234
PORTFOLIO ANALYSIS OF CARBON SEQUESTRATION TECHNOLOGIES AND BARRIERS TO ADOPTION Jillian D. YOUNG-LORENZ Bachelor of Arts (summa cum laude) This thesis is presented for the degree of Master of Science (Geoscience) at The University of Western Australia School of Earth and Environment 2013

PORTFOLIO ANALYSIS OF ARBON · PORTFOLIO ANALYSIS OF CARBON SEQUESTRATION TECHNOLOGIES AND BARRIERS TO ADOPTION Jillian D. YOUNG-LORENZ Bachelor of Arts (summa cum laude) This thesis

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

PORTFOLIO ANALYSIS OF CARBON

SEQUESTRATION TECHNOLOGIES AND

BARRIERS TO ADOPTION

Jillian D. YOUNG-LORENZ

Bachelor of Arts (summa cum laude)

This thesis is presented for the degree of Master of Science (Geoscience)

at The University of Western Australia

School of Earth and Environment

2013

I

ABSTRACT

The effective targeting of investment funds and research efforts to reduce industrial carbon

dioxide (CO2) emissions, while preserving access to fossil fuel energy resources, requires

quantitative methods to assess and compare a diverse range of carbon capture and sequestration

(CCS) technologies in order to rank, prioritise and make sound policy and investment decisions.

Proposed CCS methods for the stabilisation of atmospheric CO2 levels need to be analysed for

project risks across a range of factors, from the socio-economic to the purely technical. As all

CCS technologies face uncertainties and risks to implementation for global commercial-scale

adoption, I advocate a portfolio approach for carbon management to spread the risk. The main

objective of this thesis is to develop a CCS technology assessment method, and use it to make

recommendations for an optimum carbon dioxide emissions reduction portfolio. I develop a

semi-quantitative methodology, performing the initial design and calibration on the CCS

technology of Geosequestration.

The methodology provides a standardised format for decision makers to assess CCS

technologies from different business, legal, social, scientific and engineering disciplines. It is

intended as a first pass risk assessment and big picture comparison of the current global status

of alternative carbon mitigation options for the purpose of developing a carbon management

strategy and building a portfolio of carbon sequestration technologies to spread risk. Flexibility

in the design of the methodology framework allows for review and modification to account for

technology progress and increased amounts of public domain data. This tool provides a multi-

criteria feasibility analysis which can be used by governments and industry to rank competing

or complementary carbon sequestration options. It can be applied across projects, scientific

disciplines and technologies (portfolio analysis) and with some modification, within a specific

technology or project (project analysis).

I apply the methodology to three carbon sequestration options: Geosequestration, Algae CCS

and Biochar CCS, to assess their technical readiness as carbon mitigation options. Based on the

application of the methodology, the three technologies are evaluated in terms of probability of

adoption from the near to mid-long term. The main differentiators for the top-ranked

technologies are non-technical factors like public acceptance, economics, and the existence of

regulatory frameworks.

II

III

TABLE OF CONTENTS

ABSTRACT ................................................................................................................................. I

TABLE OF CONTENTS ........................................................................................................... III

LIST OF FIGURES .................................................................................................................. VII

LIST OF TABLES .................................................................................................................... XV

ACKNOWLEDGMENTS ..................................................................................................... XVII

DECLARATION OF CANDIDATE CONTRIBUTION ....................................................... XIX

CHAPTER 1 INTRODUCTION ........................................................................................... 1

OVERVIEW .................................................................................................................. 1

1.1 BACKGROUND AND MOTIVATION ........................................................................ 3

1.1.1 Greenhouse gases and climate change ............................................................... 3

1.1.2 Climate mitigation strategies ............................................................................. 8

1.1.3 Carbon capture and sequestration .................................................................... 10

1.1.4 Portfolio Approach ........................................................................................... 11

1.1.5 Needs of LNG Operators NW Shelf of Australia ............................................. 12

1.2 RESEARCH OBJECTIVES........................................................................................... 13

1.3 THESIS ORGANISATION ...........................................................................................13

REFERENCES ............................................................................................................. 15

CHAPTER 2 METHODOLOGY DEVELOPMENT ........................................................ 19

OVERVIEW ................................................................................................................. 19

2.1 INTRODUCTION ........................................................................................................ 21

IV

2.2 MULTI-CRITERIA METHODOLOGY DEVELOPMENT ......................................... 22

2.3 MODEL DEVELOPMENT PROCESS ........................................................................ 26

2.4 FRAMEWORK COMPONENTS ................................................................................. 30

2.5 RANKING SYSTEM .................................................................................................... 35

2.5.1 Sub-Criteria Scorecard ..................... ............................................................... 36

2.6 DISCUSSION ............................................................................................................... 47

2.7 CONCLUSIONS ...........................................................................................................49

REFERENCES ............................................................................................................. 51

CHAPTER 3 APPLICATION OF METHODOLOGY:

GEOSEQUESTRATION:

CARBON CAPTURE AND STORAGE IN GEOLOGICAL FORMATIONS .................. 57

OVERVIEW ................................................................................................................. 57

3.1 INTRODUCTION ........................................................................................................ 59

3.2 METHODOLOGY ....................................................................................................... 62

3.2.1 Technology Element 1: Site Identification and Characterisation ................... 63

3.2.2 Technology Element 2: Capture ...................................................................... 68

3.2.3 Technology Element 3: Transportation ...........................................................72

3.2.4 Technology Element 4: Injection and Storage ................................................73

3.2.5 Technology Element 5: Measurement, Monitoring and Verification .............78

3.3 RESULTS ......................................................................................................................79

3.3.1 Public Acceptance ............................................................................................79

3.3.2 Regulatory Frameworks ...................................................................................81

3.3.3 Economics ....................................................................................................... 85

3.3.4 Storage Quality ................................................................................................ 88

V

3.3.5 Environmental Impact ..................................................................................... 92

3.3.6 Science and Technology: Sub-Matrix Assessment ......................................... 93

3.3.7 Summary of results: Overall Technology Rating .......................................... 102

3.4 DISCUSSION ............................................................................................................ 103

3.5 CONCLUSIONS ........................................................................................................ 107

REFERENCES ........................................................................................................... 109

CHAPTER 4 APPLICATION OF METHODOLOGY: ALGAE CCS .......................... 125

OVERVIEW ............................................................................................................... 125

4.1 APPLICATION OF METHODOLOGY: ALGAE CCS ............................................. 127

4.2 INTRODUCTION ...................................................................................................... 127

4.3 RESULTS .................................................................................................................. 133

4.3.1 Science and Technology: Sub-Matrix Assessment ...................................... 139

4.3.2 Summary of results: Overall Technology Rating .......................................... 146

4.4 DISCUSSION.............................................................................................................. 147

4.5 CONCLUSIONS ........................................................................................................ 148

REFERENCES ........................................................................................................... 151

CHAPTER 5 APPLICATION OF METHODOLOGY: BIOCHAR CCS ..................... 157

OVERVIEW................................................................................................................. 157

5.1 APPLICATION OF METHODOLOGY: BIOCHAR CCS ........................................ 159

5.2 INTRODUCTION ...................................................................................................... 159

5.3 RESULTS ................................................................................................................... 167

5.3.1 Science and Technology: Sub-Matrix Assessment ....................................... 176

5.3.2 Summary of results: Overall Technology Rating............................................ 183

VI

5.4 DISCUSSION ............................................................................................................ 185

5.5 CONCLUSIONS ........................................................................................................ 187

REFERENCES ........................................................................................................... 189

CHAPTER 6 DISCUSSION AND CONCLUSIONS ....................................................... 199

OVERVIEW ............................................................................................................... 199

6.1 DISCUSSION ............................................................................................................ 199

Methodology................................................................................................................ 199

CCS Technology Comparisons .................................................................................. 202

Example of Carbon Management Portfolio .............................................................. 206

6.2 CONCLUSIONS ........................................................................................................ 209

REFERENCES ........................................................................................................... 211

VII

LIST OF FIGURES

FIGURE PAGE

CHAPTER 1 INTRODUCTION

1.1 Historical trends in carbon dioxide concentrations and temperature,

on a geological and recent time scale ........................................................................................... 4

Ahlenius, H. (2007), Historical trends in carbon dioxide concentrations and temperature,

on a geological and recent time scale: from Collection: Global Outlook for Ice and Snow,

UNEP/GRID-Arendal Maps and Graphics Library.

http://maps.grida.no/go/graphic/historical-trends-in-carbon-dioxide-concentrations-and-

temperature-on-a-geological-and-recent-time-scale (accessed 7 May 2011).

1.2 Additional infrastructure costs in Alaska due to thawing of the permafrost ……….......….…… 6

Larsen, P., Goldsmith, S., Smith, O., Wilson, M., Strzepek, K., Chinowsky, P. and

Saylor, B. (2007), Estimating Future Costs for Alaska Public Infrastructure at Risk to

Climate Change: ISER Report, Institute of Social and Economic Research,

University of Alaska, Anchorage, AK, USA.

1.3 Satellite image of the heatwave of the Russian Federation (9 August 2010) ................................ 7

Image source: NASA.

http://eoimages.gsfc.nasa.gov/images/imagerecords/45000/45069/russialsta_tmo_2010208.jpg

(accessed 24 March 2013).

1.4 Rainfall over Pakistan between 26 and 29 July 2010 .................................................................... 7

Pakistan Meteorological Department (2010), Super Flood 2010 in Pakistan:

Brief by Mahmood, A., Director General, Pakistan Meteorological Department.

http://eit.dmi.gov.tr/FILES/sunular/Pakistan.ppt (accessed 24 March 2013).

1.5 Mismatch between computer modelled predictions of rising global mean

temperatures in °C and actual data recorded (adapted from Stott et al., 2013) ............................. 9

The Economist (2013), A sensitive matter: The Economist – Climate Science,

[from the print edition], 30 March 2013.

http://www.economist.com/news/science-and-technology/21574461-climate-may-be-heating-up-

less-response-greenhouse-gas-emissions (accessed 19 April 2013).

VIII

1.6 Pacala and Socolow: Stabilisation Wedges Triangle Model ......................................................... 9

Pacala, S., and Socolow, R. (2004), Stabilization Wedges: Solving the Climate Problem

for the Next 50 Years with Current Technologies, Science, 305(5686), 968-972.

1.7 Global Carbon Pools and fluxes .................................................................................................. 10

Figure adapted from: Lal, R. (2008), Carbon Sequestration: Philosophical Transactions of the

Royal Society B: Biological Sciences, 363, 1492, 815-830.

CHAPTER 2 METHODOLOGY DEVELOPMENT

2.1 Methodology Development Model .............................................................................................. 29

2.2 Master Matrix Step 1 ................................................................................................................... 30

2.3 Science and Technology Sub-Matrix........................................................................................... 30

2.4 Master Matrix Spreadsheet Template........................................................................................... 31

2.5 Science and Technology Sub-Matrix Spreadsheet Template for Geosequestration …................ 31

2.6 Schematic of Master Matrix Scorecard, Sub-matrix Scorecard, and Rating Scale ..................... 33

2.7 Portfolio Analysis Taxonomy....................................................................................................... 35

Adapted from Young-Lorenz, J.D. and Lumley, D. (2013), Portfolio analysis of

carbon sequestration technologies and barriers to adoption: General methodology

and application to geological storage: Energy Procedia, 37, 5063-5079.

2.8 Carbon prices over time in Europe for the period April 2010 to April 2012 .............................. 42

Image source: van Renssen, S. (2012), Saving EU climate policy:

Nature Climate Change, 2, 392-393.

http://www.nature.com/nclimate/journal/v2/n6/images/nclimate1561-f1.jpg

(accessed 15 April 2013).

2.9 Qingdo algae bloom off the coast of Qingdao in the Yellow Sea, China in 2008 ...................... 46

Image source: Earth Observatory, NASA

http://eoimages.gsfc.nasa.gov/images/imagerecords/8000/8897/qingdao_amo_2008180.jpg

(accessed 18 April 2013).

CHAPTER 3 APPLICATION OF METHODOLOGY: GEOSEQUESTRATION: CARBON

CAPTURE AND STORAGE IN GEOLOGICAL FORMATIONS

3.1 Schematic diagram of the Geosequestration Process................................................................... 60

Image source: Scottish Centre for Carbon Storage.

http://www.geos.ed.ac.uk/sccs/public/teachers/CCS-IV.jpg (accessed 7 May 2011).

IX

3.2 Map of Geosequestration Projects, 2012 .................................................................................... 61

Image source: CO2CRC http://www.co2crc.com.au/images/imagelibrary/gen_diag/selected-

large-scale-storage-projects_media.jpg (accessed 9 October 2012).

3.3 Injection of carbon dioxide into depleted oil and gas reservoir ................................................. 64

Image source: Scottish Centre for Carbon Storage.

http://www.sccs.org.uk/public/teachers/Injection-full.jpg (accessed 1 October 2012).

3.4 Phase diagram of CO2 illustrating temperature versus pressure and critical point ..................... 65

Image source: ChemicaLogic Corporation (1999)

http://www.chemicalogic.com/Documents/co2_phase_diagram.pdf (accessed 1 April 2013).

3.5 CO2 Depth versus Density Chart ................................................................................................ 66

Image source: CO2CRC

http://co2crc.com.au/images/imagelibrary/stor_diag/supercritical_highres.jpg

(accessed 1 October 2012).

3.6 Carbon Dioxide Capture Processes for LNG Processing, Post-Combustion,

Pre-Combustion and Oxyfuel....................................................................................................... 70

Image source: CO2CRC

http://co2crc.com.au/images/imagelibrary/cap_diag/Captureprocesses_media.jpg

(accessed 13 July 2012).

3.7 Absorption – Desorption Process................................................................................................. 72

Image source: CO2CRC

http://www.co2crc.com.au/images/imagelibrary/cap_diag/solvent_absorption_media.jpg

(accessed 13 July 2012).

3.8 CO2 Trapping Mechanisms - CO2 Trapped, and Applicability versus Time ............................ 74

Image source: Elements – Geoscience World

http://elements.geoscienceworld.org/content/4/5/325/F7.large.jpg (accessed 5 April 2013).

3.9 Structural and Stratigraphic Trapping ........................................................................................ 75

Image source: CO2CRC. http://www.co2crc.com.au (accessed 2 October 2011).

3.10 Residual Saturation /Capillary Trapping .................................................................................... 76

Image source: CO2CRC. http://www.co2crc.com.au (accessed 2 October 2011).

X

3.11 Solubility Trapping ....................................................................................... ............................. 77

Image source: CO2CRC

http://www.co2crc.com.au/images/imagelibrary/stor_diag/co2_dispersion_post_inject_media.jpg

(accessed 2 October 2011).

3.12 CO2 sources, pipelines and projects in the USA ......................................................................... 86

Image source: Melzer, L.S. (2011), Emergence Of Residual Zones, Price And Supply Factors

Usher In New Day In CO2 EOR: American Oil and Gas Reporter, February 2011 Cover Story.

http://www.aogr.com/index.php/magazine/cover-story/emergence-of-residual-zones-price-and-

supply-factors-usher-in-new-day-in-co (accessed 17 April 2013).

3.13 Carbon prices over time in Europe for the period April 2010 to April 2012 .............................. 88

Image source: van Renssen, S. (2012), Saving EU climate policy: Nature Climate Change, 2,

392-393.

http://www.nature.com/nclimate/journal/v2/n6/images/nclimate1561-f1.jpg

(accessed 15 April 2013).

3.14 The Gorgon Project Development Concept ................................................................................. 89

Image source: Chevron Australia

http://www.chevronaustralia.com/ourbusinesses/gorgon/globalsignificance.aspx

(accessed 17 April 2013).

3.15 Geophysical Techniques to characterise and monitor CO2 storage reservoirs ............................ 91

Image source: Subsurface Imaging Group, The Ohio State University

http://www.geology.ohio-state.edu/~jeff/geophysics1.htm (accessed 23 April 2012).

3.16 CO2 storage potential in Australia .............................................................................................. 94

Image source: CO2CRC (2011)

http://www.co2crc.com.au/images/imagelibrary/gen_diag/Australia_Storage_Potential_2011_me

dia.jpg (accessed 17 April 2013).

3.17 Pipeline Infrastructure Map of Western Australia as at 30 June 2010 ........................................ 97

Image Source: Economic Resource

http://www.erawa.com.au/access/gas-access/pipeline-infrastructure-map/

(accessed 17 April 2013).

XI

3.18 Onshore vs. Offshore storage costs for depleted oil and gas fields in Europe ............................ 99

Image source: Global CCS Institute

http://www.globalccsinstitute.com/publications/global-storage-resources-gap-analysis-policy-

makers/online/83236 (accessed 17 April 2013).

CHAPTER 4 APPLICATION OF METHODOLOGY: ALGAE CCS

4.1 Algal Biofuel and Bioenergy Pathways .................................................................................... 129

Figure adapted from: NNFCC (2012), Research Needs in Ecosystem Services to Support Algal

Biofuels,Bioenergy and Commodity Chemicals Production in the UK: A project for the Algal

Bioenergy Special Interest Group [Smith, N. and Higson, A.]: NNFCC Project Number: 12-008.

4.2 Example of open raceway pond in Israel for the cultivation of microalgae for biofuel ............ 130

Image source: Seambiotic

http://www.seambiotic.com/research/microalgae-speices/ (accessed 24 April 2013).

4.3 Example of a closed column tubular photobioreactor for microalgae

cultivation in China ............................................................................................ ...................... 130

Image source: Algae Energy Group

http://www.bioenergychina.org/ea/Facility.html (accessed 24 April 2013).

4.4 Aerial view of Cyanotech algae farm and facilities at Keahole Point, Hawaii ………………. 139

Image source: Cyanotech Corporation http://www.cyanotech.com/company/facility.html

(accessed 24 April 2013).

4.5 The Green Crude Farm facility being developed in Columbus, New Mexico, USA

by Sapphire Energy, the US Department of Energy [US DOE] and the

US Department of Agriculture ................................................................................................. 140

Image source: Sapphire Energy

http://www.sapphireenergy.com/locations/green-crude-farm.html

(accessed 25 April 2013).

4.6 Horizontal tubular photobioreactor microalgae cultivation system and biorefinery

at AlgaePARC Research Facility, Wageningen University and Research Centre,

Netherlands ............................................................................................................................... 145

Image source: AlgaePARC, Wageningen UR

http://www.wageningenur.nl/en/Expertise-Services/Facilities/AlgaePARC/Themes/Microalgae-

biorefinery.htm (accessed 24 April 2013).

XII

CHAPTER 5 APPLICATION OF METHODOLOGY: BIOCHAR CCS

5.1 Example of biochar.................................................................................................................... 160

Image source: Australia and New Zealand Biochar Researchers Network

http://www.anzbiochar.org/images/char600x358.jpg (accessed 24 April 2013).

5.2 Microscopic image of biochar showing its micropore structure............................................... 161

Image source: Biochar Project (Jocelyn)

http://biocharproject.org/wp-content/uploads/2011/08/Jocelyn-biochar-electron-microscope-

images-1.jpg (accessed 24 April 2013).

5.3 Potential ecosystem benefits from biochar production and application systems ...................... 163

Figure adapted from: Waters, D., Zwieten, L., Singh, B. P., Downie, A., Cowie, A. L.,

and Lehmann, J. (2011), Biochar in soil for climate change mitigation and adaptation:

In Soil Health and Climate Change: Springer Berlin Heidelberg, 15, 345-368.

5.4 Overview of the sustainable biochar concept ........................................................................... 163

Woolf, D., Amonette, J.E., Street-Perrott, F.A., Lehmann, J., and Joseph, S. (2010),

Sustainable biochar to mitigate global climate change: Nature Communications: 1, 56, 1-9.

5.5 Potential impacts of Biochar application to plant-soil systems ................................................. 165

Figure adapted from: Waters, D., Zwieten, L., Singh, B. P., Downie, A., Cowie, A. L.,

and Lehmann, J. (2011), Biochar in soil for climate change mitigation and adaptation:

In Soil Health and Climate Change: Springer Berlin Heidelberg, 15, 345-368.

5.6 Bales of corn stover collected from grain fields after harvesting .............................................. 171

Image source: Science Daily

http://images.sciencedaily.com/2007/07/070709103138-large.jpg (accessed 24 April 2013)

5.7 Modern pyrolysis stove ............................................................................................................. 175

SHALIN Finland in collaboration with Aalto University (link), Household Level Pyrolytic

StoveLinking energy, Forestry and Agriculture for local level climate change mitigation

and improved food security: SHALIN Finland in collaboration with Aalto University,

[Kuria, P. and Kagiri, E. (eds.), Lindberg, A., Mäkinen, I., Mäkelä, M., Ottelin, J., Kuria, P.

and Kagiri, E.]. http://www.into-ebooks.com/download/250/ (accessed 22 April 2013).

XIII

5.8 Traditional three-stone stove used in Kenya ............................................................................ 176

SHALIN Finland in collaboration with Aalto University (link), Household Level Pyrolytic

Stove Linking energy, Forestry and Agriculture for local level climate change mitigation

and improved food security: SHALIN Finland in collaboration with Aalto University,

[Kuria, P. and Kagiri, E. (eds.), Lindberg, A., Mäkinen, I., Mäkelä, M., Ottelin, J., Kuria, P.

and Kagiri, E.]. http://www.into-ebooks.com/download/250/ (accessed 22 April 2013).

5.9 500kW biomass gasification plant in Merced, California, USA .............................................. 181

Image source: International Biochar Initiative [IBI] (2012), Phoenix Energy’s business

model: Building small profitable plants: [in September 2012 newsletter][photo courtesy of

Phoenix Energy]. http://www.biochar-international.org/profile/Phoenix_Energy

(accessed 24 April 2013).

XIV

XV

LIST OF TABLES

TABLE PAGE

CHAPTER 1 INTRODUCTION

1.1 Predicted decrease in near-surface permafrost area and increase in thickness

of the active layer ........................................................................................................ ................. 5

Table source: United Nations Environment Programme [UNEP] (2012), Policy

Implications of Warming Permafrost: [Report], United Nations Environment

Programme, Nairobi, Kenya. http://www.unep.org/pdf/permafrost.pdf

(accessed 22 March 2013).

CHAPTER 2 METHODOLOGY DEVELOPMENT

2.1 Identified Evaluation Criteria....................................................................................................... 22

2.2 Examples of Stakeholder Diversity.............................................................................................. 23

2.3 Associated Issues for Regulatory Frameworks............................................................................ 24

2.4 Economic and Financial Factors ................................................................................................. 25

2.5 Evaluation Criteria Rating Scale and Equivalents ...................................................................... 33

2.6 Evaluation Criteria Scorecard ................................................................................ ...................... 36

2.7 Sub-Criteria Scorecard ............................................................................................................... 37

CHAPTER 3 APPLICATION OF METHODOLOGY: GEOSEQUESTRATION:

CARBON CAPTURE AND STORAGE IN GEOLOGICAL FORMATIONS

3.1 Rock properties affecting ability to sequester CO2...................................................................... 67

3.2 Sedimentary basin assessment factors ........................................................................................ 68

3.3 Illustration of average score of sub-criteria input into Public Acceptance column of

the Master Matrix ....................................................................................................................... 81

3.4 Measurement, Monitoring and Verification (MMV) technologies and methods....................... 100

XVI

3.5 Technology Elements of Geosequestraton: Science and Technology

Sub-Matrix Assessment ............................................................................................................ 102

3.6 Overall risk assessment rating of Geosequestration: Master Matrix ........................................ 102

CHAPTER 4 APPLICATION OF METHODOLOGY: ALGAE CCS

4.1 Science and Technology Sub-Matrix Scorecard: Results of Application of Methodology

to Algae CCS ............................................................................................................................ 146

4.2 Overall technology rating of Algae CCS: Master Matrix ......................................................... 146

CHAPTER 5 APPLICATION OF METHODOLOGY: BIOCHAR CCS

5.1 Typical product yields (dry wood basis) obtained by different modes of wood pyrolysis ....... 162

Adapted from: International Energy Agency Bioenergy [IEA Bioenergy] (2007),

Biomass Pyrolysis: IEA Bioenergy [Bridgwater, T]: Aston University, United Kingdom,

T34:2007:01.

5.2 Science and Technology Sub-Matrix Scorecard: Results of Application of

Methodology to Biochar CCS.................................................................................................... 183

5.3 Overall technology rating of Biochar CCS: Master Matrix ...................................................... 183

CHAPTER 6 DISCUSSION AND CONCLUSIONS

6.1 Master Matrix: Overall Technology Ratings for Geosequestration, Algae CCS

and Biochar CCS ....................................................................................................................... 205

6.2 Science and Technology Sub-Matrix Scorecard: Results of Application of

Methodology to Geosequestration ……………................………...………………………… 205

6.3 Science and Technology Sub-Matrix Scorecard: Results of Application of

Methodology to Algae CCS ...................................................................... ................................ 206

6.4 Science and Technology Sub-Matrix Scorecard: Results of Application of

Methodology to Biochar CCS .................................................................... ............................... 206

XVII

ACKNOWLEDGEMENTS

Firstly, I would like to thank my coordinating supervisor, Winthrop Professor David Lumley,

for his insight, guidance and talent for understanding the “whole picture.” He has the ability to

to bring together the scientific and business perspectives needed to understand the multi-

disciplinary challenges in the adoption of carbon capture and sequestration technologies. I have

learned an inordinate amount from various scientific disciplines, notably geoscience, and

possibly more of maths and physics than I thought possible! I also thank my co-supervisor

Winthrop Professor David Day from the UWA Business School of Management and Leadership

for his continued support.

My research would not have been possible without the generous support of the University of

Western Australia, School of Earth Environment, the Centre for Petroleum Geoscience and CO2

Sequestration, and the UWA: Reservoir Monitoring Research Consortium. My thanks and

appreciation.

