ 1.40 per liter of
diesel.
A set of 2,000 simulations, with 25 different values for the variables of
sensitivity, were performed. The operation was repeated without PV modules,
batteries and converter and without connection to the grid. The results are
presented and discussed in the next section.
Figure 19.6 Optimization space for the system of Figure 19.5, with the ability to sell
excess power to the grid.
The solutions of combining the wind turbines and batteries, including diesel
generators are more favourable in the range of lowest price of the fuel. The
polygonal line that divides gray and light gray hatched areas shows the costs in
these two areas are very similar. Few solutions do not require the use of batteries.
PV modules do not appear in the solution space.
The solution corresponding to the average values of wind speed and cost of
diesel observed in the region is obtained with a combination of three 5 kW diesel
generators, 16 batteries with 200 Ah each, two wind turbines and a converter
with 4 kW. This solution represents a total investment of US$ 467,374, an upfront
investment of US$ 173,100 and an energy cost of US$ 0.195 per kWh.
Levelized costs of the energy required in all cases are shown in Figure 19.7.
Obviously, the higher costs correspond to systems based on diesel consumption
and lower costs correspond to the solutions associated with higher wind speeds. In
the intermediate region, the costs are distributed in horizontal bands, meaning that
they depend more on the wind speed than the cost of fuel.
Similar simulations were performed assuming that the system is connected to
the grid but without the ability to sell the excess energy (Figure 19.8). As expected,
the combination of diesel generator and batteries is preferable at higher wind
speeds compared to Figure 19.6. Similarly, the solutions corresponding to higher
wind speeds and lower costs of diesel also occupy a region slightly larger in the
optimization space. A consequence of the inability to sell excess power is the increase
in the black-stripped region, or a larger number of solutions in the intermediate region
without diesel fuel consumption and with the energy storage in batteries.
Figure 19.7 Levelized costs of energy for the system of Figure 19.5, selling excess
power to the grid.
Figure 19.8 Optimization space for the system of Figure 19.5, without the ability
to sell excess power.
In this case, the solution corresponding to the average values of wind speed and
cost of diesel observed in the region is obtained with a combination of three 5 kW
diesel generators, 16 batteries with 200 Ah each, no wind turbines and a converter
with 4 kW. This solution represents a total investment of US$ 477,381, an capital
investment of US$ 11,100 and a cost of energy of US$ 0.209 per kWh.
Figure 19.9 Optimization space for the system of Figure 19.5, without connection
to the grid.
In this case, the solution corresponding to the average values of wind speed and
cost of diesel observed in the region is obtained with a combination of three 5 kW
diesel generators, 16 battery with 200 Ah each, two wind turbines and a converter
with 4 kW. This solution represents a total investment of US$ 867,023, an initial
investment of US$ 197,000 and a cost of energy of US$ 0.350 per kWh.
The solution which presents the lowest cost was the first, involving a combination
of diesel generators, wind turbines and energy storage in batteries, besides the
connection to the network. The costs of these solutions were high, higher than
those obtained with the simple power supply of the interconnected system. The
need for plant operation without failure of the power supply contributes to higher
costs. Unfortunately, the PV modules do not appear among the feasible solutions.
Table 19.1 summarizes the main results. Three basic configurations of the hybrid
system were simulated. The three settings were composed of 16 200 Ah batteries,
a converter with 4 kW and three 5 kW diesel generators. One of the differences
between the first configuration and the other two is that the former was simulated
including the connection to the grid. Another difference is that the first and third
configurations include three wind turbines, while the second configuration did not
have wind energy.
The lower capital cost of the second configuration is due to the absence of
wind turbines. The small difference compared to the cost of energy of the first
configuration, combined with limited financial availability, can enable this solution.
The difference of the total cost over the initial cost is mainly due to fuel
consumption. The third configuration involves a high annual fuel consumption, to
maintain the supply of electricity for 100% of the time. Better operation strategy
can yield better results.