I thank the School of Earth and Environment’s Graduate Research School Coordinators,

Andrew Rate and Eun Jun Holden, for their practical advice and support. The same applies to

the Graduate Research School Education Officers, Krys Haq and Jo Edmonston. You were all

very helpful at all stages of my candidature, and I appreciate all your assistance. I would also

like to thank Adjunct Associate Professor Jo Voola for giving me my initial start at UWA in the

School of Mechanical and Chemical Engineering.

I thank my colleagues on the 4D Research Team and at the Centre for Petroleum Geoscience.

Big cheers to Helen Nash, Lisa Gavin, Philbilly Cilli, Mohammad (MoMo) Emami, Matty Saul,

Mike Dentith, James Deeks, Rafael de Souza, Wendy Young, Jeff Shragge, Nader Issa, Mudasa

Muhammad Saqab, UGuen Jang, Alan Pitman, and Polly Mahapatra. What a great team! Other

friends who have always been there when I needed them are Mao Luo, Tom Hoskins, Patricia

Alessi, and the administrative team at SEE North. And I cannot forget Michael Djohan; FNAS

IT would just break down without you! I value everybody’s advice, friendship, love and support

these past two years. I will miss the fun times, the jolly decent food at Uniclub, the

opportunities to bounce ideas, and even offering feedback on geophysics problems!

I thank God for his many blessings, including the support of my family at home in New

Zealand, and friends from St. Andrews and around the world. And finally, I would like to thank

my beloved husband, Lars Lorenz and fluffy son, Jaeger. Lars, I would not have been able to do

it without your loving support and fine editing skills. You helped to lessen the impact of my

complicated and convoluted German-style sentences. Thank you for your constant strength,

XVIII

humour, wisdom and love. Jaeger, you keep me sane, give me heaps of joy and playful fun, and

kept the loneliness at bay since your Daddy has been away.

Thank you one and all.

XIX

DECLARATION OF CANDIDATE CONTRIBUTION

This thesis is my own work, and no part of it has been presented for a degree at this, or at any

other, university.

This thesis contains published work and/or work prepared for publication, some of which has

been co-authored. The bibliographical detail of the work is outlined below (with percentage

contributions from authors in parentheses). This paper is based on material from Chapters 2 and

3 of this thesis.

Young-Lorenz, J.D. (70%) and Lumley, D. (30%) (2013) Portfolio analysis of carbon

sequestration technologies and barriers to adoption: General methodology and application to

geological storage: Energy Procedia, 37, 5063-5079.

We hereby declare that the individual authors have granted permission to the candidate

(Jillian D. Young-Lorenz) to use the results presented in this publication.

Student Signature .............................................................................

Coordinating Supervisor Signature .....................................................

XX

Introduction

1

CHAPTER 1

INTRODUCTION

OVERVIEW

This chapter introduces the background context of greenhouse gas (GHG) emissions and their

possible links to global warming and climate change. Although the causal link between

anthropogenic GHG and climate change cannot yet be definitively proven, the potential risks to

environment, economies, food and energy sources justify the development of carbon mitigation

strategies and tools for preventative measures. A global goal, shared by many governments and

international bodies, is the reduction of carbon dioxide (CO2) emission levels in the atmosphere.

Carbon capture and sequestration (CCS) technologies through biotic or abiotic means are

examples of such strategies. As all CCS technologies face uncertainties and risks to

implementation for global commercial-scale adoption, I advocate a portfolio approach for

carbon management to spread the risk. I introduce the development of a CCS technology

assessment method to be applied to various CCS technologies and to make recommendations

for an optimum CO2 emissions reduction portfolio. A description of the thesis organisation

concludes this chapter.

Introduction

2

Introduction

3

1.1 BACKGROUND AND MOTIVATION

1.1.1 Greenhouse Gases and Climate Change

Global emissions of greenhouse gases (GHGs) covered by the United Nations Framework

Convention on Climate Change [UNFCCC] have increased approximately 70% from 1970–

2004, and by 24% from 1990–2004 (Intergovernmental Panel on Climate Change [IPCC],

2007a) (Figure 1.1). GHGs absorb infrared radiation trapping heat within the Earth’s

atmosphere, and include the six common gases: carbon dioxide (CO2), methane (CH4), nitrous

oxide (N2O), hydrofluorocarbons (HFCs), perfluorocarbons (PFCs), and sulphur hexafluoride

(SF6). These gases can be compared by multiplying the emissions of a given gas by its global

warming potential to derive a common unit of CO2-equivalent (CO2-e) (IPCC, 2007a; Forster et

al, 2007). CO2 is thought to contribute ~64% of the total GHGs and radiative forcing, primarily

from combustion of fossil fuels, deforestation and changes in land-use since 1750 (IPCC,

2007a; Forster et al., 2007; World Meteorological Service [WMO], 2012a). Methane is the

second most important GHG contributing ~18% of total GHGs since 1750, with 60% of

methane emissions from human activities (cattle-rearing, rice planting, fossil fuel exploitation

and landfills), and 40% from natural sources (wetlands, permafrost)(WMO, 2012a). The third

largest contributor is nitrous oxide at ~6% of total GHGs, from anthropogenic (biomass

burning, fertiliser manufacture and application, industrial processes) and natural (oceans)

sources. The impact of nitrous oxide on the climate is ~300 times greater than equal emissions

of CO2 (IPCC, 2007a; Forster et al., 2007; WMO, 2012a), and in addition, this gas contributes

to the destruction of the stratospheric ozone layer.

The effect of CO2 on the global temperature was first described at the end of the nineteenth

century by Svante Arrhenius in 1896 as a response to a discussion at the Physical Society of

Stockholm about the possible cause of the Ice Age. He described the effect of a change in CO2

concentration on temperature in the following manner:

“If the quantity of carbonic acid increases in geometric progression, the augmentation

of the temperature will increase nearly in arithmetic progression.” (Arrhenius, 1896)

According to the IPCC Report (2007a) Climate Change 2007: The Physical Science Basis,

the CO2 concentration in the atmosphere in 2005 far exceeded “the natural range over the last

650,000 years (180 to 300 parts per million (ppm)) as determined from ice cores” (Figure 1.1).

The Report states that the increase in concentrations of CO2, CH4 and N2O are very likely the

causes of the increased total radiative forcing of the Earth’s climate since 1750. By 2011, the

Introduction

4

globally averaged mole fractions (number of molecules of the gas per million/billion molecules

of dry air) of CO2, CH4, and N2O in the atmosphere reached new highs, with CO2 reaching

390.9 ± 0.1ppm, CH4 1813 ± 2 parts per billion (ppb), and N2O 324.2 ± 0.1 ppb (WMO, 2012a).

Figure 1.1. Historical trends in carbon dioxide concentrations and temperature, on a geological and

recent time scale. Increasing temperatures are notable from the middle ages onward as the world

emerged from the Little Ice Age (LIA), around 1850. With the industrial era, human activities have at the

same time increased the level of carbon dioxide (CO2) in the atmosphere, primarily through the burning

of fossil fuels (IPCC, 2007a; Stern, 2007). The top part of the CO2 measurements, the observations, are

referred to as the 'Mauna Loa curve' or the 'Keeling curve' in reference to the dramatic increase in

atmospheric CO2 levels documented by Charles David Keeling from the Mauna Loa volcano research

station in Hawaii (Image source: Ahlenius, 2007).

The WMO, in its Provisional Statement on the Global Climate in 2012, describes the

previous twelve years as the warmest since records began in 1850. January to October 2012 was

the ninth warmest such period with global land and ocean surface temperatures for the period

about 0.45°C above the corresponding 1961–1990 average of 14.2°C (World Meteorological

Introduction

5

Organisation (WMO), 2012b). One of the most marked events attributed to the increase in

global temperatures, is the rapid loss of Arctic sea ice. The Report by Koç et al. in 2009 states

that rapid sea ice loss is an indicator of global climate change, with the Arctic shrinking by 40%

in the period from 1979 to 2009 (13% pa), and sea ice cover diminishing significantly faster

than climate model predictions. Between March and September 2012, the Arctic experienced a

sea ice loss of 11.83 million km2

(an area larger than the USA), the largest seasonal ice extent

loss in the 34-year satellite record (WMO, 2012c). Climate models predict that summer sea ice

may disappear in the Arctic within four to forty years (Plumer, 2012). A recent study of ice

cores from the Antarctic suggests that an accelerated snow melt is occurring in the Southern

Hemisphere as well. Even though it is not the only recorded period of increased snow melt, the

intensity of the melting since the beginning of the 20th century is unprecedented in the 1000 year

record analysed (Abram et al., 2013).

Another aspect of increasing average temperatures in the Arctic zone is the deepening of the

active layer of permafrost, the layer which undergoes annual thawing through the summer

months. By increasing the thickness of the layer, more organic matter which was previously

sequestered in the permafrost will be accessible for decay. This process is referred to as

permafrost carbon feedback and could potentially contribute 43 – 135 Gigatons (Gt) CO2-e by

2100 and 246-415 Gt CO2-e by 2200 to global GHG emissions. A large range of uncertainties is

related to these estimates as these feedback processes are currently not well understood, and

estimates for the affected portion of the permafrost differ widely (Table 1.1). A more immediate

impact will be felt in the form of damage to infrastructure and buildings which rely on the

permafrost as a solid foundation. A report by Larsen et al. from 2007 attempts to estimate the

additional economic costs for infrastructure in Alaska, concluding that the additional costs are

in the order of 10% - 15% (Figure 1.2).

Table 1.1. Predicted decrease in near-surface permafrost area and increase in thickness of the active

layer (United Nations Environment Programme [UNEP], 2012)

Introduction

6

Figure 1.2. Additional infrastructure costs in Alaska due to thawing of the permafrost (Larsen et al.,

2007).

Based on data and observations gathered from 1950, the IPCC stated that there are

indications of changes in some extreme weather and climate events (IPCC, 2012), but due to the

rarity of such extreme events and the globally varying quality of the available data, no firm

conclusions can currently be drawn. The following excerpt from the 2012 IPCC Special Report

on Managing the Risks of Extreme Events and Disasters to Advance Climate Change

Adaptation reflects their assessment, based on available scientific evidence:

It is likely that anthropogenic influences have led to warming of extreme daily minimum

and maximum temperatures at the global scale. There is medium confidence that

anthropogenic influences have contributed to intensification of extreme precipitation at the

global scale. It is likely that there has been an anthropogenic influence on increasing

extreme coastal high water due to an increase in mean sea level. The uncertainties in the

historical tropical cyclone records, the incomplete understanding of the physical

mechanisms linking tropical cyclone metrics to climate change, and the degree of tropical

cyclone variability provide only low confidence for the attribution of any detectable changes

in tropical cyclone activity to anthropogenic influences. Attribution of single extreme events

to anthropogenic climate change is challenging.

The multiple weather and climate extremes and record-breaking events observed in recent

years, as detailed by the WMO in 2012 in its annual Provisional Statement on the State of the

Global Climate, resulted in increased awareness and sensitivity of the public to the effects of

climate change (Donner and McDaniels, 2013). Major heat waves and extremely high

temperatures impacted the northern hemisphere; droughts and wildfires affected nearly two-

thirds of the United States, western Russia (Figure 1.3), and southern Europe (WMO, 2011,

2012a, 2012b). The weather in West Africa, South East Asia and Pakistan (Figure 1.4) was

characterised by extremely heavy rainfall and severe flooding. Some regions, such as East

Introduction

7

Africa and Russia, were affected by multiple extreme events (WMO, 2012a). From late 2012 to

early 2013, Australia experienced record-breaking temperatures, bushfires, as well as intense

rainfall and flooding (Commonwealth of Australia, 2013). While these events are challenging to

causally link to global warming and GHG emissions as detailed by the IPCC, they find their

way into all levels of everyday society and politics as illustrated by their inclusion in the Press

et al. (2013) Australian Strategic Policy Institute, Special Report 49:

Climate change: The more severe effects of climate change, in particular the increase in

frequency and severity of natural disasters, compounded by competition over scarce natural

resources, may contribute to instability and tension around the globe, especially in fragile

states.

Figure 1.3. Satellite image of the heatwave of the Russian Federation (9 August 2010) (Image source:

NASA, 2010).

Figure 1.4.

Rainfall over Pakistan between 26

and 29 July 2010 (Image source:

Pakistan Meteorological

Department)

Introduction

8

The increased awareness of the general public and governments about the potential or real

economic and environmental costs of global warming, and further increases of GHGs drove the

implementation of the United Nations [UN] Kyoto Protocol in 1998 for the reduction of GHGs.

In recent years, more attention was directed to the impacts of climate change with the

implementation of emission limits (European Parliament and the Council of the European

Union [EU], 2001), carbon taxes (The Commonwealth of Australia, 2012), cap and trade

schemes (EU, 2013), or voluntary commitments for CO2 reduction (Aldy and Stavin, 2011).

1.1.2 Climate Mitigation Strategies

Carbon dioxide (CO2) is the largest contributor to increases in GHG concentrations, having

grown by about 80% (IPCC, 2007a). In order to limit temperature increases from pre-industrial

levels to 2°C, several global climate models suggest that atmospheric concentrations of GHG

must be stabilised at 450ppm of CO2-e by 2050 (IPCC, 2007a). In 2013, at the World Economic

Forum in Davos, Lord Stern said that he underestimated the risks in 2007, and should have been

more “blunt” about the threat to the economy from rising temperatures in The Stern Review: The

Economics of Climate Change. The Review indicated a 75% chance of global temperatures

rising by 2°C - 3°C, but Lord Stern now believes it will rise by 4°C – 5°C, as the atmosphere

and planet are not absorbing as much carbon as expected, and CO2 emissions are rising strongly

(Stewart and Elliott, 2013). Recent studies and failed climate negotiations highlight that it is

unlikely global average temperatures will be restricted to a rise of below 2°C with the current

trajectory of CO2 emissions (Peters et al., 2012; UNFCCC, 2011). Peters et al. state that

“Unless large and concerted global mitigation efforts are initiated soon, the goal of

remaining below 2°C will very soon become unachievable.”

On the other hand, a recent article in The Economist (2013) notes that other recent studies

observe that global mean temperatures have not risen as predicted (Figure 1.5) (Stott et al.,

2013) although GHG concentrations continue to rise. The 2007 IPCC Report on Climate

Change: The Physical Science Basis estimates that global temperatures would increase from

pre-industrial times in the range of 2°C to 4.5°C, with a best estimate around 3°C and an

increase of less than 1.5°C would be very unlikely. However, numerous researchers have

modelled various predictions revising the sensitivity down to a range from 1.2° to 4.1°C (The

Economist, 2013).

Introduction

9

Figure 1.6. Stabilization Triangle Model (Image source: Pacala and

Socolow, 2004).

In 2004, Pacala and Socolow published their Stabilisation Triangle Model (Figure 1.6),

outlining in a schematic manner a set of suggested measures to restrict the atmospheric CO2

concentration to 500+/-50 ppm (less than twice the pre-industrial concentration of 280 ppm).

The difference between a business-as-usual scenario and a constant 2004 base level is reflected

by the Stabilisation Triangle. The Triangle is subdivided into seven wedges, each representing a

current technology with a reduction potential of 1 Gigatonnes of carbon per year (GtCpa) by

2054 (Figure 1.6). Mitigation technologies include building, vehicle and coal-plant efficiencies,

use of nuclear power, and

renewables such as wind,

carbon capture and

sequestration, biofuels, and

several natural sink

wedges (afforestation,

biofuel energy plantations,

and no till/conservation

tillage systems). Although

mitigation strategies have

been recommended by

various international

bodies to address climate

Figure 1.5.

Mismatch between

computer modelled

predictions of rising global

mean temperatures in °C

and actual data recorded

(Image source: The

Economist, adapted from

Stott et al., 2013).

Introduction

10

change and offset the effects of GHGs, such measures have not been undertaken in the volumes

necessary (Brito and Stafford Smith, 2012; IEA, 2008; Global CCS Institute, 2012).

1.1.3 Carbon Capture and Sequestration

Definition

Carbon capture and sequestration (CCS), refers to the capture of carbon dioxide from the

atmosphere and its transfer into other carbon forms including oceanic, pedologic (soils), biotic

(plants and animals) and geological formations, for long term storage. The principal global

carbon stores and movement of carbon among these reservoirs (fluxes) are illustrated in Figure

1.6. Capture and sequestration of atmospheric CO2 can be achieved through engineered means

(e.g. chemical trapping of CO2 from industrial flue gases), or abiotic (mineralisation) and biotic

processes (photosynthesis). For the remainder of this thesis, I will use the terms carbon

sequestration and CCS interchangeably for engineered, abiotic and biotic methods.

Figure 1.7. Principal global carbon pools and fluxes between them. SOC = soil organic carbon (humus,

inert charcoal, plant and animal residues, microbiota). SIC = soil inorganic carbon (elemental carbon,

carbonate minerals, e.g. calcite, primary and secondary carbonates). Carbon pool sizes are given in

petagrams (Pg), which are also known as a Gigaton (Gt). A petagram is equal to one quadrillion

(1,000,000,000,000,000 or 1015

) grams (g). The transfer of carbon fluxes are shown in Pg per annum (Pg

pa) (adapted from Lal, 2008).

Introduction

11

The CCS option covers a diverse range of technologies. For example, ocean fertilisation

technologies aim to spur the growth of plankton by the addition of nutrients and increase the

dead biomass and naturally sequestered CO2 at the ocean floor. Biofuel production aims to

provide a carbon neutral or negative energy source by using biomass as the basis for fuel

generation instead of hydrocarbons. The conversion of any bio-material into biochar creates a

material with very high carbon content which can be used as a fuel, for soil enrichment, or as a

space efficient long term storage solution. One of the most promising options is carbon capture

and storage in geological formations (Geosequestration). Geosequestration involves the capture

of CO2 at industrial emission source points (coal-fired power generation, steel, chemical,

fertiliser, cement, mineral, or LNG plants), and transport of the captured CO2 via pipeline or

alternative method to suitable sites for underground injection and long-term geological storage.

According to the IEA 2008 Report, CCS in geological formations may be capable of accounting

for about 20% of the recommended CO2 emissions reduction by 2054. But significant

challenges to widespread adoption remain with very slow progress since 2004.

1.1.4 Portfolio Approach

Many people in society desire to mitigate the risks of climate change by reducing industrial

carbon dioxide (CO2) emissions while preserving access to fossil fuel energy sources. Most

projections estimate fossil fuels will continue to be the primary source of energy until at least

2050 (IPCC, 2005; IEA, 2008). Global energy demand, mainly from electricity generation, is

forecast to increase 35% for the period 2010 to 2040, with developing nations expected to

account for 65% of the rise (ExxonMobil, 2013).

All of the currently proposed CCS technologies to mitigate atmospheric GHG face barriers

to wide scale commercial adoption. A multitude of reasons can be found for this, including

industrial readiness and feasibility of possible mitigation technologies which originate from a

widely diverse range of scientific and technical/engineering disciplines. Feasibility could also

be related to economic, capacity, legal or social acceptance issues. The challenges and

uncertainties involved in the implementation of CCS technologies suggest industry and

governments should manage their risk exposure by utilising a portfolio of CO2 sequestration

technologies to spread the risk, rather than relying on a single method (Grubb, 2004). The risk

associated with reliance on a single CCS option is illustrated by the challenges Statoil

encountered with the Snøhvit Project. Due to lower reservoir permeability than originally

predicted, CO2 injectivity was also lower, resulting in decreased injection rates for the processed

Introduction

12

CO2 from the liquefied natural gas (LNG) plant (Molde, 2011). This significantly impacted the

project on both technical and economic levels as the operator did not have an alternate storage

option or carbon offset strategy to rely upon.

The risk-reward assessment by a project operator may vary widely depending upon its

dependency on a project or overall corporate strategy. For example, an operator who has the

opportunity to vent CO2 even if a penalty must be paid, or who can rely upon carbon reduction

credits generated from other CCS technologies in its carbon management strategy, is likely to be

more risk tolerant than an operator who has to stop or reduce production in the case of lower

than anticipated CO2 storage volumes. I suggest that rather than committing only to one solution

such as Geosequestration, a diversified portfolio approach may provide a lower risk exposure

and better outcomes.

1.1.5 Specific Needs of Liquefied Natural Gas (LNG) Operators on the

North West Shelf of Australia

Increased regulatory requirements and the introduction of a carbon tax are concerns for oil

and gas (O&G) operators, and more specifically for LNG operators as they consider

investments of approximately AUD250 billion to develop the multiple giant gas fields of the

North West (NW) Shelf of Australia (Lumley, 2010). The fields contain more than 150 trillion

cubic feet (Tcf) of natural gas which is intended for export via LNG (Geoscience Australia and

Bureau of Resources and Energy Economics (BREE), 2012). As some of these offshore gas

fields, such as the Greater Gorgon field in the Carnarvon Basin, and the Browse and Bonaparte

Basins, can be characterised by a high CO2 content, for example, up to 27% for Evans Shoal

(Geoscience Australia, 1998; Lumley, 2010), operators of these fields face the challenge of

disposing of the produced CO2, either by venting, which may become more heavily regulated in

the future, or by implementing a more climate friendly solution.

To offset the potentially substantial costs of a carbon tax on emissions (high CO2 content and

LNG processing), operators need an affordable option that is deployable in remote areas.

Ideally, such a solution can take advantage of existing technology and experience, with limited

additional infrastructure build, and readily available storage sites. As an alternative to

atmospheric venting, operator Chevron and its joint venture partners of the Gorgon Project,

being undertaken to develop the Gorgon and Jansz-lo Gas Fields on the NW Shelf, have opted

for Geosequestration. The CO2 Injection Project will inject up to 4.1 million tonnes per annum

Introduction

13

(Mtpa) of CO2 from the ~15% CO2 content in the raw gas into the Dupuy Formation, which is

about 2200 metres below Barrow Island, a Class A Nature Reserve (Scott et al., 2010).

1.2 RESEARCH OBJECTIVES

To facilitate the most effective targeting of investment funds and research efforts, as well as

manage the risk exposure across a variety of projects, both governments and industry need

quantitative methods to assess and compare carbon capture and sequestration (CCS)

technologies in order to rank, prioritise and make sound policy and investment decisions.

Proposed CCS methods for the stabilisation of atmospheric CO2 levels need to be analysed for

technology and project risks across a range of factors, from the socio-economic to the purely

scientific or technical. The many factors impacting the wide scale commercial adoption of CCS

methods need to be analysed within a consistent framework.

The main objective of this thesis is to develop a CCS technology assessment method, and

use it to make recommendations for an optimum carbon dioxide emissions reduction portfolio.

This general objective contains several more specific goals: 1) To identify the critical

parameters affecting commercial uptake and barriers to adoption of carbon capture and

sequestration technologies; 2) To develop a semi-quantitative methodology to assess qualitative

and quantitative data across various scientific and engineering disciplines for carbon

management decision-making/risk assessment; 3) To identify, compare and analyse the current

state and feasibility of carbon emission reduction technologies from a general global

perspective; 4) To rank carbon sequestration methods to better target scientific, industry and

governmental investment funds, policy, and research and development (R&D) efforts; 5) To

improve the understanding of, and differences between, carbon sequestration technologies and

the barriers to adoption for each, 6) To make recommendations for a diversified portfolio mix

for LNG operators on the North West Shelf of Australia.

1.3 THESIS ORGANISATION

This thesis consists of five chapters, including this Introduction. The second chapter covers

the development process of the semi-quantitative methodology which can be used to evaluate

and compare carbon sequestration technologies from multiple disciplines in a general or site-

specific manner. Examples are given to demonstrate how to apply the methodology. In the third

Introduction

14

chapter, I apply the methodology to Geosequestration. Chapter 4 evaluates two alternative

carbon sequestration technologies, Algae CCS and Biochar CCS. Chapters 3 and 4 include a

separate discussion and conclusion for each specific technology. The Discussion and

Conclusions Chapter provides a synthesis of the methodology and suggests directions for

further improvements. The three CCS technologies are compared to each other and conclusions

given. Portfolio recommendations are provided for LNG operators to consider when planning

their carbon management strategy and investments. References are provided at the end of each

chapter.

Certain content from my thesis has been presented at conferences and seminars, and is

included in the 2013 paper by Young-Lorenz and Lumley.

Introduction

15

REFERENCES

Abram, N.J., Mulvaney, R., Wolff, E.W., Triest, J., Kipfstuhl, S., Trusel., L.D., Vimeux, F., Fleet, L. and

Arrowsmith, C. (2013), Acceleration of snow melt in an Arctic Peninsula ice core during the twentieth

century, Nature Geoscience, 6, 5, 404-411.

Ahlenius, H. (2007), Historical trends in carbon dioxide concentrations and temperature, on a geological

and recent time scale: from Collection: Global Outlook for Ice and Snow, UNEP/GRID-Arendal Maps

and Graphics Library. http://maps.grida.no/go/graphic/historical-trends-in-carbon-dioxide-

concentrations-and-temperature-on-a-geological-and-recent-time-scale (accessed 7 May 2011).

Arrhenius, S. (1896), On the Influence of Carbonic Acid in the Air upon the Temperature of the Ground,

Philosophical Magazine and Journal of Science, 5, 41, 237–276.

Brito, L. and Stafford Smith, M. (2012), State of the Planet Declaration: New Knowledge Towards

Solutions, in Proceedings of the Planet Under Pressure, 26 – 29 March, 2012, London.

The Commonwealth of Australia (2012), The Clean Energy Legislative Package: The Commonwealth

of Australia. http://www.climatechange.gov.au/en/government/clean-energy-future/legislation.aspx

(accessed 12 August 2012).

Commonwealth of Australia (2013), The Angry Summer, [Steffan, W. (Climate Commission)],

Commonwealth of Australia, Department of Climate Change and Energy Efficiency, Canberra, ACT,

Australia.

Donner, S.D. and McDaniels, J. (2013), The influence of national temperature fluctuations on opinions

about climate change in the U.S. since 1990, Climactic Change, 118, 3-4, 537–550.

Exxon Mobil (2013), The Outlook for Energy: A View to 2040: Exxon Mobil Corporation, Texas, USA.

Forster, P., Ramaswamy, V., Artaxo, P., Berntsen, T., Betts, R., Fahey, D.W., Haywood, J., Lean, J.,

Lowe, D.C., G. Myhre, G., Nganga, J., Prinn, R., Raga, G., Schulz, M. and Van Dorland, R., (2007),

Changes in Atmospheric Constituents and in Radiative Forcing. In: Climate Change 2007:The

Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the

Intergovernmental Panel on ClimateChange [Solomon, S., D. Qin, M. Manning, Z. Chen, M.

Marquis, K.B. Averyt, M.Tignor and H.L. Miller (eds.)]. Cambridge University Press, Cambridge,

UK and New York, NY, USA.

Introduction

16

Geoscience Australia and Bureau of Resources and Energy Economics [BREE] (2012), Australian

Gas Resource Assessment 2012: [Bradshaw, M. and Hall, L. (Geoscience Australia), Copeland, A.,

and Hitchins, N. (BREE)], Canberra, ACT, Australia.

Geoscience Australia (1998), Well Report – The Summary: Evans Shoal 2: Canberra, ACT, Australia.

Global CCS Institute (2011), The Global Status of CCS: 2010, Global CCS Institute, Canberra, ACT,

Australia.

Global CCS Institute (2012), The Global status of CCS: 2012: Global CCS Institute, Canberra, ACT,

Australia.

Grubb, M. (2004), Technology Innovation and Climate Change Policy: An Overview of Issues and

Options: Keio Economic Studies, 41, 2, 103-132.

International Energy Agency [IEA] (2008a), CO2 capture and storage: A key carbon abatement option:

International Energy Agency, OECD, Paris, France.

Intergovernmental Panel on Climate Change [IPCC] (2005), Intergovernmental Panel on Climate

Change Special Report on Carbon Dioxide Capture and Storage: Prepared by Working Group III of

the Intergovernmental Panel on Climate Change [Metz, B.O., Davidson, H., de Coninck, C., Loos, M.

and Meyer, L.A. (eds.)]: Cambridge University Press, Cambridge, United Kingdom and New York,

NY, USA.

Intergovernmental Panel of Climate Change [IPCC] (2007a), Climate Change 2007: The Physical

Science Basis: Contribution of Working Group I to the Fourth Assessment Report of the

Intergovernmental Panel on Climate Change, [Solomon, S., Qin, D., Manning, M., Marquis, M.,

Averyt, K., Tignor, M. M. B., Miller Jr., H.L., and Chen, Z.], Cambridge University Press,

Cambridge, UK and New York, NY, USA.

Intergovernmental Panel on Climate Change [IPCC] (2007b), Climate Change 2007: Mitigation:

Contribution of Working Group III to the Fourth Assessment Report of the Intergovernmental Panel

on Climate Change, [Metz, B., Davidson, O., Bosch, P., Dave, R. and Meyer, L. (eds.)], Cambridge

University Press, Cambridge, UK and New York, NY, USA.

Intergovernmental Panel on Climate Change [IPCC] (2012), Summary for Policymakers. In:

Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation: A

Special Report of Working Groups I and II of the Intergovernmental Panel on Climate Change, [Field,

C.B., Barros, V., Stocker, T.F., Qin, D., Dokken, D.J., Ebi, K.L., Mastrandrea, M.D., Mach, K.J.,

Plattner, G.K., Allen, S.K., Tignor, M., and Midgley, P.M. (eds.)]. Cambridge University Press,

Cambridge, UK, and New York, NY, USA, 1-19.

Introduction

17

Koç, N., Njåstad, B., Armstrong, R., Corell, R.W., Jensen, D.D., Leslie, K.R., Rivera, A., Tandong, Y.

and Winther, J.G. (eds.) (2009), Melting Snow and Ice: A Call for Action, Centre for Ice, Climate and

Ecosystems, Norwegian Polar Institute, Polar Environmental Centre, Tromsø, Norway.

Larsen, P., Goldsmith, S., Smith, O., Wilson, M., Strzepek, K., Chinowsky, P. and Saylor, B. (2007),

Estimating Future Costs for Alaska Public Infrastructure at Risk to Climate Change: ISER Report,

Institute of Social and Economic Research, University of Alaska, Anchorage, AK, USA.

Lumley, D. (2010), 4D seismic monitoring of gas production and CO2 sequestration, North West

Australia: PESA News Resources, 104, 14-17.

Molde, A.I. (2011), Reservoaret klarer ikke mer Snøhvit-CO2: Petro.no: 20 May 2011.

http://www.petro.no/modules/module_123/proxy.asp?C=14&I=16697&D=2&mid=195 (accessed 26

May 2011).

Pacala, S., and Socolow, R. (2004), Stabilization Wedges: Solving the climate problem for the next 50

years with current technologies: Science, 305, 5686, 968-972.

Peters, G.P., Andrew, R.M., Boden, T., Canadell, J.G., Ciais, P., Le Quéré, C., Marland, G., Raupach,

M.R. and Wilson, C. (2012), The challenge to keep global warming below 2 [deg] C: Nature Climate

Change [opinion and comment][advance online publication], 1-3.

Plumer, B. (2012), When will the Arctic be ice-free in the summer? Maybe four years. Or 40: The

Washington Post, 20 September 2012.

http://www.washingtonpost.com/blogs/wonkblog/wp/2012/09/20/when-will-the-arctic-be-ice-free-

maybe-four-years-or-40/ (accessed 22 March 2013).

Press, T., Bergin, A. and Garnsey, E. (2013), Heavy weather: Climate and the Australian Defence Force:

Special Report Issue 49, Australia Strategic Policy Institute, Canberra, ACT, Australia.

Scott, K.C., Parker, D.J., Trupp, M., and Clulow, B. (2010), Setting New Environmental, Regulatory, and

Safety Benchmarks: The 2009 Gorgon CO2 3D Baseline Seismic Project, Barrow Island, Western

Australia: in Proceedings of the SPE Asia Pacific Oil and Gas Conference and Exhibition, edited,

Brisbane, Queensland, Australia.

Stern, N. (2007), The Stern Review: The Economics of Climate Change, Cambridge University Press,

Cambridge, UK.

Stewart, H., and Elliott, L., (2013) , Nicholas Stern: 'I got it wrong on climate change – it's far, far

worse': The Observer, 26 January 2013.

Introduction

18

http://www.guardian.co.uk/environment/2013/jan/27/nicholas-stern-climate-change-

davos?goback=.gde_67451_member_208430439 (accessed 28 January 2013).

Stott, P., Good, P., Jones, G., Gillett, N., and Hawkins, E. (2013), The upper end of climate model

temperature projections is inconsistent with past warming: Environmental Research Letters, 8, 1,

014024.

The Economist (2013), A sensitive matter: The Economist – Climate Science, [from the print edition], 30

March 2013. http://www.economist.com/news/science-and-technology/21574461-climate-may-be-

heating-up-less-response-greenhouse-gas-emissions (accessed 19 April 2013).

United Nations Environment Programme [UNEP] (2012), Policy Implications of Warming

Permafrost: [Report], United Nations Environment Programme, Nairobi, Kenya.

http://www.unep.org/pdf/permafrost.pdf (accessed 22 March 2013).

United Nations Framework Convention on Climate Change [UNFCCC] (2011), Establishment of an

Ad Hoc Working Group on the Durban Platform for Enhanced Action: Report of the Conference of

the Parties on its seventeenth session:United Nations Framework Convention on Climate Change,

CP/2011/9/Add.1. http://unfccc.int/resource/docs/2011/cop17/eng/09a01.pdf#page=2 (accessed 21

February 2013).

World Meteorological Organisation [WMO] (2011), Weather Extremes in a Changing Climate:

Hindsight on Foresight; World Meteorological Organisation, WMO-1075.

World Meteorological Organisation [WMO] (2012a), Greenhouse Gas Bulletin: The State of

Greenhouse Gases in the Atmosphere Based on Global Observations through 2011, World

Meteorological Organisation, No. 8.

World Meteorological Organisation [WMO] (2012b), Provisional Statement on the State of Global

Climate in 2012: presented at the 18th Conference of the Parties to the United Nations Framework

Convention on Climate Change (UNFCCC).

World Meteorological Organisation [WMO] (2012c), 2012: Record Arctic Sea Ice Melt, Multiple

Extremes and High Temperatures, Press Release No. 966, 18th Conference of the Parties to the

United Nations Framework Convention on Climate Change.

Young-Lorenz, J.D. and Lumley, D. (2013), Portfolio analysis of carbon sequestration technologies and

barriers to adoption: General methodology and application to geological storage: Energy Procedia, 37,

5063-5079.

Methodology Development

19

CHAPTER 2

METHODOLOGY DEVELOPMENT

OVERVIEW

The effective targeting of investment funds and research efforts to reduce industrial carbon

dioxide emissions, while still preserving access to fossil fuel energy resources, requires

quantitative methods to assess and compare a diverse range of carbon capture and sequestration

(CCS) technologies from a global perspective. I develop a quantitative methodology comprised

of multiple components to assess the feasibility of CCS technologies. The methodology aims to

identify and characterise the most critical evaluation criteria affecting the wide-scale adoption of

carbon sequestration technologies. It is designed for use in the early stages of the CCS portfolio

planning process therefore a matrix scorecard approach was chosen for ease of use. It combines

both quantitative and qualitative measures into numeric scores to facilitate comparison. The

initial design and calibration is performed using the CCS technology of Geosequestration.

Flexibility in the design of the methodology framework allows for review and modification to

account for technology progress and increased amounts of public domain data.

Interpretation of the results produced through application of the methodology highlights both

positive aspects, and more problematic areas which can be addressed to improve ratings, and

provides a quick overview for a CCS investment portfolio analysis. The methodology can also

be used to quickly re-evaluate the impact of recent events on a single project or project

portfolio. This tool provides a multi-criteria feasibility analysis which can be used by

governments and industry to rank carbon sequestration options. It can be applied across

projects, scientific disciplines and technologies (portfolio analysis) and with some modification,

within a specific technology or project (project analysis).

Methodology Development

20

Methodology Development

21

2.1 INTRODUCTION

Many people in society desire to mitigate the risks of climate change by reducing industrial

carbon dioxide (CO2) emissions while preserving access to fossil fuel energy sources. Industrial

carbon emitters are faced with increasing regulation and carbon tax or debt while the available

carbon sequestration options face significant challenges and uncertainties. To facilitate the most

effective targeting of CCS investment funds and research efforts, both governments and

industry need quantitative methods to assess and compare CCS technologies in order to rank,

prioritise and make sound policy and investment decisions. Proposed CCS methods for the

stabilisation of atmospheric CO2 levels need to be analysed for project risks across a range of

factors, from the socio-economic to the purely technical. No easy-to-use reference methodology

to review or analyse CCS technologies was identified in a public domain literature search. The

goal of this thesis is to develop such a methodology in a quantitative format.

I develop a semi-quantitative methodology framework and matrix scorecard rating system

utilising a strategy similar to an “inversion by forward modelling” (Boschetti and Moresi, 2001)

or manual inversion approach. The methodology under development is used to create a

prediction (synthetic dataset) about the probability of success of a given technology and the

result is compared to an expectation scenario based on case studies or current understanding of a

technology. If differences are present, a manual update of the methodology is performed after

analysis of the causes of the differences. In order to develop a methodology to rate and rank the

most feasible options, I define the following objectives:

1. Identification of common critical parameters affecting the uptake of CCS technologies

at commercial scales

2. Identification of common measures or metrics applicable to evaluate CCS technologies

from a variety of scientific and engineering disciplines

3. Ease-of-use and reference

4. Flexibility for updating of the methodology

5. Application across scientific disciplines and technologies (scope - portfolio analysis)

6. Adaptable for application across projects (scope - project analysis within a CCS

technology)

Achieving these objectives requires an iterative process with ongoing research and review,

with possible updates to the model and methodology components, due to the rapidly changing

development cycle stages and diversity of CCS technologies. The iterative process is limited to

the updating of the model to ensure relevance and applicability for all CCS technologies.

Methodology Development

22

2.2 MULTI-CRITERIA METHODOLOGY DEVELOPMENT

I have developed this technique by conducting an extensive review of the literature and

project case studies, gathering data, developing a model and rating system in an iterative

manner, and providing a new method of analysis. I identified and grouped the parameters

affecting the wide-scale adoption of CCS technologies into six critical evaluation criteria (Table

2.1): 1) public acceptance, 2) regulatory frameworks, 3) economics, 4) science and technology,

5) storage quality, and 6) environmental impact.

These key evaluation criteria are selected because they have the most impact on whether a

carbon sequestration project or technology will proceed, suffer delays, or be implemented at a

commercial and widespread scale. They must also be general enough to be applicable to all

possible CCS technologies. For instance, as the goal of the methodology is to assess and

compare divergent technologies, during the iterative process to determine critical evaluation

criteria relevant to all CCS technologies, a criterion such as long-term liability would not

qualify as such, because this specific challenge is limited to Geosequestration. This particular

issue would still be analysed under the evaluation criteria of Economics and Regulatory

Frameworks. In this section I briefly define the evaluation criteria, and in later sections discuss

and provide case studies for illustration as to how to use the rating system and assign scores.

Table 2.1. The identified evaluation criteria affecting widescale adoption of carbon sequestration

technologies.

Methodology Development

23

Criterion 1: Public Acceptance

“Public Acceptance” refers to public, political and other stakeholder acceptance and is

increasingly seen by governments and industry as crucial to CCS technology adoption. This is

one of the main reasons for demonstration projects, such as the Mid-West Midwest Regional

Carbon Sequestration Project in Greenville, Ohio, USA, being cancelled at late stages after

years of investment from multiple stakeholders (Kessler, 2009). Stakeholders for carbon

sequestration technologies cover a diverse range of individuals and organisations, ranging from

the business community to individual property owners. Some examples of stakeholders are

given in Table 2.2.

Table 2.2. Some examples of stakeholder diversity under the “Public Acceptance” criterion for carbon

sequestration technologies and projects. This list is not exhaustive but included to provide a general idea

of the broad range of individuals and organisations affected.

Criterion 2: Regulatory Frameworks

“Regulatory Frameworks” describes a system of rules and the means of enforcement that can

favour or impede the uptake of CCS technologies. They are established by local, state and

federal governments as well as international bodies to regulate a specific activity and range from

the ease of land access, to policies for monitoring and remediation. Table 2.3 lists some

associated issues included under Regulatory Frameworks.

Methodology Development

24

Table 2.3. A range of associated issues related to “Regulatory Frameworks” which affect the uptake of

carbon sequestration technologies.

Criterion 3: Economics

“Economics” refers to the economic and financial aspects affecting CCS technologies,

projects and investment (Table 2.4). Individual projects must be profitable to provide the

necessary capital expenditure (CAPEX) and ongoing operational expenditure (OPEX). This can

be achieved as a stand-alone business, by trading in carbon credits, or through large scale

government incentives.

Criterion 4: Science and Technology

“Science and Technology” refers to the available knowledge and experience in research,

design and development (R&D) activities, and the ability to demonstrate the scientific concepts

and engineering solutions at different scales in the lab, and in the field for demonstration and

commercial scale projects. It includes the understanding of the impact of a CCS technology in

mitigating atmospheric CO2 associated with the implementation of such a technology.

Methodology Development

25

Table 2.4. Economic and financial factors affecting the adoption of carbon sequestration technologies,

projects and investment.

Criterion 5: Storage Quality

“Storage Quality” is defined as the capacity and quality of space available for secure storage

of CO2. It covers the sequestration rate, volume, scalability, security, and length of time the CO2

remains sequestered. If a CCS technology or project is not capable of providing minimum

volumes of safe, long-term storage, the commercial viability will be decreased. This will reduce

the priority of the project for governments and also affect internal resource allocation by

industry. In addition, projects which cannot refer to minimal required volumes will be more

susceptible to negative feedback from stakeholders as the cost benefit evaluation becomes

negative. All these factors results in significant project risk.

Criterion 6: Environmental Impact

“Environmental Impact” covers positive or negative environmental impacts of a CCS

technology or project. It includes changes to eco- and water systems, the environmental

footprint due to infrastructure health and safety, and food, water and energy sources and their

Methodology Development

26

security. These impacts may be restricted to the immediate vicinity, or could impact locations at

significant distances from the source CCS technology activity. The concerns about any CS

technology are whether any side effects are unanticipated or could cause irreversible damage to

the environment. Baseline measurements, appropriate modelling, measurement monitoring and

verification (MMV), and remediation measures must be in place to predict and counter such

negative effects should they occur.

2.3 MODEL DEVELOPMENT PROCESS

The methodology aims to provide an assessment tool for a wide range of technologies, so it

has to account for the differences between the technologies without rendering the comparison

process completely arbitrary. To achieve this goal, an iterative approach is used to update the

methodology model. If differences are present between the expectation or reference data set and

the predicted outcome of the methodology application, updates to the model may include

redefinition of a sub-criterion or revision of sub-criteria measures. This is necessary as only one

well documented technology was available to calibrate the model against commercial scale case

studies, Geosequestration. All the other technologies dealt with an expectation scenario which

was derived from various sources but could not be ground truthed against the actual

performance of commercial scale deployments.

The general sequence of the methodology development first identified the critical parameters

(evaluation criteria) across which all technologies should be compared (as introduced in Section

2.2). The criteria and sub-criteria are then defined for which the given technology should be

ranked. The latter stage also comprises definitions of the cut-off points for the ranking.

Wherever possible, quantitative cut-off measures are preferred to qualitative measures. After

this procedure was performed for Geosequestration, and reflected the current status as given by

case studies and other data, the resulting outcome was applied to the next CCS technology. If

this application did not provide the indicated results as documented in the available literature

(the expectation scenarios) for this or any of the following technologies assessed, the process

was iterated. In this iteration, the evaluation criteria and sub-criteria were modified and the

initial calibration against Geosequestration repeated. Passing this stage, the modified evaluation

criteria and sub-criteria, were then applied to Algae CCS and Biochar CCS. Any further

mismatch between expectation scenarios for technologies and the assessment outcome resulted

in a new iteration.

Methodology Development

27

Figure 2.1 shows a simplified chart of the process which is discussed in this section. To

calibrate this process and define reference points, Geosequestration is used as the initial CCS

technology due to its more advanced stage of development, and associated larger volume of data

and case studies available in comparison to other CCS technologies under evaluation. The steps

to develop the methodology model are as follows:

Step 1: Identify evaluation criteria

From the literature review of a range of CCS technologies, the evaluation criteria

affecting technology adoption are identified and defined. These criteria form the

basis across which all technologies in the Master Matrix (Figure 2.2) and the

Science and Technology Sub-Matrix are scored (Figure 2.3) and should be

applicable to all CCS technologies to ease comparison. These criteria are used to

define general cut-off points on a scale from 1 to 3, and used as the basis of the

rating system.

Step 2: Define Science and Technology Sub-Matrix

The scientific and engineering elements specific to a CCS technology are identified

and defined. The initial Science and Technology Sub-Matrix is calibrated with

Geosequestration which provides the model technology elements which the

Science and Technology Sub-Matrices for all other technologies should ideally

replicate to allow for comparison. Where this is not feasible because a technology

element proves to be irrelevant, e.g. injectivity for Biochar CCS, it is substituted

with a relevant technology element of equivalent purpose or importance. Under the

Sub-Matrix, the evaluation criterion for Science and Technology is named

Demonstrated/Ready to better reflect the characteristic being measured.

Step 3: Establish ranking system and sub-criteria

Identify the specific questions or issues to evaluate the technology’s status for the

specific sub-criteria. Each is scored on a scale of 1 to 3, and the

compilation/average of all answers form the numerical value for the specific

technology evaluation criterion. Relevant cut-off points for the CCS technologies

are introduced.

Step 4: Develop reference dataset/benchmark

Derive a reference data set against which the outcome of the ranking process

should be compared. In the case of Geosequestration, case studies were used to

Methodology Development

28

define an expectation scenario. For later steps when assessing CCS technologies

which have not left the laboratory stage, available data is used to extrapolate an

expectation outcome which should be reflected by the ranking process.

Step 5: Apply ranking system

The ranking system is applied and the derived numerical values for both Science

and Technology Sub-Matrix and Master Matrix values are compiled. These values

are then compared to the reference data set which is provided by the expectation

scenario or extrapolated expectation scenario. I evaluate how well the quantitative

assessment of the methodology in the form of the derived combination of criteria

and sub-criteria (test data) is reflected by, or matches, the reference dataset.

Step 6: Test dataset matches reference dataset?

The test dataset is calibrated with the benchmark. The evaluation of the outcome of

the comparison is not restricted purely to “match”/ “does not match”. If the

reference dataset matches the analysis outcome through the application of the

ranking system, the next technology will be evaluated. If this is not the case, the

“no match” outcome is present, a form of reality check is introduced before the

result is discarded. This point has proven necessary because of the lack of, or very

limited data and case studies for most CCS technologies. To account for this, an in-

depth analysis of the results of the ranking methodology is performed to ensure that

the critical points which cause the disconnect between the expectations scenario

and ranking are not due to overly optimistic or pessimistic assumptions in the

reference data set.

Step 6a: Test with next CCS technology

If the ranking is successfully applied to a specific CCS technology, i.e.

Geosequestration as the initial and calibrating entity, the next CCS technology will

be evaluated (from Step 1 onwards). The evaluation criteria and sub-criteria are

reviewed for applicability. If any of the criteria or sub-criteria are proven not to be

applicable to this new technology, the whole process is reinitiated, and iterated,

starting with Geosequestration, and the whole Matrix scoring system is updated

until it is relevant to the next technology being evaluated.

Methodology Development

29

Step 6b: Reality check

If the methodology outcome and reference dataset do not match, the reality check

is performed. It aims to evaluate if the mismatch could be connected to a bias in the

reference dataset, e.g. only considering “safe projects,” too small a data subset, or

ignoring certain subsets of factors impacting technology adoption.

If the methodology outcome is regarded as realistic, the reference dataset will be

updated.

If the methodology outcome is regarded as unrealistic, then the model development

process must be reviewed. If the technology under consideration is

Geosequestration, an update of the criteria and sub-criteria is performed. If any

other technology is evaluated, the model development process is reiterated to

ensure the ranking system is applicable for this technology. The decision to

perform a complete re-start of the process, going back to Geosequestration, was

made because the methodology should be general, i.e. applicable to all

technologies.

Step 7: Assess next CCS technology

Perform the evaluation of the next technology from Step 1 onwards.

Fig 2.1. Methodology Development Model describes the flow of the methodology as an iterative process.

(S/T = science and technology)

Methodology Development

30

Figure 2.2. Master Matrix Step 1: Identification of evaluation criteria, across which CCS technologies

and technical elements are scored in the Master Matrix and Science and Technology Sub-Matrix,

respectively.

Figure 2.3. Science and Technology Sub-Matrix.

2.4 FRAMEWORK COMPONENTS

Following the identification and grouping of the evaluation criteria, the results of my

research have led to the development of several methodology framework components which

operate together to comprise the whole ranking system. In addition to the iterative process (as

discussed in 2.3), I develop a ratings system to generate ratings from quantitative and qualitative

data, and a matrix scorecard framework that allows for ease of reference and application. In

developing a tool which can be easily utilised and referenced, a matrix scorecard format was

chosen because this framework is quickly accessed and understood. It also generates

quantitative numerical scores from lengthy qualitative data assessments thereby allowing for a

quick measure to compare projects, similar to the approach used to analyse 4D seismic project

feasibility (Lumley et al., 1997).

Methodology Development

31

Master Matrix Scorecard

The Master Matrix Scorecard acts as the primary portfolio analysis comparison tool of the

different CCS technologies. Each technology is ranked for each of the evaluation criterion, as

the same issues affect the adoption of all CCS methods. Behind the scenes is the Master Matrix

Spreadsheet Template where the calculations are made (Figure 2.4), averaging the total of the

six parameters to generate an overall risk rating for each of the analysed CCS technologies.

Figure 2.4. Master Matrix Spreadsheet Template

Sub-Matrix Scorecard (Science and Technology Sub-Analysis)

The Science and Technology evaluation criterion has an additional in-depth assessment in

the form of a Sub-Matrix because of the broad diversity of technologies which are considered

for carbon sequestration. This Sub-Matrix contains a set of scientific and technical elements,

unique for each of the CCS technology options. For example, Figure 2.5 illustrates the technical

elements for Geosequestration. While it would be beneficial to align the Sub-Matrix as much as

possible with the analysed carbon sequestration technologies for ease of comparison, this will

not always succeed if the CCS technology portfolio is too diverse.

Figure 2.5. Science and Technology Sub-Matrix Spreadsheet Template for Geosequestration

Methodology Development

32

Each technology element is assessed across the evaluation criteria with the exception of the

“Storage Quality” criterion as it is not applicable to all elements (but is important to all CCS

technologies). The Science and Technology column of the Sub-Matrix is defined as

“Demonstrated Technical Readiness” to minimise confusion and characterise the criterion. The

overall parameter score is taken from the average column and input into the Science and

Technology column in the Master Matrix in order to provide the average Science and

Technology rating of the CCS technology as a whole (Figure 2.6 below).

In addition, the technical elements can score very differently across the evaluation criteria.

For example, in the case of Geosequestration, “Public Acceptance” of “Injection and Storage” is

ranked poorly due to community concerns about issues such as storage site integrity and loss in

property values. However, public perception of the “Capture” component is rated highly

because it presents no perceived risk and beneficially prevents industrial generated CO2 from

entering the atmosphere. The Sub-matrix analysis identifies these types of differences within a

respective CCS option as well as for specific projects. It highlights areas which require a more

in-depth analysis to define focus points which are critical to the technology rating. These areas

will form the focus for development of most efficient strategies for potential upgrading of that

technical aspect. This will thus improve the overall rating of that CCS technology.