Beluco and Daronco (2013) provide more details on the results.
19.6 Final Remarks
This chapter evaluated alternatives for energy supply at a sewage treatment
plant located in a small town in southern Brazil. The plant is located near the
center of the city and therefore connected to the grid. Even so, some alternatives
were analyzed, including wind turbines and PV modules, using the software
Homer.
The combination of wind turbines and a diesel generator with PV modules, with
the ability to sell excess power to the grid, results in a cost of US$ 0.195 per kWh.
This cost will raise to US$ 0.209 per kWh if there is no opportunity to sell the
excess energy to the grid. The cost of energy would raise further to US$ 0.350 per
kWh if no PV modules and storage batteries are included.
This feasibility study is only the first step in designing a hybrid system
based on renewable resources for power supply to the sewage treatment plant
considered in this chapter. The large sunny area suggests the installation of PV
modules and the typical wind potential in the region suggests the installation of
wind turbines, but the cost for connection to the grid will certainly represent the
threshold to be reached.
19.7 References
Beluco A. and Daronco G. (2013). Simulation Results with Homer on a Wind PV Hybrid
System for Power Supply of a Sewage Treatment Plant in a Small Town in Southern
Brazil. Internal Report, IPH, UFRGS. Available at www.beluco.net/reports/homer-
alto-alegre.pdf (accessed 22 August 2013).
Beluco A., Souza P. K. and Krenzinger A. (2008). A dimensionless index evaluating the
time complementarity between solar and hydraulic energies, Renewable Energy, 33,
2157–2165.
Beluco A., Souza P. K. and Krenzinger A. (2012). A method to evaluate the effect of
complementarity in time between hydro and solar energy on the performance of
hybrid hydro PV generating plants. Renewable Energy, 45, 24–30.
Caetano Branco Company (2012). Branco Generators Set, www.branco.com.br/us (accessed
28 December 2012)
Cosgrove J. J. (1909). History of Sanitation. Pittsburgh, p. 124.
Daronco G. (2013). Wastewater Treatment Plant – Alto Alegre. Internal Report, CORSAN.
Available at www.beluco.net/reports/wwtp-alto-alegre.pdf (accessed 22 August 2013).
Google Inc. (2014). Google Maps. Available at maps.google.com (accessed 8 August 2014).
Homer Energy (2009). Software HOMER, version 2.68 beta, The Micropower Opyimization
Model. Available at www.homerenergy.com (accessed 9 September 2009).
Lambert T. W., Gilman P. and Lilienthal P. D. (2005). Micropower system modeling with
Homer. In: Integration of Alternative Sources of Energy, F. A. Farret and M. G. Simões
(eds), John Wiley & Sons, pp. 379–418, ISBN: 0471712329.
Lilienthal P. D., Lambert T. W. and Gilman P. (2004). Computer modeling of renewable
power systems. In: Encyclopedia of Energy, Cleveland C. J. (ed.), Elsevier, 1, pp. 633–
647, NREL Report CH-710-36771.
Marco R., Menendez P. E. and Black & Veatch. (2013). How to Use Energy at Waste Water
Plants and How We can Use Less. Available at www.ncsafewater.org/Pics/Training/
AnnualConference/AC10TechnicalPapers/AC10_Wastewater/WW_T.AM_10.30_
Menendez.pdf (accessed 12 December 2013).
Science Channel (2013). What is the history of sewage treatment systems? Available at
curiosity.discovery.com/question/what-history-sewage-treatment-systems (accessed
25 August 2013).
Silveira D. A., Beluco A. and Wartchow D. (2010). Projeção do consumo energético no
tratamento de esgotos sanitários. 2. Congresso Internacional de Tecnologias para o
Meio Ambiente.
Vision Battery (2013). Vision 6FM200D Model. Available at www.vision-battery.com
(accessed 2 April 2013).
iwapublishing.com
@IWAPublishing
ISBN: 9781780405018 (Hardback)
ISBN: 9781780405025 (eBook)