Consequently, use of the Sub-matrix emphasises the importance of the different technical

aspects for each of the CCS methodologies.

Evaluation Rating Scale

A simple numerical ranking value is used to generate quantitative data, facilitate ease of use

and align with the reference table character of the methodology (Table 2.5). Each CCS

technology in the Master Matrix, and technical element in the Sub-Matrix, is assigned a rating

across the evaluation criteria on a scale of 1 - 3, with 1 being the lowest rating and 3 the highest.

The cut-off points between low and medium are at 1.5, medium between 1.5 and 2.5, and high

above 2.5. For the general methodology development, all criteria have equal weighting. A

subjective weighting scheme could be introduced if the conditions require it, for example, in

densely populated areas like Western Europe, public acceptance may receive a higher weighting

while in sparsely populated areas, transportation may be more critical.

Methodology Development

33

Table 2.5. Evaluation Criteria Rating Scale and Equivalents

The schematic shown in Figure 2.6 summarises and illustrates the link between the three

discussed components of the methodology, and the interaction between the Master Matrix

Scorecard and Science and Technology Sub-Matrix. The average score of the technical elements

assessed in the Science and Technology Sub-Matrix is transferred into the Science and

Technology column of the Master Matrix. The overall technology risk assessment rating is

shown in the Technology Rating column.

Figure 2.6. Schematic showing how the Master Matrix Scorecard, Sub-matrix Scorecard, and Rating

Scale work in relation to each other.

Assessment Grade Numerical Average Range Equivalent

Scorecard Symbol Equivalent

Low <1.5 X

Medium >1.5 to <2.5 -

High >2.5 √

Methodology Development

34

Methodology Development

35

2.5 RANKING SYSTEM

This section first describes the Portfolio Analysis Taxonomy (Figure 2.7) which provides an

overview of the overall methodology structure, then defines the Evaluation Criteria (Table 2.6)

and Sub-Criteria Scorecards (Table 2.7) for the purposes of portfolio or project analysis.

Portfolio Analysis Taxonomy

The Portfolio Analysis Taxonomy (Figure 2.7) illustrates the overall methodology structure

(Young-Lorenz and Lumley, 2013). The six evaluation criteria, discussed in Section 2.2, are

defined in the Evaluation Criteria Scorecard (Table 2.6) and used to assign quantitative scores.

The sub-criteria provide the specific questions or issues to assess each evaluation criteria (Table

2.7) and represent the characteristics, used to establish the numerical score. Quantitative

measures, case studies, reports and other data are used to guide ranking and calibrate scores.

The average ratings of all sub-criteria provide the respective evaluation criteria score, while the

average of all evaluation criteria ratings provides the overall technology rating. The technical

readiness of the given technology is analysed in a separate sub-matrix to take into account the

technology elements specific to each CCS technology.

Figure 2.7. Portfolio Analysis Taxonomy showing elements (evaluation criteria and sub-criteria)

assessed in my methodology for each carbon sequestration technology in order to perform a portfolio

evaluation.

Methodology Development

36

Evaluation Criteria Scorecard

Each of the evaluation criteria is kept general to ease comparison across technologies. The

general interpretation of the rating for each criterion is provided in Table 2.6. Cut-off points for

the score assignment are preferably quantitative versus qualitative to enforce applicability of the

methodology across technologies and limit the influence of subjective judgment.

Table 2.6. Rating System Evaluation Criteria Scorecard illustrating the rating interpretation for each

criterion. This Scorecard works together with the Rating System Sub-criteria Scorecard to provide a

semi-quantitative analysis for each CCS technology.

2.5.1 Sub-Criteria Scorecard

The sub-criteria provide the specific questions or issues which must be answered under each

parameter criterion in order to evaluate the technology. They represent the characteristics used

to generate quantitative data (numerical scores) (Table 2.7), with quantitative measures, case

studies, reports and other data used to guide ranking and cut-off points. Likewise, when

applying the methodology, the assessor determines the overall rating for each sub-criterion

based on case studies and other relevant data. The compilation of all the sub-criterion scores

represents the criterion score.

Methodology Development

37

Table 2.7. Rating System Sub-Criteria Scorecard with measures to establish ratings.

Public Acceptance

The definition of “Public Acceptance” refers to public, political and other stakeholder

acceptance, and is often referred to as “social licence to operate.” Therefore, I analysed the

reported reasons for a lack of public acceptance in various case studies in order to determine

suitable sub-criteria.

How does the CCS method affect me?

This question refers to the impact on the individual, and entails how much the project is

likely to cause a “not in my backyard” (NIMBY) reaction related to concerns about property

devaluation, economic impacts, or health, safety and environmental risks (Kessler, 2009;

Kuijper, 2011; Schlumberger Business Consulting Energy Institute [SBC Energy Institute],

2013; Reuters, 2011). For example, an Australian household may appreciate the potential

benefits of adopting CCS technologies for climate change mitigation, but object to rising

Methodology Development

38

electricity prices (Farr, 2012). The Sleipner Project in the North Sea (Reiner, 2008) would

represent a high scoring example due to its remote location offshore and no direct impact on

individuals; while the now cancelled Barendrecht (Netherlands) onshore project in a densely

populated area would represent a low scoring example (Kuijper, 2011; IEA GHG, 2009).

How natural is the CCS method?

A CCS technology, closely resembling a natural process in nature’s carbon cycle, is more

likely to receive a higher level of public acceptance from the local community and

environmental groups. Reforestation would represent a high scoring project as it is considered a

reestablishment of a natural balance, enhancing the natural carbon cycle, and a “safe” carbon

mitigation option (Silver et al., 2010: Ipsos-MORI, 2010); whereas Ocean Fertilisation would

be a low scoring CCS technology due to risks of disruption of the natural balance of marine life

systems on a global scale (IEA GHG, 2002).

Science and technology comprehension

An educated public with a solid comprehension of the scientific and engineering concepts

underpinning a CCS technology is less influenced by misinformation and special interest

groups. In addition, educated and engaged stakeholders are more likely to get involved and take

ownership of a project while sidelined stakeholders are more likely to resist a technology’s

adoption (Miller et al., 2008). The active involvement and communications of the Otway

Project with the public would represent a high score (Ashworth et al., 2010); while the

inadequate communications, education and engagement with the local community that resulted

in unjustified fears about the Barendrecht Project and its ultimate cancellation would represent a

low score (Kuijper, 2011; IEA GHG, 2009).

Methodology Development

39

Regulatory Frameworks

“Regulatory Frameworks” refers to a system of rules, regulations and legislation, and the

means to enforce them. These rules are established by state and federal governments as well as

international bodies and range from comprehensive frameworks to rules for very specific

activities.

Enabling regulation

The existence of enabling legislation or flexibility in the adaptation of existing rules to new

requirements, will favour the uptake of CCS technologies. In 1995, Porter and van der Linde

advocated the view that effective environmental legislation has positive effects on resource

productivity and innovation thus leading to competitive advantage. Enabling regulation could be

in the form of funding incentives such as the European Union’s New Entrance Reserve (NER

300) programme; or taxation schemes, like the Commonwealth of Australia’s Clean Energy

Legislative Package and Carbon Pollution Reduction Scheme. Regulatory requirements which

encourage industry compliance can positively affect adoption of a CCS technology as they

create a base need for research and development as well as deployment of a technology. High

scores are given to flexible regulators, like the amendment of existing pipeline legislation

(Western Australian Petroleum Pipelines Act 1969) to allow for CO2 transportation by the State

Government of Western Australia (The Parliament of Western Australia, 2011), and the

required retrofitting of CCS equipment to coal-fired power plants, as in the United Kingdom

(UK) (Department of Energy and Climate Change [DECC], 2010). A low scoring example

would be the absence or retraction of planned legislation, or the passing of politically

controversial regulations which are dependent on single party support.

Restrictive or negative regulation

The existence of restrictive or negative legislation, or a comprehensive but negative

regulatory framework, can negatively impact the adoption of certain technical aspects of CCS

technologies. A high scoring example may be a country with no negative or restrictive

regulation. A low scoring scenario may be a government banning access to a significant portion

of technically and economically suitable storage sites, such as in the Netherlands (ICIS Heren,

2011).

Methodology Development

40

Regulatory certainty

Regulatory certainty refers to the presence or absence of a comprehensive regulatory

framework. High policy uncertainty can result in a significant increase of risk and planning

uncertainty (costs, liability and project delays), which can motivate commercial enterprises to

delay or cancel projects and thus further hinder the adoption of a CCS technology (Bachu,

2008). A high score is given for the existence of either positive or negative comprehensive

legislation, because this provides the ability to conduct economic analysis and planning with a

significant degree of certainty. In the USA, some corporations and industries are pushing for

unified federal and state legislation while setting voluntary greenhouse gas (GHG) reduction

targets for near-term strategic advantage as well as preparing for future mandatory requirements

(Hoffman, 2005). A low score is assigned where the absence of a regulatory framework yields

high risk for a project since it is difficult to plan in the face of regulatory uncertainty. For

example, the delay in enactment of proposed CCS legislation by the German government,

contributed to the decision of Swedish power utility Vattenfall to cancel the Jänschwalde project

in Brandenburg (Helseth, 2011; Reuters, 2011).

Economics

“Economics” refers to the economic and financial aspects affecting CCS technologies,

projects and investment.

Viable business case

Significant funds are required to invest in CCS technologies (CAPEX, OPEX, monitoring,

regulatory compliance and carbon accounting) especially at the early adopter stage, and

operators need these to be offset by consistent revenue streams and predictable, sustainable

business models. Regulatory requirements which enforce industry compliance act to define both

corporate strategy and cost, not only in the initial stages for CAPEX but also in the long-term

with OPEX and MMV expenses. A high scoring example may be the presence of sufficient

carbon tax savings through the use of CCS technology as is the case at Sleipner (Torp and

Brown, 2002). A low scoring example would be a project which is dependent on a highly

uncertain revenue stream, especially to cover OPEX. Dependence on direct government funding

of OPEX, or dependency on floating CO2 surcharges to electricity prices would be additional

illustrations of low scoring cases.

Methodology Development

41

Government Funding/Incentives

This sub-criterion refers to the level of dependence upon government funding or other

incentives. A high scoring example would be a project that is not reliant upon such subsidies to

encourage the adoption of the CCS technology. Enhanced oil recovery (EOR) projects with

sufficient increased oil production due to CO2 injection are currently one of the few profitable

examples which can achieve a high score (SBC Energy Institute, 2012). If a project is not a

stand-alone viable business, then a heavy reliance on government incentives or intervention

make a project susceptible to a business strategy review or regulatory change. This is illustrated

by the Longannet CCS project in Fife, Scotland, after the takeover of Scottish Power by

Iberdrola. This project was cancelled after the UK Government would not increase the funding

support and the project partners would not agree to cover the additional GB₤520,000 risk,

especially as coal-fired power plants did not fit with Iberdrola’s business focus which is on

renewable energy (Iberdrola Group, 2013; Haszeldine, 2011; BBC Online 2011a: BBC Online,

2011b).

A high reliance on subsidies or incentives would score a low rating, such as a project which

cannot proceed without significant government financial intervention. For example, plans to

develop the South West CO2 Sequestration Hub project in Western Australia can proceed only

with the financial support under the Commonwealth of Australia’s Carbon Capture and Storage

Flagships programme (Government of Western Australia, 2011). This research and

demonstration project brings together government and large-scale CO2 emitters from different

industries to capture and inject CO2 into an onshore storage site and thereby build crucial

experience and capabilities (Stalker et al., 2013). Financial support from both the State and

Federal Governments encourages participation by industry by limiting the financial risk of the

industry partners, and enables this pilot project to demonstrate the commercial-scale feasibility

of CCS technologies in Australia (Zero Emission Resource Organisation [ZERO], 2013).

Carbon Market

This sub-criterion is defined by the level of sensitivity to the carbon market. The carbon

market is still immature with a decreased international price of carbon (Figure 2.8), which is

predicted to stay low until 2020 as a result of economic weakness in Europe and an oversupply

of emissions abatement permits (Maher, 2012). A high score would be given to a business

model which does not rely upon trade or sale of carbon credits. A high sensitivity to trade

Methodology Development

42

volumes and carbon price is a disadvantage and would thus merit a low score as the market is

currently artificial and dependent on political will and cooperation.

Figure 2.8. Carbon prices over time in Europe for the period April 2010 to April 2012 (Image source:

van Renssen, 2012).

Science and Technology

“Science and Technology” is defined as the capability and efficacy of a CCS technology in

mitigating atmospheric CO2, and understanding of the associated environmental risks, with a

basis in sound scientific concepts and engineering solutions. Technical Readiness and

Demonstrated Evidence are both required for industrial scale adoption as no economic

assessment is feasible without knowledge of these sub-criteria. If risks are present, their impact

has to be quantifiable, and means for remediation available.

Technical Readiness

Technical readiness refers to the status of a CCS application, that is, whether it needs more

research and development, or is ready for demonstration, or is already available at commercial

Methodology Development

43

scale. The cut-off points approximate the adoption times of those of the Gartner S-curve

technology development life cycle, as illustrated in the Evaluation Criteria Scorecard (Table

2.6), with timelines from concept to market (Fenn et al., 2009). Technical readiness can also be

illustrated by the complementary Grubb curve which outlines the initial increase then decrease

in costs and effort as induced endogenous technological changes and learning progress beyond

the emerging stage (Grubb et al. 2002). The lower the level of Technical Readiness of a CCS

technology, the higher the economic and implementation risks. Public awareness and

acceptance are also impacted by whether a technology is immature or perceived as novel and

untested. One method of assessing the level of technical readiness of a CCS solution is the

availability of commercial products and tools, such as software and services, which are already

used by industry for this or similar purposes. The adaptation of CO2 capture technologies used

by the oil and gas (O&G) industry for gas processing for utilisation by other industries, such as

coal-fired power generators, would be an example of a high scoring case. Another high score

illustration would be the utilisation or modification of knowledge and tools for CCS purposes

from the O&G industry which has over forty years of knowledge, experience and tools gained

from safely injecting CO2 into oil and gas reservoirs to enhance hydrocarbon recovery. A low

scoring example would be a technology like Algae CCS with significant development needs to

ascertain the most suitable microalgae strains and achieve efficient high volume production

levels (Global CCS Institute, 2011).

Demonstrated Evidence

Demonstrated evidence refers to the established nature of scientific concepts, engineering

solutions, benefits and safety for each technical element specific to a given CCS option both in

the lab and field. A skilled labour force is also needed to implement and prove the benefits and

safety of all technical elements for a particular CCS option through measurement, monitoring

and verification (MMV) activities, management and remedial plans. A high scoring example

would be Geosequestration, which has been safely demonstrated for over forty years of EOR for

CCS purposes via numerous large-scale demonstration projects (Eiken et al., 2004, 2011;

Hovorka et al., 2011). A low scoring example would be an immature or unproven scientific

concept or technical solution, such as the deposition of CO2 in solid phase on the deep ocean

floor (Brewer et al., 1999), which may work in theory, however field trials are unlikely due to

the inability to qualify and quantify the risks associated with it.

Methodology Development

44

Storage Quality

“Storage Quality” is defined as the capacity and the quality of the space available for long-

term secure storage of CO2. This sub-criteria reflect the standards proposed by international

agencies for the qualification of storage sites by considering capacity, injectivity, containment,

and monitoring potential (Bachu, 2000; Bradshaw et al. 2007; Det Norske Veritas [DNV],

2012a, 2012b; Aarnes et al., 2009).

Capacity

Capacity refers to the total volume of CO2 which can be stored and scaleability of the

technology. A high score example would be given to a massive deep saline formation with an

assessed storage volume capable of supporting a number of major commercial CCS projects, for

example 10+ GtC, such as the Mount Simon Sandstone (North American Carbon Storage Atlas

[NACSA], 2012). A low scoring example would be a very localised storage site, such as a small

depleted hydrocarbon reservoir, or an unreasonably large dependency on areal space

requirements such as that required for algae ponds (Global CCS Institute, 2011; Benneman,

2003).

Rate

Rate is the volume of CO2 that can be stored per unit time, or the speed at which CO2 can be

sequestered. A high scoring example would be the ability to store CO2 at the same rate as which

the CO2 is captured or produced, such as Geosequestration is capable of doing, for example at

the Sleipner Project where more than 1 MtC have been stored each year for 15+ years (Eiken et

al., 2011). A low scoring example could be Reforestation since the rate of CO2 storage is

constrained by the relatively slow growth of trees. The average carbon sequestration rate

estimate per unit of tree cover in the USA is 0.277 kilograms carbon per square metre per year

(kg m2pa) and the net sequestration rate 0.205 kg m

2pa accounting for the decomposition of

dead trees (Nowak et al., 2013).

Duration

Duration refers to the length of secure CO2 storage in time, and is therefore related to

containment and leakage issues. A high scoring example may Biochar CCS as radiocarbon

dating of the Amazonian Terra Preta soils measured the mean residence time (MRT) of carbon

Methodology Development

45

as ranging from 500-7000 years (Neves et al., 2003). A low scoring example may be Algae

CCS, as the CO2 is only stored temporarily during the growth cycle, which ranges from days to

months depending on climatic conditions and algae strain, until the algae biomass either

decomposes (Jewell and McCarty, 1971) or is harvested and converted into the desired end-

product, e.g. biofuels.

Verification

Verification is the ability to monitor the sequestered CO2 for storage behaviour,

environmental interactions, containment, leakage and economic auditing. A high scoring

example would be Geosequestration due to the commercial availability of geophysical methods

and equipment used by the energy industry for exploration, production and enhanced

hydrocarbon recovery (Hovorka et al. 2011; Hill et al. 2013). A low scoring example may be

Afforestation, which is the establishment of a forest in a location where there was none

previously, because it is difficult to access and monitor large tracts of land to check if the

planted trees are still standing, and to accurately account for the carbon they are sequestering.

Environmental Impact

“Environmental Impact” relates to the possible positive or negative environmental impacts

that a CCS technology or project may have locally, regionally or globally. Other issues related

to this criterion are health and safety, damage to resources, competition for groundwater and

land resources, such as agriculture versus biomass cultivation, as well as food and energy

sources and security.

Potential impact on ecosystems and water systems

Impacts include any potential side-effects or changes experienced in eco- and water systems

and may even occur thousands of kilometres away from the project site. A high scoring example

would be Geosequestration, where any CO2 leakage is likely to be small and constrained to the

immediate site vicinity because of the rigorous process of site characterisation and monitoring

(Bachu, 2000; CSA Group and IPAC-CO2 Research Incorporated [CSA IPAC-CO2], 2012).

Ocean Fertilisation may be a low scoring example, since the effects of micronutrients at

commercial scales on marine systems and tidal effects are unknown, and could impact coastal

areas and food sources tens of thousands of kilometres from the site of fertilisation. An analogue

for the global footprint of ocean fertilisation could be found in the June 2008 and July 2011

algae blooms in the Yellow Sea, China (Ghosh, 2011; Marris, 2008) (Figure 2.9).

Methodology Development

46

Figure 2.9. Algae bloom observed off the coast of Qingdao, in the Yellow Sea, China on 28 June 2008

(Image source: Earth Observatory, NASA, 2008).

Environmental footprint

Environmental footprint is defined, under the framework of this methodology, as the

physical infrastructure a CCS technology has on the ground, for example, storage facilities,

other buildings or roads, etc. This sub-criterion is more easily comprehended by stakeholders

because of visibility, proximity and measurability. A high scoring example may be Ocean

Fertilisation since little, if any, physical infrastructure is needed. A low scoring example may be

Algae CCS since it requires large ponds next to emissions sources and significant infrastructure

for overnight storage of the CO2 (Global CCS Institute, 2011; Benneman, 2003).

Methodology Development

47

2.6 DISCUSSION

This quantitative methodology is designed as a first pass assessment of CCS technologies for

the purpose of providing a quick overview for a carbon management investment portfolio. Of

course, this can be followed up with more detailed assessment of individual evaluation criteria

or specific projects, and would take significantly more time, financial analysis and project-

specific information, e.g. exact costs from service providers. It is intended to provide a broad-

brush first selection mode for decision makers to: a) identify the CCS technologies which may

be applicable in a respective situation, and b) obtain an understanding of the main challenges

and risks which various CCS technologies may experience. The development and primary

calibration of the methodology is based on Geosequestration. This CCS option comprises the

most advanced set of carbon mitigation technologies, thus providing better availability of public

domain literature and case studies. While my preference is to use as much quantitative data as

possible, in large parts, such quantitative data is not easily accessible, and could not be

generated through statistical analysis due to the absence or very limited availability of data in

the public domain.

Applications and Future Improvements

Some applications of, and future improvements to, my methodology are as follows:

The in-depth analysis of the Master Matrix and Science and Technology Sub-Matrix

can highlight beneficial as well as less certain/higher risk areas that require further

improvement, R&D or assessment. This could include potential business opportunities,

e.g. sale of capture technologies by the O&G industry to other heavy emitter industries

such as coal-fired power generators or steel manufacturers; or the need to improve

certain practices, e.g. earlier and better engagement of local communities for improved

public acceptance rates.

The flexibility of the methodology allows for efficient reevaluation and updating of

criteria and sub-criteria to take into account the impact of recent events on a project

portfolio. For example, the Netherlands government legislation banning onshore storage

has forced Vattenfall to delay or cancel projects (Barendrecht, Eemscentrale and

Magnum) and rush to find economically and technically viable offshore storage sites

(ICIS Heren, 2011).

Methodology Development

48

The methodology can be tailored for a specific portfolio assessment or improved by

introducing weights for country or region specific issues. This would need further

detailed evaluation by the user, but the methodology is flexible enough to account for

this. With some adjustment this methodology can also compare projects based on the

same CCS technology. The application of the technique in this manner could provide

the most quantitative or least subjective comparative assessment.

The evaluation criteria currently have equal weighting. The methodology could be

improved and further extended by introducing variable weighting.

Flexibility in the methodology framework and design means that as a technology

progresses and more quantitative data becomes available, the methodology can be

updated and the sub-criteria modified if necessary.

Challenges

Some of the challenges I experienced in developing the methodology and rating system, and

issues to consider when applying it, are as follows:

The complexity of establishing suitable standardised measures or metrics and cut-off

points to compare across multiple CCS technologies encompassing multiple scientific

and technical disciplines.

The quantitative assessment of qualitative data (case studies, published literature and

reports) and transformation into quantitative scores, may still reflect user bias in the

decision making process, i.e. is still subjective to a certain degree.

One possibility to derive a more quantitative assessment would be a statistical analysis

of cancelled and delayed projects, however, such a database and/or analysis is not

readily available for Geosequestration as the calibrating CCS technology, and this

measure is not applicable for other technologies still at laboratory stage.

Additional calibration with other CCS technologies should be performed, especially if

in-house information is available.

To apply the methodology to other CCS technologies requires a time-consuming

literature review. To manage the large quantity and variety of data from multiple

disciplines, support documentation, such as fact sheets, analysis tables and databases, are

Methodology Development

49

also developed in order to make a full assessment. However, once the methodology has

been applied and analysis is complete (for that moment in time), the results can quickly

be updated as the hard work with the initial literature review has already been done.

Despite the challenges discussed above, the new availability of a standardised semi-

quantitative methodology to compare various CCS solutions should significantly assist with

focusing of research efforts and limited resources in order to help reduce atmospheric CO2

emissions.

2.7 CONCLUSIONS

I develop and present a new methodology to assess and compare multiple CCS technologies

and their barriers to adoption. I have defined a general evaluation criteria and ranking/scoring

system for CCS projects calibrated to the specific CCS technology of Geosequestration. This

semi-quantitative methodology can be applied to multiple alternative CCS technologies in both

a general and site-specific comparative manner, allowing the ranking of competing or

complementary technologies across a CCS project portfolio, and highlighting benefits or areas

that can yield improvement in the respective score. My new methodology should be both

practical and useful for CCS portfolio analysis of commercial-scale CO2 sequestration solutions.

In the next chapter, I apply the new methodology to evaluate Geosequestration as a carbon

mitigation option.

Methodology Development

50

Methodology Development

51

REFERENCES

Aarnes, J.E., Selmer-Olsen, S., Carpenter, M.E., and Flach, T.A. (2009), Towards guidelines for

selection, characterization and qualification of sites and projects for geological storage of CO2: Energy

Procedia, 1, 1, 1735–1742.

Ashworth, P., Rodriguez, S., and Miller, A. (2010), Case study of the CO2CRC Otway Project: CSIRO,

EP 103388, commissioned by Global CCS Institute, Canberra, ACT, Australia.

Bachu, S. (2000), Sequestration of CO2 in geological media: criteria and approach for site selection in

response to climate change: Energy Conversion and Management, 41, 9, 953–970.

Bachu, S. (2008), CO2 storage in geological media: Role, means, status and barriers to deployment,”

Progress in Energy and Combustion Science, 34, 254–273.

Benneman, J. (2003), Technology Roadmap–Biofixation of CO2 with Microalgae: Final Report

(7010000926) to the US Department of Energy, National Energy Technology Laboratory,

Morgantown-Pittsburgh, USA.

Boschetti, F., and Moresi, L. (2001), Interactive inversion in geosciences: Geophysics, 66, 4, 1226-1234.

Bradshaw, J., Bachu, S., Bonijoly, D., Burruss, R., Holloway, S., Christensen, N.P., and Mathiassen,

O.M., (2007), CO2 storage capacity estimation: Issues and development of standards: International

Journal of Greenhouse Gas Control, 1, 1, 62–68.

Brewer, P. G., Friederich, G., Peltzer, E. T., & Orr, F. M. (1999), Direct experiments on the ocean

disposal of fossil fuel CO2: Science, 284, 5686, 943-945.

Chevron Australia (2012), Injecting Billions of Dollars across Australia: Chevron Australia Pty. Ltd.

http://www.chevronaustralia.com/ourbusinesses/gorgon.aspx (accessed 20 June 2012).

CSA Group and IPAC-CO2 Research Incorporated [CSA IPAC-CO2] (2012), CSA Group and IPAC-

CO2 Research Inc. Announce World’s First Bi-National Standard for Geologic Storage of Carbon

Dioxide: [news release] 15 November 2012, CSA Group IPAC-CO2: Regina and Toronto, Canada.

http://www.ipac-co2.com/uploads/File/PDFs/CSA%20IPAC-

CO2%20CCS%20Standard%202012%20Final%20News%20Release%20v1%20copy.pdf (accessed

30 November 2012).

Department of Energy and Climate Change [DECC] (2010), Clean coal: an industrial strategy for the

development of carbon capture and storage across the UK: Her Majesty’s Govenment, Department of

Energy and Climate Change, URN 10D/570.

Methodology Development

52

Det Norske Veritas [DNV] (2012a), Recommended Practice DNV-RP-J203: Geological Storage of

Carbon Dioxide.

Det Norske Veritas [DNV] (2012b), DNV Service Specification DNV-DSS-402: Qualification

Management for Geological Storage of CO2.[replace dnv in text?]

Eiken, O., Zumberge, M., Torkjell, T., Sasagawa, G., and Nooner, S. (2004), Gravimetric monitoring of

gas production from the Troll Field: Presented at SEG 74th

Annual Meeting.

Eiken, O., Ringrose, P., Hermanrud, C., Nazarian, B., Torp, T.A., and Loier, L. (2011), Lessons learned

from 14 years of CCS operations: Sleipner, In Salah and Snohvit: Energy Procedia, 4, 5541-5548.

Farr, M. (2012), Carbon tax to blame for rise in electricity costs,” Herald Sun, 13 June 2012.

http://www.heraldsun.com.au/news/more-news/carbon-tax-to-blame-for-hike-in-electricity-

costs/story-e6frf7l6-1226394637354 (accessed 18 June 2012)

Fenn, J., Raskino, M. and Gammage, B. (2009), Gartner's Hype Cycle Special Report for 2009: ID No:

G00169747, Gartner Inc.

http://www.bnlcp.org.uk/sites/bnlcp/files/report/Gartners_hype_cycle_special_2009.pdf (accessed 21

June 2012).

Ghosh, P. (2011), China’s Eastern Coastline Teeming With Green Algae Bloom: International Business

Times: Science, 26 July 2011. http://www.ibtimes.com/articles/186963/20110726/china-algae-bloom-

qingdao.htm (accessed 17 May 2012).

Global CCS Institute (2011), Accelerating the uptake of CCS: Industrial use of captured carbon

dioxide: [Parsons Brinckerhoff], Global CCS Institute, Canberra, ACT, Australia.

Government of Western Australia (2011), Collie-South West CO2 Geosequestration Hub – Project and

Activity Progress Report for the Global Carbon Capture and Storage Institute: Government of

Western Australia, Department of Mines and Petroleum, Perth, Western Australia, Australia: Report

No. DMPNOV11_1578.

Grubb, M., Köhler, J., and Anderson, D. (2002), Induced technical change in energy and environmental

modeling: Analytic approaches and policy implications, Annual Review of Energy and the

Environment, 27, 1, 271-308.

Haszeldine, S. (2011), Why the UK buried a world-first carbon-capture scheme: New Scientist: Opinion,

21 October 2011. http://www.newscientist.com/article/dn21080-why-the-uk-buried-a-worldfirst-

carboncapture-scheme.html?full=true (accessed 7 June 2012).

Methodology Development

53

Helseth, J. (2011), CCS in Germany – is it game over?: Global CCS Institute,

http://www.globalccsinstitute.com/community/blogs/authors/jonashelseth/2011/10/10/ccs-germany-it-

game-over (accessed 7 June 2012).

Hoffman, A.J. (2005), Climate Change Strategy: The Business Logic Behind Voluntary Greenhouse Gas

Reductions: California Management Review, 47, 3, 21-46.

Hill, B., Hovorka, S., and Metzer, S. (2013), Geologic carbon storage through enhanced oil recovery: in

Proceedings of the 11th International Conference on Greenhouse Gas Control Technologies (GHGT-

11). http://www.fossiltransition.org/filebin/images/GHGT11_Hill_Hovorka_Melzer-2.pdf (accessed

1 April 2013) [in press].

Hovorka, S.D., Meckel, T.A., Trevino, R.H., Lu, J., Nicot, J.P., Choi, J.W., Freeman, D., Cook, P.,

Daley, T.M., Ajo-Franklin, J.B., Freifeild, B.M., Doughty, C., Carrigan, C.R. La Brecque, D.,

Kharaka, Y.K., Thordsen, J.J., Phelps, T.J. Yang, C. Romanak, K.D., Zhang, T., Holt, R.M., Lindler,

J.S., and Butsch, R. J. (2011), Monitoring a large volume CO2 injection: Year two results from

SECARB project at Denbury’s Cranfield, Mississippi, USA: Energy Procedia, 4, 3478-3485.

Iberdrola Group (2013), Lines of business.

http://www.iberdrola.com/webibd/corporativa/iberdrola?IDPAG=ENWEBCONLINEA&codCache=1

3662759340511511 (accessed 18 April 2013).

ICIS Heren (2011), Dutch CCS in disarray as ‘on land’ storage ruled out: ICIS Heren, 15 February 2011.

http://www.icis.com/heren/articles/2011/02/15/9435644/dutch-ccs-in-disarray-as-on-land-storage-

ruled-out.html (accessed 15 June 2012).

International Energy Agency Greenhouse Gas R&D Programme [IEA GHG] (2002), Ocean Storage

of CO2: 2nd Edition (revised), IEA Greenhouse Gas R&D Programme [Ormerod, W.G., Freund, P.

and Smith, A. (compilers), Davison J. (revisor)],

http://www.ieaghg.org/docs/general_publications/oceanrep.pdf (accessed 15 October 2011).

International Energy Agency Greenhouse Gas R&D Programme [IEA GHG] (2009), Natural and

Industrial Analogues for Geological Storage of Carbon Dioxide. IEA Greenhouse Gas R&D

Programme [Stenhouse, M. (compiler)], Cheltenham, Gloucestershire, UK.

Intergovernmental Panel on Climate Change [IPCC] (2005), Intergovernmental Panel on Climate

Change Special Report on Carbon Dioxide Capture and Storage: Prepared by Working Group III of

the Intergovernmental Panel on Climate Change [Metz, B.O., Davidson, H., de Coninck, C., Loos, M.

and Meyer, L.A. (eds.)]: Cambridge University Press, Cambridge, United Kingdom and New York,

NY, USA.

Methodology Development

54

Ipsos-MORI (2010), Experiment Earth: Report on a Public Dialogue on Geoengineering: Natural

Environment Research Council, Swindon, United Kingdom.

www.nerc.ac.uk/about/consult/geoengineering-dialogue-final-report.pdf (accessed 21 February 2013).

Jewell, W.J. and McCarty, P.L. (1971), Aerobic decomposition of algae: Environmental Science and

Technology, 5, 10, 1023-1031.

Kessler, R.A. (2009), Midwest group cancels Ohio CCS project despite DOE funding: Recharge, 21

August 2009. http://www.rechargenews.com/business_area/innovation/article186348.ece (accessed 30

May 2012).

Kuijper, M. (2011), Public acceptance challenges for onshore CO2 storage in Barendrecht: Energy

Procedia, 4, 6226-6233.

Lumley, D.E., Behrens, R.A., and Wang, Z. (1997), Assessing the technical risk of a 4-D seismic project:

The Leading Edge, 16, 9, 1287-1291.

Maher, S. (2012), Carbon Floor Price Warning: The Australian, 22 May 2012.

http://www.theaustralian.com.au/national-affairs/climate/carbon-floor-price-warning/story-e6frg6xf-

1226362855884 (accessed 20 June 2012).

Marris, E. (2008), Scientists identify algae that almost swamped the Olympics: Nature, 4 August 2008.

http://www.nature.com/news/2008/080804/full/news.2008.998.html (accessed 17 May 2012).

Miller, E., Summerville, J., Buys, L., and Bell, L. (2008), Initial public perceptions of carbon

sequestration: implications for engagement and environmental risk communication strategies:

International Journal of Global Environmental Issues, 8, (1), 147-164.

Neves, E.G., Petersen, J.B., Bartone, R.N. and Silva, C.A.D. (2003), Historical and socio-cultural origins

of Amazonian Dark Earths: In Amazonian Dark Earths: Origin, Properties, Management, [Lehmann,

J., Kern, D.C., Glaser, B, and Woods, W.I. (Eds.)]: Kluwer Academic Publishers, Dordrecht, The

Netherlands, 29-50.

North American Carbon Storage Atlas [NACSA] (2012), The North American Carbon Storage Atlas

2012: First Edition. http://www.netl.doe.gov/technologies/carbon_seq/refshelf/NACSA2012.pdf

(accessed 17 July 2012).

Nowak, D. J., Greenfield, E. J., Hoehn, R. E., and Lapoint, E. (2013), Carbon storage and sequestration

by trees in urban and community areas of the United States: Environmental Pollution, 178, 229-236.

Pacala, S., and Socolow, R. (2004), Stabilization Wedges: Solving the Climate Problem for the Next 50

Years with Current Technologies: Science, 305, 5686, 968-972.

Methodology Development

55

Porter, M.E. and van der Linde, C. (1995), Green and Competitive: Ending the Stalemate: Harvard

Business Review, Reprint 90557, 119-134.

Reiner, D.M. (2008), A looming rhetorical gap: A survey of public communications activities for carbon

dioxide and storage technologies: Electricity Policy Research Group, Working Papers, No. 801.

Reuters (2011), Update 2 – Vattenfall drops carbon capture project in Germany: Reuters, 5 December

2011. http://www.reuters.com/article/2011/12/05/vattenfall-carbon-idUSL5E7N53PG20111205

(accessed 7 June 2012).

Schlumberger Business Consulting Energy Institute [SBC Energy Institute] (2012), Leading the

Energy Transition: Bringing Carbon Capture and Storage to Market, [Soupa, O., Lajoie, B., Long,

A., and Alvarez, H.]: Schlumberger Business Consulting Energy Institute.

Schlumberger Business Consulting Energy Institute [SBC Energy Institute] (2013), Leading the

Energy Transition: Carbon Capture and Storage: Bringing Carbon Capture and Storage to Market:

Factbook Edition, January 2013 Update: Schlumberger Business Consulting Energy Institute.

Scott, K.C., Parker, D.J., Trupp, M., and Clulow, B. (2010), Setting New Environmental, Regulatory, and

Safety Benchmarks: The 2009 Gorgon CO2 3D Baseline Seismic Project, Barrow Island, Western

Australia: in Proceedings of the SPE Asia Pacific Oil and Gas Conference and Exhibition, edited,

Brisbane, Queensland, Australia.

Silver, W.L., Ostertag, R., and Lugo, A.E. (2000), The Potential for Sequestration Through Reforestation

of Abandoned Tropical and Pastural Lands: Restoration Ecology, 8, 4, 394-407.

Stalker, L., Varma, S., Van Gent, D., Haworth, J. and Sharma, S. (2013), South West Hub: a carbon

capture and storage project, Australian Journal of Earth Sciences: An International Geoscience

Journal of the Geological Society of Australia, 60, 45-58.

The Parliament of Western Australia (2011), Petroleum and Geothermal Energy Legislation

Amendment Bill 2011: Draft 3: Perth, Western Australia, Australia.

http://www.dmp.wa.gov.au/documents/Petroleum_and_Geothermal_Energy_Legislation_Amendment

_Bill_2011.pdf (accessed 1 April 2013).

Torp, T. and Brown, K. (2002), CO2 Underground Storage Costs as Experienced at Sleipner and

Weyburn: in Proceedings of the 7th International Conference on Greenhouse Gas Control

Technologies (GHGT-7).

Young-Lorenz, J.D. and Lumley, D. (2013), Portfolio analysis of carbon sequestration technologies and

barriers to adoption: General methodology and application to geological storage: Energy Procedia, 37,

5063-5079.

Methodology Development

56

Zero Emission Resource Organisation [ZERO] (2013), Collie South West CO2 hub.

http://www.zeroco2.no/projects/collie-south-west-hub (accessed 18 April 2013).

Geosequestration

57

CHAPTER 3

APPLICATION OF METHODOLOGY:

GEOSEQUESTRATION: CARBON CAPTURE AND

STORAGE IN GEOLOGICAL FORMATIONS

OVERVIEW

In this chapter, the methodology and rating system discussed in Chapter 2 are applied to

evaluate Geosequestration (GS); i.e. carbon capture and storage (CCS) in subsurface geological

formations. Although the rating of projects is site-specific, particularly when comparing onshore

and offshore, the storage capacity of geological formations to sequester large amounts of carbon

dioxide (CO2) within a relatively short time frame is recognised as a major advantage of this

carbon management strategy. The technical capacity of storage sites is defined by the capacity

which is available to reduce greenhouse gas (GHG) emissions with current technologies, and is

not equivalent to economic capacity. Economic capacity is the available storage capacity within

economic and social constraints. The quality of storage available worldwide is a major

advantage of GS in mitigating atmospheric CO2. Surprisingly, the expectation of a high rating

for the Science and Technology criterion of GS was not upheld, even though there are forty

years of relevant petroleum industry experience in injecting and producing fluids, drilling,

finding traps, and characterising and monitoring reservoirs. A lack of social licence to operate

for onshore CO2 storage, and the high costs of capture and long-term monitoring, lower the

rating. Additional issues relate to technology novelty in applying the technology for the storage

of CO2 rather than for oil and gas extraction, and technology transfer to other high emitter

industries. My results indicate that the most critical barriers to widespread commercial adoption

of GS are not technology- or capacity-related, but instead relate to issues of public acceptance

and economics. Despite these concerns, the analysis suggests that GS has a medium to high

probability of success as a commercial-scale CO2 storage option.

Geosequestration

58

Geosequestration

59

3.1 INTRODUCTION

As discussed in the Introduction chapter, global emissions of greenhouse gases (GHG)

covered by the Kyoto Protocol have increased approximately 70 percent (%) from 1970–2004,

and by 24% from 1990–2004. Carbon dioxide (CO2) is the largest contributor to this increase,

having grown by about 80% (IPCC, 2007). In order to limit temperature increases from pre-

industrial levels to 2 degrees Celsius (°C) several global climate models suggest that

atmospheric concentrations of GHG must be stabilised at 450 parts per million (ppm) of CO2

equivalent by 2050 (Global CCS Institute, 2011a). In the Stabilisation Triangle Model,

presented by Pacala and Socolow in 2004, GS represents one of the most promising strategies to

mitigate CO2.

According to the International Energy Agency’s [IEA] Energy Technology Perspectives

2012, GS may be capable of accounting for about 17% of the recommended CO2 emissions

reductions required by 2050. GS involves the capture of CO2 at industrial emission source

points, transport of the CO2 to a storage site, and injection of the CO2 into a subsurface

geological storage reservoir. Industrial CO2 source points include coal-fired power generation,

steel, chemical, fertiliser, cement, mineral, and liquified natural gas (LNG) plants. Transport of

the captured CO2 occurs via pipeline or alternative methods such as shipping, trucking etc.

Suitable storage sites are typically geological formations deeper than ~1km where a

combination of porous reservoir rock (e.g. sandstone) and non-permeable cap-rock (e.g. shale)

form a reservoir-seal pair. The 1km upper depth limit results from the pressures and

temperatures needed to keep CO2 in a compressed liquid form. The most favourable storage

formations are depleted oil and gas reservoirs and saline aquifers deep in the subsurface. Oil and

gas reservoirs are subsurface rock formations that act as natural traps of fluids and gases (fossil

fuels). They already have a proven reservoir-seal pair as hydrocarbons were held in place prior

to production for millions of years. Once the majority of the hydrocarbons have been produced

and the reservoirs end their economic life, the reservoirs are said to be depleted. Saline aquifers,

sometimes referred to as saline sinks, are porous and permeable rock formations containing very

salty brine water that are located deep in the subsurface of sedimentary basins both onshore, and

offshore on the continental shelves. The water resources are not potable (drinkable) and

therefore have limited or no economic value. The CO2 in liquid supercritical phase is injected

and sequestered by a variety of geological trapping mechanisms over time. Figure 3.2 shows a

schematic of the GS process.

Geosequestration

60

Figure 3.1. Schematic Diagram of the Geosequestration Process. Methane gas produced from offshore

gas reservoirs is piped onshore and converted to LNG for transport or burned to generate electricity).

The CO2 is separated from the gas stream and piped onshore or offshore for injection and storage in

depleted oil and gas reservoirs, and deep saline aquifers (Image source: Scottish Centre for Carbon

Storage, 2011).

There are currently only a few commercial-scale GS projects active around the world.

Commercial-scale GS projects are defined as storing 1 metric tonne of Carbon per annum

(Mtpa). A metric tonne (1000 kilogram (kg) = 2240 pounds (lbs) in comparison to an

imperial/US/short ton (2000 lbs = 907kg). A tonne is approximately 10% more than a ton.

Sleipner, a gas field in the Norwegian North Sea, was the first commercial CO2 storage project.

Since 1996, 16Mt CO2 has been injected at a rate of 1Mtpa, at about 1,000m below sea level

into the Utsira Formation, a saline sandstone formation (Eiken et al., 2011). The Weyburn-

Midale project (Saskatchewan, Canada), operational since 2000, utilises CO2 piped from the

Great Plains Synfuels Plant in North Dakota, USA for enhanced oil recovery, injecting CO2 into

the Weyburn (2.4Mtpa CO2) and into Midale (0.4Mtpa CO2) carbonate reservoir oil reservoirs

(Whittaker et al., 2011; Whittaker, 2004; Wilson and Monea, 2004). At the In Salah CCS

project in Algeria, 1.2Mtpa CO2 from gas processing has been injected 1800m into the Krechba

Formation, a depleted oil reservoir in Algeria since 2004 (Eiken et al., 2011).

The Cranfield Project in Mississippi, USA, began in 2008 as a pilot test project and by

March 2012 had injected a total of 5 million metric tonnes CO2 (3.5 million metric tonnes from

the Jackson Dome natural source and 1.5 million metric tonnes produced and recycled from

Plant Barry Power station) into the lower Tuscaloosa Formation, a deep sandstone saline aquifer

Geosequestration

61

(Hill et al., 2013). At the Snøhvit project in the Barents Sea, 0.7Mtpa CO2 from LNG

processing has been injected 2,600m depth into the saline Tubasan sandstone formation

reservoirs since the field became operational in 2008 (Eiken et al., 2011). The Decatur Project

in Illinois (USA) utilises CO2 from an industrial source, injecting 1Mt CO2 over a three year

period beginning in November 2011 from ethanol processing into the Mount Simon Sandstone

Formation, a saline sink (National Energy Technology Laboratory, United States Department of

Energy [NETL DOE], 2011; United States Department of Energy [US DOE], 2012).

New CCS projects which will soon be operational are the Quest project in Alberta, Canada,

and the Gorgon and South West Hub Projects, both located in Western Australia. Quest is part

of the Athabasca Oil Sands Project, where 1.2Mtpa CO2, representing 35% of emissions from

the Shell Scotford Upgrader, a bitumen processing facility, will be captured and injected

onshore at a depth of 2,000m - 2,500m into the Cambrian Basal Sands (Voser, 2012). Starting in

2014, at Gorgon, 3 - 4Mtpa from an LNG gas processing plant will be injected 2,200m below

Barrow Island into the Dupuy Formation, a turbidite sandstone reservoir. This will make

Gorgon the largest GS project worldwide (Scott et al., 2010). The South West Hub Project,

south of Perth, Australia, is expected to inject between 3 - 7Mtpa of CO2 from various

industrial sources, including coal-fired power plants, into the Lesueur Formation, an onshore

deep sandstone saline aquifer (Government of Western Australia, 2011). An overview of the

location of active and proposed projects as of January 2012 is given in Figure 3 (CO2CRC,

2010).

Figure 3.2. Overview on active and proposed Geosequestration projects (Image source: CO2CRC, 2012)

Geosequestration

62

A wide range of numerous small CCS pilot and demonstration projects of < 100 kilo tonnes

per annum (ktpa) have been implemented globally over the past 10+ years (Jenkins et al., 2011;

Schilling et al., 2009; Aimard et al., 2007). These include the Otway Project in Victoria

(Australia) where over 65,000 tonnes (t) CO2 has been injected at a depth of 2000m into the

Waarre-C Formation, a depleted sandstone gas reservoir (Jenkins et al., 2011). In Germany,

Ketzin became a major European CO2 storage test site in 2008, injecting a total of 60kt CO2

since inception into a shallow saline formation at about 650m depth (Schilling et al., 2009).

Since 2010, the TOTAL Lacq Project has injected 75kt of CO2 from oxyfuel combustion at

4,500m depth into an onshore depleted natural gas field (Aimard et al., 2007). A key objective

of these many test projects is to perform research experiments and develop the necessary

expertise for commercial upscaling of GS.

In the following sections, the new methodology developed in this thesis will be applied

specifically to GS to evaluate its current status and capability as part of a carbon mitigation

strategy, and to highlight its benefits and challenges for full-scale commercial adoption. Case

studies and examples will be provided to support the assessed scores. I conclude the chapter

with a discussion of the applicability of my methodology to CCS technologies in general and

the outcome of its application to GS.

3.2 METHODOLOGY

The technology elements unique to GS as a CCS option are identified and input into the Sub-

Matrix (refer to Table 3.4):

1) Site Identification and Characterisation,

2) Capture,

3) Transportation,

4) Injection and Storage, and

5) Measurement, Monitoring and Verification (MMV).

I have identified these technology elements from a major literature review of GS including

governmental, international agency and corporate reports, journal databases, websites, etc.

(IPCC, 2005; IEA GHG, 2012; Schlumberger Business Consulting Energy Institute [SBC

Energy Institute], 2012). A description of each element, including core scientific and

Geosequestration

63

engineering concepts, is discussed in this section. An assessment across the evaluation criteria

generates individual ratings and these provide the overall technical readiness score in the

Science and Technology column of the Master Matrix (Table 3.5).

3.2.1 Technology Element 1: Site Identification and Characterisation

“Site Identification and Characterisation” encompasses the assessment of the economic and

technical feasibility of geological storage sites capable of sequestering CO2, for secure, long

term storage at the volumes necessary to offset anthropogenic CO2 emissions. Many of the

scientific and technical issues discussed under this element, such as storage capacity controls

and the properties of injected CO2, also affect Technology Element 4: Injection and Storage.

CO2 can be injected into sedimentary basins that contain both porous permeable rocks (e.g.

sandstone) as a storage formation and non-permeable rocks (e.g. claystone, shale) as a seal, or

cap rock. Porosity is the measure of the void space in the rock for fluids or gases, and

permeability is the measure of the ability of the rock to allow fluid flow. These reservoir-seal

pairs provide suitable storage sites for CO2 (Figure 3.4) and include depleted oil and gas (O&G)

reservoirs and saline formations. Depleted O&G reservoirs are considered the best immediate

candidates for storage of CO2 as they have high porosity and permeability, traps and seals, and

their retention potential is proven with fossil fuels trapped for millions of years. In addition,

many of these reservoirs are already well characterised as part of hydrocarbon exploration and

production, including computer earth models developed to predict and match the movement of

the hydrocarbon displacement and history matched with the production data. Existing

infrastructure and wells can also be utilised for GS. While the number of depleted O&G fields

is limited and global capacity of known discoveries estimated between 675 - 900 GtCO2, the

global storage potential of deep saline aquifers in sedimentary basins is estimated between 1000

to 11,000GtCO2. Saline formations represent the largest storage potential but are currently not

well characterised, with many uncertainties about permeability, porosity, seals or traps. (IPCC,

2005)

Ideally, storage sites should be located close to the source point of emissions, which could be

sources such as oil and gas production characterised by high CO2 content, or an industrially

emitted source, such as coal-fired power generators or natural gas processing plants (Figure

3.1). The geological, geophysical and geochemical aspects of each basin are assessed to

estimate volume capacity and ability to securely store CO2 for the long term. The trapping

mechanisms of geological sequestration (e.g. structural and stratigraphic traps, capillary

Geosequestration

64

pressure / residual saturation, solubility and mineralisation) (Bachu, 2003) are discussed under

Section 3.2.4.2 (Technology Element 4: Injection and Storage).

Figure 3.3. Injection of carbon dioxide into depleted oil and gas reservoir (Image source: Scottish Centre

for Carbon Storage, 2012.)

Properties of Carbon Dioxide

One prerequisite of site characterisation is the identification of the depths at which CO2

changes phase and becomes a supercritical fluid. This is dependent on temperature and pressure,

with the critical point of pure CO2 being 31.1 degrees Celsius (⁰C) and 7.3 megapascals (MPa)

(Angus et al., 1973) (Figure 3.4). A 400-fold decrease in volume occurs when CO2 changes to

supercritical or dense-phase, typically at >800m depth, with the exact threshold depending on

the geothermal temperature and pressure gradient of the Basin (Figure 3.5). CO2 is transported

and injected in a supercritical state as it exhibits gas-like viscosity and liquid-like density, with

volume efficiencies and better fluid-flow diffusion properties.

When CO2 is first injected, it does not dissolve immediately in the formation water, and also

may not mix homogeneously. Instead, it is said to be free-phase or largely immiscible. In

supercritical phase, it is less dense than the formation water and due to gravity, migrates

upwards driven by buoyancy (the density difference between the CO2 and formation water) until

it becomes trapped on the seal rock surface. Buoyancy pressure (Pb) is opposed by capillary

pressure (Pc), which is the difference in pressure across the surface of two immiscible fluids.

Geosequestration

65

Sufficient buoyancy pressure is needed to exceed the capillary pressure at each pore throat so

injected CO2 can migrate through a reservoir pore system. Sealing rock, such as shale, has high

capillary pressure and low permeability which thereby stops the CO2 from migrating upwards.

Another important consideration is the need for high-permeability fluid flow migration

pathways from the injector throughout the reservoir rock to ensure that the injected CO2 fills the

available pore space in the reservoir.

Figure 3.4. Illustration of the phase changes of CO2 in response to temperature – pressure conditions.

The critical point represents the (P,T) conditions beyond which the CO2 exhibits liquid-like density at a

temp of 31.1 °C and a pressure of 73 bar = 7.3MPa, typically corresponding to >800m injection depth in

the earth (Image source: ChemicaLogic Corporation, 1999).

The volumetric equation for storage capacity of CO2 in a saline formation is based on the

standard industry method to calculate original oil in place (subsurface volumes, in situ fluid

distributions, and fluid displacement processes) (Calhoun Jr., 1982; Lake, 1989; NETL DOE,

2010):

= Volumetric storage capacity

= Total area (Basin/Region/Site) being assessed

Geosequestration

66

= Gross thickness of target saline formation defined by At

= Total porosity over thickness hg in area At

= Density of CO2 at pressure and temperature of saline formation

= Storage “efficiency factor” (fraction of total pore volume in saline aquifer filled

by CO2)

Fig 3.5. CO2 Depth versus Density

Chart, illustrating the change in CO2

volume and density as it changes from

a gas to a supercritical fluid as

temperature and pressure increase

with injection depth in the subsurface

(Image source: CO2CRC, 2012).

Rock Properties

The rock property criteria for site characterisation are: 1) Injectivity (can the CO2 be injected

into the rock?), 2) Containment (can the CO2 be retained in the rock?), and 3) Capacity (what

volume of CO2 can the rock hold?). All are controlled by porosity, permeability, and irreducible

fluid saturation, which themselves depend on rock lithology and fluid properties at reservoir

pressure and temperature. Porosity (φ) is the percent/fraction measure of the storage (void)

space in the rock for fluids or gases, divided by the total rock volume. Total porosity includes

all the voids, and effective porosity includes only the accessible voids, or connected spaces.

Permeability (k), which is measured in units of millidarcy (mD), is a measure of the ability of

fluid to flow through the rock, and is therefore related to injectivity. Permeability is strongly

affected by the geometry of the pore space, in particular the size of the small pore spaces or

gaps connecting the larger pore volumes in the rock (pore throats). Table 3.1 describes these

and other rock properties which affect the ability to sequester carbon dioxide.

Geosequestration

67

Table 3.1. Rock properties affecting the ability to sequester CO2 (Sheriff, 2002; Selley, 1998; Angus et al.,

1973; CO2CRC, 2008).

Fluid saturation refers to the relative volume (percentage/fraction) of a fluid or gas

compared to the available pore space or porosity. Subsurface rocks naturally contain some

amount of water, brine, oil and gas that must be compressed or displaced in order to accept the

injected CO2. When CO2 is injected, the water is displaced out of the rock. It will not displace

100% of the original fluid, and the amount which remains is the irreducible fluid saturation of

that phase. For example, a thin film of water remains around each grain because of surface

tension forces, thereby “wetting” the rock grains. The irreducible fluid saturation depends on

specific rock types. It is a critical control on storage capacity and relates to relative permeability

(the permeability of one fluid compared to another, usually water). A low irreducible fluid

saturation reflects a high capacity of storage availability, for example, an irreducible water

saturation (SWIRR) of 7% means 93% of connected pore space is available for storage.

Sedimentary Basin Assessment

Assessment of a sedimentary basin for CO2 storage requires evaluation of a range of factors

(Table 3.1) including regional and structural geology studies (outcrops, rock/well core samples),

Geosequestration

68

fluid tests, stratigraphic and seismic interpretation, application of other geophysical techniques,

such as gravity and electrical resistivity surveys, static and dynamic modelling (flow

simulation), and geomechanical and injection studies (CO2CRC, 2008; Bergmann et al., 2012;

Kopp et al., 2009a, 2009b). Amongst other factors, these studies aim to evaluate the reservoir

capacity, quality, and effectiveness to retain the injected CO2.

Table 3.2. Factors included in the assessment of a sedimentary basin (adapted from CO2CRC, 2008).

3.2.2 Technology Element 2: Capture

“Capture” is the removal of CO2 produced by burning of fossil fuels from large stationary

emitters before it is vented into the atmosphere. Examples of stationary emitters include gas and

coal-fired power stations, and industrial processors, such as LNG operations, steel refineries,

cement factories, and chemical plants. The focus of this chapter is on the power generators as

they are the largest industrial emitter of CO2 (IEA, 2008b; IEA GHG, 2007). The flue gas

produced from the combustion of fossil fuels to generate electricity contains mostly nitrogen

(N2) and low levels of CO2 ranging from 3% to 15% (IPCC, 2005), which must be separated

from the other gas before it can be compressed and transported for storage or reuse. This

section contains a brief overview of CO2 capture and separation technologies.

Geosequestration

69

Capture Processes

The three basic types of CO2 capture process are Post-Combustion, Pre-Combustion, and

Oxyfuel. Figure 3.6 illustrates these processes as well as natural gas processing.

Post-Combustion

Post-Combustion, or “end-of-pipe,” capture technologies can be installed on existing power

plants through retrofitting, and for future power plants (IPCC, 2005; Rackley, 2010a).

Following combustion of the fossil fuel, the flue gas passes through several processes of

separation which “clean” it through the use of liquid sorbents, such as amines, e.g.

monoethanolamine (MEA), absorbing the CO2 through chemical binding and then removing it

from the flue gas. Following the capture, the sorbent is regenerated through a desorber. This

allows the reuse of the sorbent and releases the captured CO2 for sequestration, or utilisation for

industrial purposes (IPCC, 2005; MacDowell et al., 2010).

Pre-Combustion

Pre-Combustion capture technologies, also known as Integrated Gasification Combined

Cycle (IGCC), use chemical processes to gasify the fossil fuel. A stream of almost pure oxygen

(O2) reacts with the fossil fuel through high pressure and temperature to form synthetic gas, or

“syngas,” which contains carbon monoxide (CO), CO2 and hydrogen (H2). By further

processing of the gas stream with water, the CO is converted to CO2 and H2. The H2 is extracted

and combusted through steam turbines to generate electricity, and the CO2 is captured using a

physical wash. (IPCC, 2005; MacDowell et al., 2010)

Oxyfuel

Oxyfuel combustion, also known as oxyfiring, uses pure O2 instead of air to combust the

fossil fuels, producing only CO2 and water as emissions. An air separation unit removes the N2

from the air (+/-78% volume), and produces O2. The fossil fuels are then combusted with the O2

in a boiler where steam is produced to power turbines and generate electricity. The flue gas

from this process contains high CO2 concentrations (>80% volume) which can easily and

efficiently be separated. (IPCC, 2005; MacDowell et al., 2010)

Geosequestration

70

Figure 3.6. Carbon Dioxide Capture Processes for LNG Processing, Post-Combustion, Pre-Combustion

and Oxyfuel (Image source: CO2CRC, 2012).

Capture and Separation Technologies

Absorption

Solvent-based absorption is a process in which carbon dioxide is absorbed from a gas stream

by using a liquid solvent, such as an amine like monoethanolamine (MEA), and waste heat

(Figure 3.8). Other solvents include carbonates and bicarbonates, but all are used in substantial

quantities. Recovery, or regeneration, of the solvent for reuse also requires a large amount of

power to remove impurities, for example, fly ash, soot, nitrogen oxides (NOX) and sulphur

oxides (SOX), which can reduce the absorption capacity of the solvent. The CO2 is removed

from the liquid which is recycled, and the gas stream emitted to the atmosphere. Absorption is

used in post-combustion capture (IPCC, 2005; MacDowell et al., 2010).

Geosequestration

71

Adsorption

Adsorption is a process where CO2 molecules are adsorbed from a gas stream and attracted

to accumulate on the surface of a solid or liquid (adsorbent) through physical or chemical

sorption. Examples of adsorbents include: silica gels and aluminas, metal organic frameworks,

other organic-inorganic hybrids, mineral zeolites, and meso-porous carbons. Amines can be

added to the pores of the adsorbent to attract CO2 molecules thus improving the adsorption

capacity of the material. Adsorption is used in Pre- and Post-Combustion capture that both

involve hot and wet flue gases. The three steps taken in adsorption systems are adsorption of

CO2, removal of impure gases, and desorption (IPCC, 2005; MacDowell et al., 2010).

Desorption

Desorption refers to the extraction, or recovery and release of CO2 from the liquid solvent

adsorbent with the use of membranes, temperature and pressure changes. This process is

achieved through a number of methods including: Thermal Swing Adsorption, Vacuum Swing

Adsorption, Pressure Swing Adsorption, and Electrical Swing Adsorption (Redhead, 1962;

Rackley, 2010b).

Membranes

Membranes can be used to filter out CO2 from gas streams. Made from polymers or

ceramics, the materials can be designed to work as physical gas separation membranes, or to

incorporate liquid solvent absorption stages and capture the CO2 during diffusion between the

pores of the membrane. For gas separation, the membrane acts as a semi-permeable barrier

through which the CO2 passes more easily than other gases. Enough pressure difference is

needed to drive separation across the membrane. The most common mechanisms for gas

separation in membranes are molecular sieving and solution-diffusion. Mechanisms for

membrane gas separation include spiral wound, flat sheet and hollow fibre (IPCC, 2005;

Rackley, 2010c). Systems combining the use of membranes with solvents can be used in the

absorption and desorption (Feron and Jansen, 2002).

Geosequestration

72

Figure 3.7. Absorption – Desorption is a cyclical process that captures CO2 by solvent-based absorption,

recovers the CO2, and re-utilises the solvent and waste heat (Image source: CO2CRC, 2012).

Cryogenic Technique

The use of low temperatures to cool, condense and purify CO2 from moderately concentrated

gas streams is called the Cryogenic technique and used in Oxyfuel combustion. The flue gas is

cooled to sub-zero temperatures under high pressure so only the CO2 condenses and other gases

(N2 and O2) are separated and released (Burt et al., 2009; MacDowell et al., 2010). Chilled

water is passed through cooling flue gas to form hydrate crystals with CO2 trapped inside. The

CO2 is released from the hydrates by heating (Surovtseva et al., 2011).

3.2.3 Technology Element 3: Transportation

The third technical element of GS is the safe and secure “Transportation” of the CO2 from

emissions source points to injection sites. CO2 can be transported as a solid, liquid or gas,

although gaseous CO2 is large volumetrically and not practical to handle. Methods of liquid CO2

transportation include ships, rail and road tankers, however, pipelines are the most common

Geosequestration

73

form of transportation used to transfer supercritical CO2. As CO2 is corrosive, pipelines are

generally constructed from carbon steel (Seiersten, 2001). Impurities (SO2, NO2, and H2S) are

removed and the CO2 is dried (<50ppm) before compression and liquefication to mitigate

corrosion and the risk of gas hydrate formation (IPCC, 2005).

The USA has extensive experience and a good safety record with the transportation of

supercritical CO2 by pipeline for enhanced oil recovery purposes, with ~ 6,200km of pipeline in

operation (Rackley, 2010e; IEA GHG, 2012). A high pressure between 13 - 19Mpa is

maintained for transport. The operating temperature for CO2 pipelines in the USA varies

between 4⁰C to 24⁰C (Hooper, 2011), with the CO2 alternating between supercritical and dense

phase. The alternative method of large volume transport is ships which carry liquid CO2.

Recommended transportation conditions are around 0.7MPa and -50⁰C. Offshore discharge of

CO2 requires heating and compression equipment on either the ship or offshore platforms in

order to inject the CO2 for storage (IPCC, 2005).

3.2.4 Technology Element 4: Injection and Storage

“Injection and Storage” include some of the points previously discussed under Technology

Element 1: Site Identification and Characterisation; specifically the properties of injected CO2,

and storage capacity controls (porosity, permeability and irreducible fluid saturation).

Injection

CO2 is injected at higher pressure than the initial reservoir pressure, resulting in reservoir

pressure build up. This pressure has to be kept below fracture pressure at which point the seal

rock could break and leakage could occur. The injection rate (volume/time) depends upon: the

area of the wellbore in contact with the formation, injection pressure, and permeability (Streit et

al., 2005; CO2CRC, 2008). Clean unconsolidated sandstone would be an example of a high

permeability reservoir rock with high injectivity. An example of a low permeability reservoir

rock with low injectivity would be a heavily cemented and mineralised sandstone.

Geosequestration

74

Storage

Secure long-term storage of injected CO2 in geological formations is dependent upon a

mixture of the four main types of trapping mechanisms: 1) Structural and Stratigraphic, 2)

Residual Saturation, 3) Solubility, and 4) Mineralisation. The way in which the CO2 is injected

and flows through the geological formation (plume migration), together with the time it remains

in storage, determines the relative proportion of the trapping mechanisms over time (Figure

3.9). CO2 plume migration in the reservoir depends on a number of factors including the

heterogeneity of the reservoir rock as well as the type of displaced fluid and presence of

permeability flow paths and barriers (CO2CRC, 2008; IPCC, 2005).

Figure 3.8. Illustration of physical and geochemical trapping mechanisms showing percentage of CO2

trapped and mechanism applicability versus time. The storage security of later trapping mechanisms is

higher than earlier mechanisms (Image source: Elements: Geoscience World, 2013).

Structural and Stratigraphic Trapping

Structural and stratigraphic trapping are physical mechanisms and the most familiar and best

understood form of trapping by geoscientists and thus present the lowest risk. Structural

trapping describes the process of CO2 being trapped in a local structural high like an anticline or

Geosequestration

75

tilted fault block which is covered by a non-porous, non-permeable seal rock (claystone or

shale) and thus inhibits further ascent of the fluid (Salvi et al., 2000; Selley, 1998). Stratigraphic

traps refer to geological containers where the lateral seals are provided by a change in rock type

(facies change), thereby preventing further lateral migration, and encompass features like pinch-

outs, unconformities, and channels (Selley, 1998). Structural and stratigraphic trapping occurs

within an immediate timeframe and continues as the plume migrates throughout the formation

(Figures 3.8 and 3.9) (IPCC, 2005).

Figure 3.9. Structural and Stratigraphic Carbon Dioxide Physical Trapping Mechanisms. Top left:

Structural - Fault trapping, Top right: Structural - Anticline or Dome/Fold trapping, Lower left:

Stratigraphic – Pinch-out or Facies Change trapping, Lower right: Stratigraphic - Unconformity

trapping (Image source: CO2CRC, 2011).

Residual Saturation Trapping

Residual CO2 saturation trapping during plume migration occurs when the CO2 is trapped in

pore space by surface tension or capillary pressure. It is a hydrodynamic process also referred to

Geosequestration

76

as capillary trapping. When CO2 is injected, it is an immiscible fluid such that the free-phase

CO2 plume displaces the in situ fluids in the reservoir. If the buoyancy of the CO2 or the fluid

flow and the associated pressure difference are not sufficient to overcome the capillary pressure

of the pore throats, the CO2 will be immobilised, thereby trapping the CO2 in pore space before

the seal is reached. This residually trapped CO2 can eventually dissolve into the formation

water. Residual gas saturation is a function of pore geometry, such that a high pore volume to

throat size ratio favours residual CO2 trapping (Obdam et al., 2003; Kumar et al., 2005).

Residual saturation / capillary trapping is significant in the short term (Holtz, 2003) (Figure

3.10).

Figure 3.10. Residual Saturation Trapping (Image source: CO2CRC, 2011).

Solubility

Solubility of the CO2 in the formation water is a hydrodynamic process and a migration

associated trapping mechanism. CO2 diffused in reservoir water (aqueous CO2) forms a weak

acid; carbonic acid (H2CO3):

Increasing pressure increases the solubility of CO2 in water, but it decreases with increasing

water salinity and temperature. As CO2 dissolves in the formation water, the water becomes

denser, and begins to sink downwards allowing the CO2 to become more dispersed. The extent

of mixing (convective flow) depends on the thickness of the storage formation (Figure 3.12).

The amount of CO2 dispersed in the water (storage effectiveness) can increase over time (Figure

3.9) (Ennis-King and Paterson, 2003).

Geosequestration

77

Figure 3.11.

Solubility Trapping Mechanism

(Image source: CO2CRC,

2011).

Mineral Trapping

Mineral trapping (mineralisation) is a geochemical process closely linked with solubility

trapping. It refers to the Water – Rock – CO2 chemical interactions which occur when CO2 is

dissolved in water. The carbonic acid reacts with the reservoir rock to form and precipitate solid

carbonate minerals thus trapping CO2 in its most stable form as a mineral. Examples of solid

carbonate minerals are calcium carbonate, or calcite (CaCO3), and magnesite. The potential for

geochemical reactions to occur and form these precipitates/minerals depends on: composition of

the reservoir rock, rate of fluid flow, temperature and pressure, chemical composition of the

water, and water to rock contact area. (Gunter et al., 1993, 1997: Perkins et al., 2005)

An example of calcite mineralisation is given in the following geochemical reaction:

CO2 storage effectiveness increases with time for mineral trapping, and the process could

continue for over a thousand years after injection. Mineralisation is considered the ideal

trapping mechanism and most secure form of CO2 sequestration (IPCC, 2005).

Geosequestration

78

3.2.5 Technology Element 5: Measurement, Monitoring and Verification (MMV)

CO2 project operators, governments and other stakeholders are very concerned that the CO2

does not leak from the geological formations, potentially causing environmental damage or

posing hazardous situations and health risks (IEA, 2011; Jenkins et al., 2011; IPCC, 2005).

MMV provides a mechanism for assurance monitoring by taking measurements of geological,

geophysical and geochemical data as well as of air and water quality to ensure the technical

integrity of the storage site, and provide assurance to both regulators and the public. Operators

also need to account for the stored volume of carbon to gain economic benefits such as carbon

credits or to offset carbon debt (Hill et al., 2013).

It is important to take baseline measurements (including seasonal fluctuations) prior to

injection as there could be pre-existing, naturally occurring levels of CO2 within a reservoir or

in the surrounding environment. MMV provides an early warning system of CO2 leaks or

unexpected reservoir behaviour. MMV is used to check the accuracy of dynamic reservoir

models and predicted CO2 plume migration paths through further calibration and updating as

measurements are made over time. Operators and governments need to measure and monitor the

volumes of CO2 injected and how much remains securely contained in the geological formation

(Jenkins et al., 2011; Gorecki et al., 2012; Hill et al., 2013; IPCC, 2005).

Geosequestration

79

3.3 RESULTS

In this section, I apply the general concepts of my methodology for a semi-quantitative

assessment to the specific CCS technology of GS at its current state of development for a

general overview and from a global perspective. The more in-depth assessment for Science and

Technology is presented last for ease of reading/reference (Section 3.3.1). These results

represent an analysis of data from available case studies and regional bias could be attributed to

project location and public information. This analysis follows the one presented in Young-

Lorenz and Lumley (2013).

3.3.1 Public Acceptance

Public, stakeholder and political acceptance is increasingly seen by governments and

industry as crucial for the adoption of carbon mitigation projects to proceed. It has proven

difficult for GS project proponents to establish credibility and trust without extensive public

engagement (Kuijper, 2011; Itaoka et al. 2012). A lack of social licence to operate can result in

vocal public opposition and rejection by stakeholders, contributing to a negative perception.

Lack of public acceptance is compounded by involvement from non-local political

organisations, e.g. environmental activists, and media interactions which amplify possible risk

events and can even spread scientific misinformation (Itaoko et al., 2012).

How does GS affect me?

The CCS benefits of GS are generally accepted by the wider public (Shackley et al., 2007;

Ashworth et al., 2009; Itaoka et al., 2004; Reiner et al., 2006), and storage in remote and

offshore locations is generally deemed to be publicly acceptable, as for example at Sleipner

(Reiner, 2008). Concerns about onshore storage due to feared loss in property values, and health

and safety issues, often result in a not in my backyard (NIMBY) response that can lead to

cancellation of projects such as Barendrecht (Kuijper, 2011) and the Midwest Regional Carbon

Sequestration Project (Itaoka et al., 2012). Other local community fears include induced

seismicity, undesirable property access, and potential leakage into potable water supplies

(Kessler, 2009; Kuijper, 2011; SBC Energy Institute, 2013; Reuters, 2011; Global CCS

Institute, 2011b).

Geosequestration

80

The Barendrecht demonstration project in the Netherlands was cancelled at a late stage

because of fears of CO2 leakage, despite the long-term presence and production of methane

previously stored in the two depleted reservoirs. Failure to effectively engage the local

community resulted in demonstrations by active local groups which utilised internet activity to

generate opposition to the project (Kuijper, 2011). The Midwest Regional Carbon

Sequestration Project in Greenville, Ohio was also cancelled at a late stage of development. This

project would have injected 1Mtpa CO2 from a nearby ethanol plant more than 3000 feet

beneath the surface into the Mount Simon Sandstone formation. After reaching the execute

stage, the project was cancelled in August 2009 because of fears of increased seismic activity

and lowered property values, and in spite of receiving US$61 million funding from the

Department of Energy (Kessler, 2009). The balancing of public acceptance of offshore storage

but not onshore leads to an overall score of 2 for the how does GS affect me sub-criterion.

How natural is GS?

GS is viewed as an acceleration of a natural process in the carbon cycle in which CO2 is

stored naturally in geological formations for tens of thousands to millions of years (IEA GHG,

2009). Natural accumulations of CO2 exist in locations all over the world (IPCC, 2005). During

oil and gas extraction and processing, the CO2 is extracted at these locations and following

processing, is either vented or can be re-injected back into the geological site where it had been

safely stored for geological time scales. Thus, the how natural is GS sub-criterion scores highly

at 3.

Science and technology comprehension

The public generally has a limited knowledge about the scientific concepts underlying GS

and this can negatively impact the level of acceptance (Itaoka et al., 2012; Global CCS Institute,

2011b, Whitmarsh et al., 2011; European Commission, 2011). A common misconception about

geological storage is that the injection process into the subsurface is similar to blowing up a

huge balloon underground, i.e. leading to pressure build-up and potential leakage (Wallquist et

al., 2010) or induced seismicity (Global CCS Institute, 2011b). Unreasonable fears of

asphyxiation due to pressure build up and sudden CO2 release are partly responsible for the

cancellation of projects such as Barendrecht (Wallquist et al., 2010; Kuijper, 2011). For these

reasons, science and technology comprehension scores a 1.

Geosequestration

81

The combination of the sub-criteria scores results in an overall “public neutral” (2) Public

Acceptance rating. This is the score that is input into the Public Acceptance column of the

Master Matrix as illustrated in Table 3.3. Similarly, the evaluation criteria ratings will be

entered into the relevant columns of the Master Matrix.

Table 3.3 Illustration of average score of sub-criteria input into Public Acceptance column of the Master

Matrix.

3.3.2 Regulatory Frameworks

Enabling Body of Laws and Regulations

Much progress has been made in establishing regulatory frameworks for the implementation

of GS with the United Kingdom (UK), European Union (EU), Alberta (Canada), Norway, State

Government of Western Australia, and the Commonwealth of Australia leading.

The enabling body of laws and regulations fall mainly into three categories:

a) Emission limits

Emission limits are imposed on a national or regional level in the form of specific

emission limits or the introduction of a cost for CO2 emission. An example for specific

limits would be the recently introduced amendment to the UK Energy Bill, proposing a

statutory emission rate of 200g/kWh. This would require new gas-fired power generators

to install CCS technology (Bellona, 2013a). Other limits on emitters are introduced

through the setting of a carbon price, like the Australia’s Clean Energy Legislative

Package taxation scheme, which establishes a carbon price of AUD $23 until it switches

to a cap-and-trade market pricing in July 2015, linking with the EU’s Emission’s Trading

Scheme (ETS) and/or other carbon markets (The Commonwealth of Australia, 2012).

b) Funding incentives

Funding incentives are provided to companies to compensate for the higher CAPEX and

OPEX due to the introduction/integration of CCS technologies and encourage adoption.

Geosequestration

82

Examples are the EU’s New Entrance Reserve (NER 300) programme and the UK’s re-

launch of the £1 billion fund, previously allocated to Longannet, for a second CCS

Commercialisation Programme competition to develop a viable CCS industry in the UK.

c) Storage and Infrastructure

Legislation and regulation for storage and infrastructure introduces new or modifies

current laws to allow for CO2 storage and transportation, e.g. the amendment of existing

pipeline legislation in Western Australia to include the transportation of CO2 by pipeline

(The Parliament of Western Australia, 2011). In 2006, the amendment of The London

Convention and Protocol on the Prevention of Marine Pollution by Dumping of Wastes

and Other Matter, incorporated CO2 storage in geological formations under the sea bed

into international treaties. In 2007, The Convention for the Protection of the Marine

Environment of the North-East Atlantic (“OSPAR Convention”) was also amended to

allow for geological storage under the seabed; however this has yet to be ratified.

While significant progress has been made to establish regulatory frameworks for GS,

inconsistencies in regulation at both national and international levels, e.g. USA and Australia

(IEA, 2011), still negatively impact its uptake and application. The London Convention permits

storage, but does not allow for trans-border transportation impacting offshore storage in

particular (IEA, 2011). A draft Communication from the European Commission to EU member

states calls for greater commitment from both governments and industry to regulate sectoral

emission performance standards and/or mandatory CCS certificate schemes (Bellona, 2013b).

For the reasons discussed above, the enabling regulation sub-criterion scores a 2 rating.

Negative legislation?

Negative legislation in the form of restricted access to economically viable storage space, or

restrictions on the use of transport infrastructure, limits the uptake of GS. The Barendrecht,

Magnum and Eeemscentrale projects in the Netherlands are negatively impacted by the lack of

access to onshore storage sites (ICIS Heren, 2011). In the UK, access to the transport

infrastructure for CO2 is severely limited as no supercritical CO2 can be piped onshore (WA

ERA, 2010). While some of these negative legislations are historic and can be solved through

amendments to existing legislation as in Western Australia (The Parliament of Western

Australia, 2011), others are recent enactments as is the case in the Netherlands. As a result,

negative legislation scores a 1.

Geosequestration

83

Regulatory Certainty

Areas of priority for regulators are the long-term liability of GS projects and cross-border

jurisdictions, especially for onshore vs. offshore, marine dumping and transportation (IEA,

2011). High policy uncertainty for CCS can result in a significant increase of risk and planning

uncertainty (costs, liability and project delays), which may motivate commercial enterprises to

cancel projects and thus further hinder the adoption of a CCS technology (Bachu, 2008). For

example, the delay in enactment of proposed CCS legislation by the German government,

apparently led Swedish power utility Vattenfall to cancel the Jänschwalde project in

Brandenburg (Reuters, 2011; Helseth, 2011). This 1.5 billion Euro demonstration project would

have stored 1.7Mtpa CO2 in an onshore saline aquifer.

In 2008, the Commonwealth of Australia became the first country to legislate federal

exploration leases for CO2 offshore storage sites. This established the framework for temporary

ownership rights and provides operators with more certainty to plan for long-term projects.

Similar regulation was introduced in the UK, where the Crown owns the rights, and would issue

land leases for offshore CO2 storage up to the 200 mile (320 km) continental shelf international

boundary. The CO2 storage licences would be obtained through the Department of Energy and

Climate Change (DECC), with a draft licensing structure modelled on the current gas licensing

structure (Parliamentary Office of Science and Technology [POST], 2009).

Regulatory compliance to reduce CO2 emissions and/or mandate utilisation of CCS

equipment can significantly speed up the adoption process by making it a cost of doing

business. The mandated retrofitting of coal-fired power plants with CCS equipment by 2025 in

the UK (IEA, 2011) or the EU’s Large Combustion Plant Directive (European Union [EU],

2001) are two examples that provide a level competitive playing field for power plant operators.

It is estimated approximately one-third of the UK’s power requirements will be generated from

gas and coal-fired plants employing CCS by 2050 (National Audit Office, 2012). The costs of

retrofitting or inclusion of CCS equipment can be transferred to the consumer through price

increases without loss of competitiveness, if mandates to reduce CO2 emissions are legislated

and price increases approved.

The lack of certainty about long-term liability after handover is a significant barrier as it

poses significant economic risk for the operator. In the EU, any damage to the environment is

covered by the 2009 EU Environmental Liabilities Directive. The site operator is liable for

damage up to 30 years after an incident takes place, irrespective of the time the facility closes.

While not ideal due to the continuing liability beyond project end, it still provides a clearly

defined time frame for liability. The IEA’s CCS Review (2011) stated that the trend tends to be

Geosequestration

84

towards transferral of long-term liability from the operator to government, but that there are

notable differences between jurisdictions as to interpretation and processes of the transferral

requirements.

Operators in the UK and EU are responsible, under Article 18 of the EU Directive on the

Geological Storage of CO2, and must cover post-closure costs, liabilities and all other

obligations until project end when the reservoir is sealed and the facilities decommissioned.

Transfer of liability to the government only occurs once the operator provides a report to

demonstrate the following:

(a) the conformity of the actual behaviour of the injected CO2 with the modelled

behaviour;

(b) the absence of any detectable leakage;

(c) that the storage site is evolving towards a situation of long-term stability.

(European Parliament and the Council of the European Union [EU],2009)

MMV rules must also be in place for operators and governments to qualify or audit for

carbon accounting (payment of money or credits), or make it clear when legal obligations are

fulfilled, e.g. EU CCS Directive and the EU ETS Monitoring and Reporting Guidelines. An

independently managed fund may be required for CO2 storage liabilities as per the nuclear

experience (UKERC, 2012), however, to date, there is no such independent insurance fund or

special trust fund which covers the post-closure liabilities of leakage remediation (IEA, 2011) or

liability associated with emission repayments due to leakage (SBC Energy Institute, 2012).

Regulatory certainty scores a 2 rating.

Averaging the three sub-criteria thus brings the overall Regulatory Frameworks score to a

“neutral” (2) rating, which is entered into the Regulatory Frameworks column of the Master

Matrix.

Geosequestration

85

3.3.3 Economics

Viable Business Case

Oil and gas (O&G) operators use Carbon Capture and Utilisation and Storage (CCUS) to

generate supplementary income from enhanced oil recovery (EOR) which is not readily

available to other industrial CO2 emitters. Without this economic benefit, they would be

unlikely to commit the significant investment CCS represents on a large scale (with the

exception of possible R&D or test case purposes) unless regulation made it mandatory.

Combining GS with reutilisation of CO2 for hydrocarbon recovery offers the opportunity to

achieve both economic viability and CO2 mitigation. R&D efforts are underway to gain insight

about CO2 projects in EOR contexts (Gorecki et al., 2012). The Sectoral Assessment CO2

Enhanced Oil Recovery Report (2011), by Advanced Resources International for the United

Nations Industrial Development Organisation (UNIDO), estimates the long-term storage

volumes of CO2 through EOR at 318GtCO2, with the capacity of the largest 54 oil basins in the

world about 140GtCO2, and the potential recovery of 470 billion additional barrels of oil. Figure

3.13 illustrates CO2 sources, pipelines and CO2 projects in the USA.

The petroleum industry also has the advantage of technical expertise and experience with

onshore EOR in the USA since 1972, whereas the coal and other heavy CO2 emitter industries

do not, so a significant gap of GS knowledge and experience exists between the industries. The

slow adoption of GS also means that skill sets are developed but lost as experts, engineers and

technicians move to more financially stable industries. Events like the Global Finance Crisis

have negatively impacted uptake of GS technologies and projects, for example, both the

Freeport Gasification Project in Texas (Global CCS Institute, 2011c), and Phase 2 of the

American Electric Power [AEP] Mountaineer Project in West Virginia (Wald and Broder, 2011;

AEP, 2011) were cancelled because of the uncertainty of US climate policy and weak economic

conditions.

GS is costly but not prohibitive (Cook, 2012) with the most significant expenditure at the

capture stage. Other factors impacting the economics of a project include the existence of

reusable legacy wells, distance between emissions source point and storage site, and number of

emissions sources (SBC Energy Institute, 2012). The cost of CO2 avoided is lower for coal-fired

CCS power plants than for some alternatives such as renewables like solar and offshore wind

(Global CCS Institute, 2011d; SBC Energy Institute, 2012; ZEP, 2011). The levelised cost of

electricity (LCOE) of a CCS coal-fired plant is US$89-139 per megawatt hour (/Mwh) and for a

CCS natural gas plant US$107-119/MWh (Global CCS Institute, 2011d). A dollar value is put

Geosequestration

86

on emissions through the “cost of CO2 avoided,” expressed in dollars per tonne of CO2. The cost

of CO2 emissions avoided for a coal-fired power plant with CCS technology ranges between

US$23–92/t compared with US$139-201/t for solar thermal and US$90-176/t for offshore wind

(Global CCS Institute, 2011d). The 2011 European Technology Platform for Zero Emission

Fossil Fuel Power Plants [ZEP] Report on the Costs of CO2 Capture, Transport and Storage:

Post-demonstration CCS in the EU stated that CCS will be cost-competitive with other low-

carbon energy technologies after 2020.

Figure 3.12. Overview of onshore CO2 projects in the USA, together with the utilised CO2 sources and the

associated pipeline network. The CO2 is primarily used for EOR (Image source: Melzer, 2011).

In the presence of CO2 emissions penalties, the CAPEX and OPEX costs can be offset

through financial savings in penalty rates. For instance, the Norwegian Government has both

emissions legislation and a carbon price; therefore state-owned operator Statoil had economic

frameworks to assess decisions for projects like Sleipner where 1Mtpa CO2 is injected into the

reservoir until 2025, saving Statoil around US$100,000 in carbon taxes every day (Torp and

Brown, 2001). In 2012, the DECC (UK), announced plans under its proposed new energy bill,

to introduce an agreed Contract for Difference (CfD) for CCS generated power which would be

competitive with strike prices against other low carbon generation technologies and against the

traded electricity market price. The CfD would provide an instrument of long term support for

the high CAPEX and means electricity companies can make investment decisions to build CCS

equipped fossil fuel power stations without Government capital. With this, CCS and low-carbon

fossil-fuelled electricity generation has the potential to be a profitable industry. (The House of

Commons, 2013)

The discussion above indicates that the viable business case sub-criterion scores a 2 rating.

Geosequestration

87

Dependency Public Funding

There is currently a high level of dependency on government financial incentives for

projects, for example Longannet (Haszeldine, 2011; BBC News Online [BBC] 2011a, 2011b)

and the South West (SW) CO2 Sequestration Hub project (Government of Western Australia,

2011). GS projects without the supplementary income from enhanced hydrocarbon recovery are

currently only financially viable with government support, making them very susceptible to

policy changes. The Longannet Project, located in Scotland and winner of the CCS Competition

(£1billion) was the only planned commercial project in the UK, but was cancelled when the UK

Government withdrew its support after failure to agree terms with the consortium partners.

Longannet is the third largest coal-fired plant in Europe, producing 2400MW and one of the

largest polluters (7-8Mtpa CO2). A 2Mt, 260km pipeline was planned to transport CO2 from

Longannet, Fife to the North Sea for injection into depleted oil and gas reservoirs. The SW Hub

project in Western Australia was allocated AUD52 million from the Commonwealth of

Australia’s AUD1.68 billion Carbon Capture and Storage Flagships (CCS Flagships)

programme, part of the Australian Government’s $5 billion Clean Energy Initiative in June

2011. An additional AUD330 million is available should it progress past the current stage

(Government of Western Australia, 2011).

Government funding and incentives dependency earns a rating of 1.

Carbon Market?

At the Climate Summit held in Durban at the end of 2011, the Kyoto Protocol was extended

to 2017/2020 and CCS was included as a Clean Development Mechanism (CDM) that enables

the generation of Certified Emissions Reductions (CERs) carbon credits for sale on carbon

markets. As this is a relatively recent inclusion, GS is not sensitive to a weak carbon price and

immature market at this time. Inclusion of GS as a CDM will encourage developing nations to

adopt CCS technologies, but the largest GHG emitters, China, India and the USA, do not

qualify to generate CER carbon credits for trade or offset. Cap and trade systems are still under

development and cannot be relied upon for long-term reliable revenue streams (Aldy and

Stavins, 2011). The price of carbon has significantly decreased as a result of the lack of a global

regulatory framework and an oversupply of carbon credits. In 2011, the weighted average price

of carbon credits globally was only US$15/t CO2 (SBC Energy Institute, 2012). Therefore the

carbon market sub-criterion earns a neutral (2) rating.

Geosequestration

88

The sub-criteria averages out to a score of 1.7 which may represent an “affordable” economic

assessment and is entered into the Economics column of the Master Matrix.

Figure 3.13. Carbon prices over time in Europe for the period April 2010 to April 2012, showing a steady

decline since mid 2011 (Image source: van Renssen, 2012).

3.3.4 Storage Quality

Capacity

Global estimates of storage sites available for Geosquestration range from 1,700 –

11,000GtCO2 (IPCC, 2005). In 2007, Bradshaw et al. proposed guidelines to more accurately

estimate storage capacity and viability of geological storage. Worldwide research efforts are

underway to assess storage capacity in suitable geological formations (NETL DOE, 2010; North

American Carbon Storage Atlas [NACSA], 2012; The Norwegian Petroleum Directorate, 2012;

Vangkilde-Pedersen et al., 2009). In Australia, the available CO2 storage volume is estimated at

100 Mtpa for a 260 – 1120 year period on the West Coast, and 200 Mtpa for a 70 – 450 year

periods on the East Coast (Carbon Storage Taskforce, 2009). Storage volumes for Mexico,

Canada and the USA is estimated to be 1.9 metric trillion tonnes, equating to 500+ to 5000

years’ worth of storage capacity (NACSA, 2012). For this region, the low end (500 years)

scenario places the capacity at 136 billion metric tonnes (bmt) in depleted O&G fields, 1738bmt

in saline aquifers and 65 bmt in coal fields.

Geosequestration

89

Depleted O&G fields are considered the best immediate candidates for storage of CO2 as

they have high porosity and permeability, and their retention potential is proven with fossil fuels

having been trapped for millions of years. In addition, many of these fields are already well

characterised through their production history. While the number of depleted O&G fields is

limited, the global storage potential of deep saline aquifers in sedimentary basins is estimated

between 100 - 11,000 Gt, but they are currently not well characterised (IPCC, 2005).

GS is very site and situation specific, with geological and lithological aspects of a site

affecting storage capacity. Operators need certain volumes of capacity to match project needs.

For example, the Gorgon project on the North West Shelf of Australia will soon be the largest

CO2 sequestration project in the world. Operator Chevron, requires a certain volume to be

injected over the life of the project, a 15mtpa liquefied natural gas (LNG) plant on Barrow

Island and a domestic gas plant for the Western Australia market (300 terajoules per day). An

estimated 3.4 - 4 mtpa of CO2 will be injected into a deep saline aquifer 2.5km below Barrow

Island into the Dupuy Formation for approximately forty years (Scott et al., 2010; Chevron,

2012) (Figure 3.15). An estimated 250 mtCO2 storage space will be needed to meet the project’s

volume requirements. Economic and social constraints result in a differentiation of technical

and economic storage capacity because of land and resource competition and the matching of

emissions sources to logistically suitable and publically acceptable sinks; however the large

volumes of technical capacity around the world justify a volume rating of 3

.

Figure 3.14. The Gorgon Project Development Concept illustrating the Greater Gorgon gas fields, gas

plant, and pipeline network to the mainland and Barrow Island.

Geosequestration

90

Rate

Large volumes of CO2 can be injected into reservoir formations quickly at pressures below

fracturing (fracking) point. While the rate of injection is related to the number of injection wells

(Hosa et al., 2011), the injectivity rate is site specific and largely dependent on the permeability

of the reservoir rock, thickness of the formation and the pressure gradient. The technically

feasible injection rate of a storage site must be aligned with the needs of the project. In the case

of the Snøhvit gas field, the CO2 injectivity of the rock was lower than initially predicted,

resulting in pressure build up and the risk of exceeding the fracking point by September 2011

(Molde, 2011; Helgesen, 2010). The resulting reduction in injection rate had technical and

economic implications for operator Statoil which has no permission to vent. From a

comparative perspective, the rate of GS compares more favourably to other CCS technologies

such as Afforestation or Algae which depend on photosynthesis and individual growth rates,

therefore GS rate earns a 3.

Duration

Storage of CO2 in geological formations is considered long-term and secure with duration

and storage integrity increasing over time, and dependent on the four CO2 trapping mechanisms

discussed in Section 3.2.4 (IPCC, 2005). Geological uncertainty caused by a lack of site

characterisation constrains the ability to predict and model interactions between the injected

CO2 with the reservoir rock, or the trapping mechanisms of capillary, dissolution and mineral

trapping. Limited knowledge exists about the effects of mineral and rock diagenesis due to the

presence of CO2 and the associated effects on porosity and permeability beyond the extraction

time frame (WA ERA, 2010). This can be offset through experiments on core samples (e.g. core

flood, formation fluid chemistry for rates of mineral dissolution) to build up a knowledge base

for different types of reservoir rock (Ketzer et al., 2009). Extensive knowledge about the

behaviour of fluids with respect to structural and stratigraphic trapping mechanisms is available

throughout the O&G industry due to their analogies with the hydrocarbon extraction process.

For these reasons, duration earns a 3.

Verification Potential

Current commercially available MMV technologies are capable of monitoring the CO2 plume

migration, identifying the presence of CO2 leakage, and provide input to static and dynamic

Geosequestration

91

reservoir models in offshore and onshore environments. These include chemical measurements

to assess soil, air, fluid, well cores, etc. as well as geophysical techniques with Figure 3.15

providing a brief overview (Lumley, 2001, 2011; Rackley, 2010d; Hill et al., 2013, Hovorka et

al., 2011; Eiken et al., 2004, 2011; Gilfillan et al., 2009). Examples are discussed in Section

3.3.6 (Technology Element 5: Measurement Monitoring and Verification) Increasing experience

will result in improved interpretation of the measurements to address geological uncertainties.

In addition, most countries have an existing legislative framework to allow for land access for

geophysical surveying work. This legal framework can be adapted to the needs of MMV

activities which typically require longer term arrangements to accommodate for land ownership

changes, re-zoning, as well as a higher frequency of access. For safety, risk and uncertainty

management procedures, the experience of the O&G industry with hazardous gases can be

utilised. Verification also merits a 3 rating.

Storage Quality scores a “high” (3) rating which is input into the Storage Quality column of the

Master Matrix. This evaluation criterion is a major advantage of GS as a carbon mitigation

option.

Figure 3.15. Geophysical techniques to characterise sedimentary basins for suitability as CO2 storage

sites and to monitor changes in the reservoirs (Image source: Subsurface Imaging Group, The Ohio State

University, 2012).

Geosequestration

92

3.3.5 Environmental Impact

Potential impact on ecosystems and water systems

Globally, CCS prevents the discharge of GHG and contributes to the mitigation of climate

change, thereby positively impacting the environment (IPCC, 2005). Injection occurs at depths

where the CO2 is unlikely to interfere with surface resources, unlike other CCS technologies

such as Ocean Fertilisation. The localised impact of any negative event due to the defined

storage location limits the environmental impact. Public concerns about health, safety and

environmental impacts are framed by inappropriate analogy with rapid emission of large

volumes of naturally occurring CO2 in volcanic areas which resulted in fatalities in Indonesia

and West Africa, or damage to the local ecosystem in California, USA (IEA GHG, 2009).

There are no recorded incidents of sudden CO2 emissions from sedimentary basins, and

although slow natural seepage has been detected along faults or from boreholes, very few

examples of localised environmental damage have been linked to this phenomena (IEA GHG,

2008, 2009, Link 1). Hazardous CO2 release requires the build up of trapped CO2 in gas phase

(IEA GHG, Link 1), which can be identified through ongoing MMV activities.

While amine solvents, used in the capture process, could potentially produce carcinogens,

current technologies and management practices can mitigate the impact on environment and

health (Shao and Stangeland, 2009). Recent research suggests that the risk of formation and

spreading of nitrosamines is less than previously thought (Skriung, 2011). The potential risks

associated with injection of CO2 into the subsurface are that acute or chronic leakage of the CO2

could cause damage or contaminate resources in other subsurface zones, affecting access to

resources such as hydrocarbons, groundwater, mineral deposits, geothermal, minerals, coal, etc.

(Bachu, 2008; Holloway, 2001). These risks can be minimised through rigorous site selection

and characterisation. Potential impact on ecosystems and water systems scores a 3.

Environmental Footprint

There is some environmental footprint from characterisation, infrastructure, transportation

(pipelines) and MMV activities. The footprint is not larger than for conventional O&G

activities, which is commonly considered tolerable if not in highly environmentally sensitive

areas. Existing regulations and impact management plans can be modified to manage the above

factors. The concern that the injection of large volumes of CO2 in areas close to faults can result

in sufficient pressure build-up to trigger earthquakes (induced seismicity) and may lead to the

Geosequestration

93

fracturing of the reservoir seals and release of the stored CO2 (Zobeck and Gorelick, 2012)

seems highly improbable, and in any event such risks can be minimised by rigorous site

selection in tectonically stable regions, active monitoring and engineering mitigation solutions

(Bachu, 2003; McClure, 2012; Det Norske Veritas, 2012a; National Research Council, 2012),

therefore environmental footprint earns a 2 rating.

Environmental Impact thus has an overall “high” (2.5) score which is input in the

Environmental Impact column of the Master Matrix.

3.3.6 Science and Technology: Sub-Matrix Assessment

I identified the Science and Technology elements specific to GS, and evaluated each against

the evaluation criteria. The Technology elements are:

1) Site identification and characterisation

2) Capture

3) Transportation

4) Injection and Storage

5) Measurement, Monitoring and Verification (MMV)

Technology Element 1: Site Identification and Characterisation

In general, the wider public is not typically aware of the activities undertaken to gather the

information required to identify and characterise potential CO2 storage sites. Interaction with

property owners is restricted to terrestrial surveys and test drilling, but only affects a very

restricted area. Offshore or regional acquisition with air-borne surveying (onshore) is

straightforward and requires no to minimal interaction with local property owners, therefore

Public Acceptance scores a 3. There is no consistent legislation regulating these activities,

although existing legislation already covers most survey activities requiring licensing, such as

land access. The DNV CCS certification framework released in 2012 and the CSA IPAC-CO2

bi-national Canada-USA consensus standard (CSA Z741) announced the same year provide the

basis for internationally recognised processes for site characterisation based on industry best

practice and compliance with regulations, international standards and directives (DNV, 2012a,

2012b; CSA IPAC-CO2, 2012; Aarnes et al., 2009). Regulatory Frameworks scores a 2.

Geosequestration

94

Site characterisation is the most time-consuming and costly part of the site selection process,

but not unaffordable, with assessment of suitable geological storage sites being undertaken

worldwide (NASCA, 2012; The Norwegian Petroleum Directorate, 2012; Carbon Storage

Taskforce, 2009). Figure 3.17 shows the result of the technical analysis for Australia. Therefore

Economics earns a rating of 2. The technology needed to identify and characterise geological

formations is often the same as that used by the energy and minerals industries, therefore

Demonstrated Technical Readiness rates a 3. There is some footprint from characterisation

activities, particularly terrestrial geophysical surveys and test drilling so Environmental Impact

scores a 2.

The combined average rating for the Site Identification and Characterisation is very positive

(2.4).

Figure 3.16 CO2 storage potential and emissions sources in Australia (Image source: CO2CRC,

2013).

Technology Element 2: Capture

Capture involves the separation and removal of CO2 from flue gas produced by the burning

of fossil fuels at power plants and other industries prior to venting into the atmosphere. The

IEA’s Energy Technology Analysis Report (2008b) states that the long-term potential of CCS in

mitigating CO2 levels comes from power generation (50%), industrial hydrogen (30%), heavy

Geosequestration

95

industries (16%), and natural gas processing (4%). This element gets a 3 rating for Public

Acceptance because there is no perceived risk, and it is seen to benefit the environment. Few

countries have legislated industry compliance to utilise capture technologies or cap emissions,

therefore Regulatory Frameworks earns a 1.

Capture represents the most costly part of CCS because flue gas contains low concentrations

of CO2, making separation a complicated and expensive process. Current R&D efforts are

focused on improving efficiencies of capture technologies and lowering its costs (ZEP, 2011;

Burt et al., 2009; Baxter et al., 2009; CO2CRC, 2012; IEA 2008a). Examples include the

reduction of parasitic energy use, new/improved solvents, configuration of separation devices,

and incorporation of membrane liquid adsorption. For these reasons, Economics earns a 1.

Demonstrated Technical Readiness scores a 3, with post-combustion the most mature capture

method (IPCC, 2005), capturing ~90% of CO2 from industrial operations, although a system-

dependent energy penalty of 16% to 30% applies (SBC Energy Institute, 2012; IEA GHG,

2007). Capture prevents CO2 release, and current technologies and management practices can

mitigate the impact of amine solvents on environment and health (Shao and Stangeland, 2009,

Skriung, 2011) therefore Environmental Impact scores a 3.

Capture scores a “medium” (2.2) rating.

Technology Element 3: Transportation

Pipelines are the main solution for high volume transportation of CO2, although scattered

emission source points, remote offshore storage sites or environmentally hostile transit areas

may favour the use of maritime transport. Only pipelines are analysed as all active or planned

projects are based on this mode of transportation. The environmental footprint associated with

pipeline infrastructure could negatively impact Public Acceptance if located through densely

populated or environmentally sensitive areas, therefore Public Acceptance of Transportation

scores a 2. CO2 transportation can be covered by existing or modified pipeline legislation, as is

the case in Western Australia (The Parliament of Western Australia, 2011). Figure 3.18

illustrates the extent of the overall pipeline network in Western Australia. Adaptation of

regulatory health and safety codes for the design and operation of pipeline routing may be

required (Gale and Davison, 2004; IPCC, 2005; Carbon Storage Taskforce, 2009). In 2010, Det

Norske Veritas [DNV] published a Recommended Practice guideline for the design and

operation of CO2 pipelines. The availability of existing regulatory frameworks, recommended

Geosequestration

96

practice guidelines, and political flexibility to adapt legislation yields a Regulatory Framework

rating of 3.

Pipelines are the most effective and economic onshore solution. The building costs of

pipeline infrastructure are impacted more by transport distance and environment than transport

capacity, e.g. the CAPEX costs of onshore CO2 pipeline infrastructure are $0.2-0.4 million/km

for a 50cm diameter, 10MtCO2pa capacity, but 40 – 70% higher for offshore (Gale and Davison,

2004; SBC Energy Institute, 2012). North America is currently the only region with a dedicated

pipeline network for CO2 transport (IEA GHG, 2012; Watt, 2010). In Australia, more than

5,000 km of large diameter pipeline is needed to transport CO2 (Carbon Storage Taskforce,

2009). In Europe, an integrated and inter-country pipeline infrastructure (30,000 – 150,000km)

needs to be built (IEA GHG, 2005). In order to develop integrated onshore and offshore pipeline

infrastructure, passage and land acquisition rights, monitoring for pressure, leakage, and

pipeline integrity, metering, gating from different emitters, cost-effective materials, custody

transfer across county, state and country borders and associated costs should be included in any

planning (Congressional Research Service [CRS], 2009; Watt, 2010; Seevam et al., 2008).

Both economic and technical feasibility of transportation systems are dependent on the

number of source points and chemical composition of the CO2. The upgrading or reuse of

existing pipeline networks requires fit-for-purpose technical analysis as supercritical CO2 has

higher acidity and requires higher pressure than natural gas (SBC Energy Institute, 2012), but

could significantly reduce CAPEX. The OPEX of pipeline transportation for CO2 is lower than

that of natural gas or hydrogen because of its density in supercritical state (IEA, 2008a).

Pipeline networks create economies of scale to lower risk and share costs, e.g. multi-user “hub-

and-spoke” approaches are planned for both the South West Hub project, Western Australia and

CarbonNET project in Victoria (WA ERA, 2010). The Economics of Transportation earns a

score of 3.

The use of pipeline technology for the transportation of more volatile gases than CO2 is

already mature (IPCC, 2005) and the challenges of transporting supercritical CO2 are well

understood. The USA has an excellent safety record with four decades of experience in CO2-

dedicated pipeline networks for onshore EOR (>40Mtpa over 7,300km of pipeline), with the

CO2 coming mainly from naturally occurring and therefore, relatively pure CO2 (SBC Global

Institute, 2012; CRS, 2007, 2009; Watt, 2010). Offshore pipelines for Sleipner in the North Sea

and Snøhvit in the Barents Sea, the latter being a very hostile environment, provide technical

reference cases for offshore transportation (Watt, 2010). For these reasons Demonstrated

Geosequestration

97

Technical Readiness scores a 3. Environmental Impact is rated a 2 because of the

environmental footprint from pipeline networks.

The overall score for Transportation technology is (2.6) which counts as a high rating.

Figure 3.17.

Gas Pipeline Infrastructure

map of Western Australia as

at 30 June 2010

(Image source: Economic

Regulation Authority, 2010).

Technology Element 4: Injection and Storage

Injection and Storage refers to the injection of supercritical CO2 (greater than 31.1⁰C and

7.3MPa) into geological formations suitable for long-term storage and secure containment.

While the rate of injection is related to the number of injection wells (Hosa et al., 2011), the

injectivity rate is site specific and largely dependent on the permeability of the reservoir rock,

thickness of the formation and the pressure gradient. The USEPA’s Underground Control

Geosequestration

98

Program (USA) demonstrated that fracture pressures range from 11 - 21 kilo Pascals (kPa)

(IPCC, 2005). Too high a pressure can cause fracturing, thereby affecting storage estimates and

impacting the technical and economic parameters of a dependent project, as Statoil discovered

while injecting at Snøvit in the Barents Sea (Molde, 2011; Helgesen, 2010).

Frequent community concerns about economic impact, health, safety and environmental

damage from CO2 injection activities and potential leakage at onshore storage sites give a score

of 1 for Public Acceptance. Nation-specific legislation both supports (Australia) (The

Commonwealth of Australia, 2012) and restricts (Netherlands) (ICIS Heren, 2011) the ability

for storage of CO2 in geological formations. This balances out in a Regulatory Frameworks

score of 2. The costs for injection and storage infrastructure are site-specific, but these can be

approximated from the drilling of similar wells in the O&G industry (IEA, 2008a). Additional

expenses are monitoring equipment and well bore isolation (cementing) to mitigate potential

interaction of the CO2 with the reservoir. Offshore wells are more expensive than onshore (IEA,

2008a)(Figure 3.19). The Economics of Injection and Storage earns a score of 2. Demonstrated

Technical Readiness has been proven with 40+ years of EOR experience (over 100 projects) by

the O&G industry (IPCC, 2005; Beckworth, 2011) and it therefore earns a score of 3.

Environmental Impact rates a 2 because of environmental footprint from injection and

monitoring wells, and the potential of induced microseismic events.

The overall rating for Injectivity and Storage is “medium” (2).

Geosequestration

99

Figure 3.18. Bankability cost for onshore and offshore storage projects in depleted gas fields in

Europe. Onshore fields show a lower Probability of Success with a limited difference in mean cost. 90p

means that 90% of the projects are below the cost threshold, 10p that 10% of the projects are below this

threshold (Image source: Global CCS Institute, 2011).

Geosequestration

100

Technology Element 5: Measurement Monitoring and Verification (MMV)

Measurement, Monitoring and Verification (MMV) refers to the use of tools and tests to

monitor CO2 injected into geological formations by use of geophysical and chemical data

(seismic, gravity, EM, inSAR, rock, fluid, mineralisation, air, soil, etc.) to build an accurate

accounting of the stored volume and its security or retention capability, for safety and economic

purposes (IPCC, 2005; WA ERA, 2010). Table 3.4 illustrates the variety of technologies and

methods available to monitor the CO2 distribution in the subsurface and at the surface for CO2

leak detection.

Table 3.4. Measurement, Monitoring and Verification (MMV) Technologies and Methods available to

monitor the CO2 distribution in the subsurface and at the surface for CO2 leak detection.

Geosequestration

101

All surveys and tests, whether subsurface or surface based, require the acquisition of baseline

surveys, preferably with seasonal variations, to define the changes over time. These data sets

will be integrated in a similar way as for production monitoring in the O&G industry to create

and update geological and geomechanical models of the subsurface processes and define the

progress of the plume migration (Vanorio et al., 2010). Legacy wells require special attention

because they can compromise the integrity of the storage reservoir (WA ERA, 2010).

Stakeholders want assurance about the safe and secure containment of the CO2 thus Public

Acceptance of MMV is rated 3. Some regulatory framework exists, but there is no consistent

global legislation for the handover of liability, legal responsibilities and MMV obligations (IEA,

2011) and Regulatory Frameworks balances out as a score of 2. The costs of MMV activities are

affordable, however these occur over the long term (>50 years) with the frequency of

measurement significantly higher than for a typical O&G monitoring project which typically

would continue for twenty years. The frequency may not be determined by technical necessity,

but by public and political concerns. For example, expensive time-lapse seismic surveys may

be acquired annually representing a significant long-term ongoing cost unless permanent arrays

or more affordable forms of time-lapse tests become available. The exact nature and amount of

expenses will depend on the regulatory regime for the hand-over of MMV to governments and

the structure of a remediation or insurance fund. The costs of MMV can be estimated by

considering recommended practice guidelines such as the DNV Recommended Practice (RP)

J203, DNV Service Specification (DSS) 402 and CSA Z741. Therefore, Economics rates a 2.

Demonstrated Technical Readiness scores a 3 as current available technologies can be

utilised and industry best practice guidelines such as that of DNV exist to provide best practices

guidelines and certification frameworks of MMV and cover parameters such as injection

standards (maximum injection pressures, fracture gradients), well integrity, well spacing, etc.

Environmental Impact earns a 2 rating because there is some footprint with the presence of

infrastructure such as permanent monitoring arrays and monitoring stations.

MMV scores a rating of “mid-long term” (2.4).

Geosequestration

102

Summary of Science and Technology Sub-Matrix Assessment

Certain technology elements specific to GS rate poorly (Table 3.5) because of public

concerns or technology novelty. The public concerns are related to anticipation of economic

disadvantages and/or environmental, health and safety hazards. For example, Injection and

Storage is ranked poorly due to community concerns about hazardous CO2 gas build-up;

however, Capture is rated highly because there is no perceived risk, and it beneficially prevents

industrial-generated CO2 from entering the atmosphere and raising GHG levels. Technology

novelty relates to the transferability of existing CCS knowledge and technology from the O&G

industry to other high emitter industries, such as electricity generators. Therefore my previous

expectation of a high rating was not upheld, despite forty years of relevant petroleum industry

experience with enhanced hydrocarbon recovery in North America. The Science and

Technology Sub-Matrix received a “mid-long term” (2.3) rating which is entered into the

Science and Technology column of the Master Matrix. A medium score, rather than the

expected high, for the Science and Technology evaluation criterion does not diminish the

overall assessment of GS.

Table 3.5. Science and Technology Sub-Matrix Scorecard: Results of Application of Methodology to

Geosequestration.

3.3.7 Summary of Results: Overall Technology Rating

The overall rating for GS at 2.3 (Table 3.6), as detailed above, suggests a medium to high

probability of success as a commercial-scale CO2 storage option. This is consistent with recent

trends in CO2 GS, as well as published government and industry experience.

Table 3.6. Overall risk assessment rating of Geosequestration in Master Matrix.

Geosequestration

103

3.4 DISCUSSION

Observations

My assessment of GS from a general global perspective yielded a number of observations

which should be taken into account when planning a carbon management portfolio or project:

1. Public Acceptance (particularly for onshore storage and transportation pipelines) and

economics are identified as the most critical barriers to deployment of GS on a wide

scale commercial level.

2. A high level of interrelationship between some of the Evaluation Criteria is apparent,

i.e. Public and Political Acceptance very strongly affect the establishment of Regulatory

Frameworks, which in turn impacts Economics (higher planning uncertainty, risk of

delays, limitation of available storage sites or transport solutions). Although progress

has been made in the establishment of national and regional legislation and regulation,

the lack of a unified global framework reflects a lack of unity by the global community

and slows progress on adoption.

3. A clear need for regulatory guidelines was identified including mandatory compliance

and economic incentives for heavy emitters but also in areas of MMV and liability

transfer. Current commercially viable business models are limited to EOR, with most

other projects still heavily reliant on government subsidies or incentives.

4. The evaluation of project locations requires effective source-to-sink matching

procedures, balancing economics and public acceptance, to manage infrastructure and

logistics costs. Although offshore storage sites compare unfavourably with onshore in

terms of economics, higher levels of public acceptance and lower levels of resource

conflicts and land use competition are favourable.

5. A broader range of business models is required to make projects more economically

robust and less vulnerable to policy changes. While R&D efforts are focused on

lowering costs and improving efficiencies of capture technologies, the long term

ongoing costs of MMV should not be underestimated. Although many tools are

commercially available to monitor geological, geophysical and geochemical data, these

tools must be modified to act as low-cost, minimal manpower and footprint sentinel

monitoring systems that are fit-for-purpose to suit a variety of environmental scenarios.

The IEA Greenhouse Gas R&D Programme’s Overview of Monitoring Projects for

Geosequestration

104

Geologic Storage Projects Report (2004) describes a methodology for the operational

monitoring of the CO2 plume through time-lapse EM, gravity and seismic surveys over

a thirty year period (plant/project operation) and post-closure monitoring for the

following fifty years (for a total of eighty years).

6. According to the IPCC Special Report on CCS (2005), the length of sequestration time

is ideally permanent. Permanence is a concern because O&G technologies have not

previously been deployed at such time scales, as long-term storage of the CO2 has not

been the petroleum industry’s priority. Demonstrations of the different CO2 trapping

mechanisms for the reservoir categories and lithologies need more research to assess the

long-term effects of CO2 interactions in geological formations, as lessons learned from

O&G projects regarding this cannot necessarily be translated to long-term sequestration.

7. Transportation solutions require careful evaluation, including the reutilisation of

existing pipeline networks and ship designs. The latter requires more RD&D to capture

economies of scale for CO2 transportation, with one option the dual-purpose built CO2

and LNG transportation vessels. A possible combination of both offshore pipelines for

near shore locations, and ships (long distance off-shore) could be the most economic

transportation solution. Transportation by ship is flexible, scalable and may be more

cost-effective than offshore pipelines for large distances (Barrio et al., 2004). This

mode of transport is most economical at distances between 700km – 1500km, and

requires less CAPEX for infrastructure costs (IEA, 2008a). It can provide a better

solution to deal with scattered emissions source points of CO2 and distant geological

storage sites. When considering combined ship transport, LNG/LPG vs. CO2,

compression, as well as refrigeration, are needed for ships designed to transport CO2

(Aspelund et al., 2004). Currently, there are only a few dedicated CO2 vessels with a

capacity of only 1000-1500m3 for liquefied food-grade CO2, providing a total shipped

volume of 1MtCO2 (IPCC, 2005). Research, design and development of mid-range

ships are underway in Japan (Chiyoda Corporation, 2011) and Norway. A Norwegian

example is the collaboration between Statoil Hydro and SINTEF, which aims to

develop four to six tank vessels with 20,000 m3 liquefied CO2 capacity (~24kt CO2) and

injection turrets (Rackley, 2010e).

8. The recently released global standards (DNV certification framework (DNV, 2012a,

2012b) and CSA Z741 (CSA IPAC-CO2, 2012) provide internationally recognised

process methodologies for CO2 storage site selection, risk assessment, monitoring and

verification based on industry best practice and intergovernmental directives (Aarnes et

Geosequestration

105

al., 2009). These are important because international standards and experience help to

increase confidence in the technical feasibility and operability of CCS in geological

formations, site integrity, and support economic auditing.

9. Technologies are available but need to be improved for monitoring the subsurface. For

instance, seismic imaging is proven for both terrestrial and offshore environments, but

the quality of image resolution at low frequencies could be improved to use seismic data

in a quantitative manner to evaluate CO2 distribution. Both CSEM and gravity are

sensitive to CO2 volume, however, both require more research and demonstration. New

tools, techniques and technology are needed for commercial implementation as some

technology gap exists for remote sensing technologies, such as anti-corrosive equipment

and material suitable for long-term CO2 exposure in sub-surface environments

(pressure, temperature) (WA ERA, 2010). The goal is to establish passive sentinels to

minimise cost, manpower and environmental footprint with strategically placed

samplers, such as soil-gas, and permanent arrays (seismic monitoring). This requires

remote and WIFI instruments to improve data acquisition and send data back to a

central location.

Suggestions to Improve Uptake

I suggest the following as areas/means to improve or accelerate the wide-spread commercial

deployment of GS:

Early engagement and empowerment of local stakeholders by addressing concerns (e.g.

legacy issues, like public opposition to existing plants), potential risks and highlighting

socio-economic benefits to the local community such as employment opportunities and

stimulation of the local economy (Miller et al., 2008; de Best-Waldhober et al., 2009;

Anderson et al., 2007; Roberts and Mander, 2011; Reiner, 2008). Scientists/academics

could be utilised as providers of information, as they are perceived as more credible,

trustworthy and independent than government and industry (Global CCS Institute,

2011b; European Commission, 2011).

Placing GS and the CCS industry in the context of climate change as an important part

of a diverse, low-carbon, energy mix/strategy for energy security as well as climate

change mitigation to increase levels of public acceptance (Ashworth et al., 2009, 2010;

Anderson et al., 2007; European Commission, 2011), i.e. not at the expense of

renewable energy solutions (Shackley et al., 2009; Itaoka et al., 2004).

Geosequestration

106

Improved cooperation and collaboration between stakeholders to educate and make

more informed decisions, e.g. counter misinformation about scientific concepts, learn

from shared experience, and scale up technology transfer from the petroleum to other

heavy emitter industries and within the CCS industry. Regulators and operators should

also be better prepared to educate and develop a working relationship with the media

and environmental non-governmental organisations (ENGO’s) (Wong-Parodi et al.,

2008).

The effective use of media, especially online social media, which can be quickly and

effectively utilised to communicate a series of successful case histories of good practice

or any unexpected leakage to prove efficiency of early detection and mitigation

strategies and develop an atmosphere of transparency with stakeholders.

A balance between financial support mechanisms and international standard of

regulations to provide legal certainty and a flexible regulatory framework that is

adaptable to changing circumstances. This should include faster, more efficient and

transparent permit approval and bureaucratic processes.

Suggested means to improve business models:

The capture and reutilisation of CO2 for hydrocarbon recovery and production of

unconventionals (methane, previously-fracked brown fields, shale beds,

geothermal). Recent studies discuss the possibility that the hydraulic fracturing

industry could extract more natural gas from spent fields by injecting liquid or

supercritical CO2 instead of pressurised water into methane-rich shale formations

(Ishida et al., 2012). The CO2 is better at fracturing the rock, releasing the methane,

even in previously fracked shale formations, so the rock releases more methane in

favour of the more chemically attractive CO2 which would remain stored in the

shale (McKenna, 2012). A combination of fracking with CCS to maximise recovery

from depleted reservoirs improves project economics, decreases the volume of

water needed, and encourages the commercial adoption of GS by the petroleum

industry.

Cost reductions in capture and monitoring technologies; e.g., the recent

development of a cheaper more efficient non-toxic absorber instead of expensive

amines (Wickham, 2012).

Improved source-sink matching, and matching states or countries with an

abundance of available storage sites with project owners and emitters with a lack of

Geosequestration

107

same. Leasing of the storage space could provide an additional revenue stream for

those with excess capacity. The CO2 could be transported by ship as a method to

counter the distance problem.

Shared infrastructure to improve economies of scale, increase efficiency and decrease

risk, e.g. R&D knowledge sharing and infrastructure hubs. This may require the

overcoming of technical challenges, such as the different quality of CO2 coming from a

variety of emissions sources when utilising a pipeline hub-and-spoke network.

3.5 CONCLUSIONS

I apply my new methodology to assess GS as a carbon mitigation option and the barriers to

its wide scale adoption. My analysis suggests that GS has a medium to high probability of

success as a commercial-scale CO2 storage option. GS will form an integral part of any

sequestration portfolio because of its technical capability to quickly capture and store large

volumes of CO2 in a number of locations worldwide and at the volumes necessary to mitigate

GHG levels. The high quality of storage (large available global capacity, ability to inject high

volumes in a short timeframe, long-term sequestration duration, and available MMV capability)

is a major advantage of GS as a carbon mitigation option. This technology allows for continued

access to fossil fuels resources and utilisation of existing energy generating and industrial

infrastructure with industrial scale deployment anticipated in the medium term. The comparably

well-understood interactions taking place at the storage site in comparison to other CCS

technologies, and the localised environmental impact of projects make GS a suitable bridging

technology for global carbon management for the next fifty years until renewables are available

at commercial scales. Recent research indicates the potential to utilise CO2 for unconventional

hydrocarbon production (methane, shale gas, geothermal) thereby strengthening the economics

of GS by expansion of the carbon capture utilisation and storage (CCUS) business case beyond

EOR. GS should be an essential component of any low-energy carbon management portfolio

strategy for regulators and industry.

Geosequestration

108

Geosequestration

109

REFERENCES

Aarnes, J.E., Selmer-Olsen, S., Carpenter, M.E., and Flach, T.A. (2009), Towards guidelines for

selection, characterization and qualification of sites and projects for geological storage of CO2: Energy

Procedia, 1, 1, 1735–1742.

Aimard, N., Lescanne, M., Mouronval, G., and Prebende, C. (2007), The CO2 pilot at Lacq: An

integrated oxycombustion CO2 capture and geological storage project in the South West of France:

Presented at the International Petroleum Technology Conference.

Aldy, J.E. and Stavins, R.N., (2011) The Promise and Problems of Pricing Carbon: Theory and

Experience: Discussion Paper 2011-12, Cambridge, Mass, USA.

American Electric Power Service Corporation [AEP] (2011), CCS Business Case Report:

Mountaineer Commercial Scale Carbon Capture and Storage (CCS) Project: American Electric

Power [Tamms, K. and Gilliland, D.], Project No. PRO 004.

Anderson, J., de Coninck, H., Curnow, P., Flach, T., Groenenberg, H., Norton, C., Reiner, D., Shackley,

S., Upham, P., Eldevik, F., and Sigurthorsson, G. (2007), Multidisciplinary analysis and gap-filling

strategies: The ACCSEPT Project, Report No. BRINO912OGSIG2007-1242.

Angus, S., Armstrong, B., and de Reuck, K.M. (1973), International Thermodynamic Tables of the Fluid

State Volume 3 Carbon Dioxide, International Union of Pure and Applied Chemistry, Division of

Physical Chemistry, Pergamon Press, London, 266–359.

Ashworth, P., Boughen, N., Mayhew, M., and Millar, F. (2009), An integrated roadmap of

communication activities around carbon capture and storage in Australia and beyond: Energy

Procedia, 1, 1, 4749-4756.

Ashworth, P., Boughen, N., Mayhew, M., and Millar, F. (2010), From research to action: Now we have

to move on CCS communication: International Journal of Greenhouse Gas Control, 4, 2, 426-433.

Aspelund, A., Sandvik, T.E., Krogstad, H., and De Koeijer, G. (2004), Liquifaction of captured CO2 for

ship-based transportation: in Proceedings of the 7th International Conference on Greenhouse Gas

Control Technologies (GHGT-7).

Bachu, S. (2003), Screening and ranking of sedimentary basins for sequestration of CO2

in geological

media in response to climate change: Environmental Geology, 44, 277-289.

Geosequestration

110

Bachu, S. (2008), CO2 storage in geological media: Role, means, status and barriers to deployment,”

Progress in Energy and Combustion Science, 34, 254–273.

Barrio, M., Aspelund, A.,Weydahl, T., Sandvik, T.E., L.R. Wongraven, L.R., Krogstad, H., and

Henningsen, R., Mølnvik, M., and Eide, S.I. (2004), Ship-based transport of CO2: in Proceedings of

7th International Conference on Greenhouse Gas Control Technologies (GHGT-7).

Batzle, M. and Wang, Z. (1992), Seismic properties of pore fluids: Geophysics, 57, 11, 1396-1408.

Baxter, L., Baxter, A. and Burt, S. (2009), Cryogenic CO2 Capture as a Cost-Effective CO2 Capture

Process: Presented at International Pittsburgh Coal Conference.

BBC News Online [BBC] (2011a), Concern grows over Longannet carbon capture plans: BBC News

Online, 6 October 2011. http://www.bbc.co.uk/news/uk-scotland-scotland-business-15207268

(accessed 9 October 2012).

BBC News Online [BBC] (2011b), Longannet carbon capture scheme scrapped: BBC News Online

19 October 2011. http://www.bbc.co.uk/news/uk-scotland-north-east-orkney-shetland-15371258

(accessed 9 October 2012).

Beckworth, R. (2011), Carbon Capture and Storage: A Mixed Review, Journal of Petroleum Technology,

63, 5, 42-45.

Bellona (2013a), The UK Labour party introduces an amendment to require CCS for new gas power

plants: Bellona Environmental CCS Team, 1 February 2013. http://bellona.org/ccs/ccs-news-

events/news/article/the-uk-labour-party-introduces-an-amendment-to-require-ccs-for-new-gas-power-

plants.html (accessed 12 February 2013).

Bellona (2013b), New EU CCS Policy: "The most important document that the European Commission

has produced on CCS for the last two years", says Frederic Hauge: The Bellona Foundation,13

February 2013. http://www.bellona.org/articles/articles_2012/1358108616.37 (accessed 12 February

2013).

Bergmann, P., Schmidt-Hattenberger, C., Kiessling, D., Rücker, C., Labitzke, T., Henninges, J,

Baumann, G. and Schütt, H. (2012), Surface-downhole electrical resistivity tomography applied to

monitoring of CO2 storage at Ketzin, Germany: Geophysics, 77, 6, B253-B267.

Bradshaw, J., Bachu, S., Bonijoly, D., Burruss, R., Holloway, S., Christensen, N.P., and Mathiassen,

O.M., (2007), CO2 storage capacity estimation: Issues and development of standards: International

Journal of Greenhouse Gas Control, 1, 1, 62–68.

Geosequestration

111

Burt, S., Baxter, A. and Baxter, L. (2009), Cryogenic CO2 Capture to Control Climate Change

Emissions: in Proceedings of the 34th International Technical Conference on Clean Coal and Fuel

System: The Clearwater Clean Coal Conference.

Calhoun Jr., J.C. (1982), Fundamentals of Reservoir Engineering: University of Oklahoma Press:

Norman, OK, USA.

Carbon Storage Taskforce (2009), National Carbon Mapping and Infrastructure Plan – Australia:

Concise Report: Department of Resources, Energy and Tourism, Canberra, ACT. Australia.

Chevron Australia (2012), Injecting Billions of Dollars across Australia: Chevron Australia Pty. Ltd.

http://www.chevronaustralia.com/ourbusinesses/gorgon.aspx (accessed 20 June 2012).

Chiyoda Corporation (2011), Preliminary Feasibility Study on CO2 Carrier for Ship-based CCS,

Chiyoda Corporation, Report supported by Global CCS Institute.

http://www.globalccsinstitute.com/publications/preliminary-feasibility-study-co2-carrier-ship-based-

ccs (accessed 21 November 2013).

Cooperative Research Centre for Greenhouse Gas Technologies [CO2CRC] (2008), Storage

Capacity Estimation, Site Selection and Characterisation for CO2 Storage Projects: Report No.

RPT08-1001 [Kaldi, J. G., and Gibson-Poole, C. M. (Eds.)], CO2CRC, Canberra, ACT, Australia.

Cooperative Research Centre for Greenhouse Gas Technologies [CO2CRC] (2012), New system

reduces carbon capture costs by $20 per tonne : CO2CRC [Media release] (6 June 2012).

http://co2crc.com.au/dls/media/12/UNOClearwater.pdf (accesssed 17 June 2012).

Congressional Research Service [CRS] (2007), CRS Report for Congress: Underground Carbon

Storage: Frequently Asked Questions: [Folger, P.] Order Code RL34218.

Congressional Research Service [CRS] (2009), CRS Report for Congress: Carbon Dioxide (CO2)

Pipelines for Carbon Sequestration: Emerging Policy Issues: [Parfomak, P.W., Folger, P., Vann, A.]

Order Code RL33971.

Cook, P.J. (2012), Clean energy, climate and carbon: Carbon Management, 3, 3, 259–263.

CSA Group and IPAC-CO2 Research Incorporated [CSA IPAC-CO2] (2012), CSA Group and IPAC-

CO2 Research Inc. Announce World’s First Bi-National Standard for Geologic Storage of Carbon

Dioxide: [news release] 15 November 2012, CSA Group IPAC-CO2: Regina and Toronto, Canada.

http://www.ipac-co2.com/uploads/File/PDFs/CSA%20IPAC-

CO2%20CCS%20Standard%202012%20Final%20News%20Release%20v1%20copy.pdf (accessed

30 November 2012).

Geosequestration

112

de Best-Waldhober, M., Daamen, D., and Faaij, A. (2009), Informed and uninformed public opinions on

CO2 capture and storage technologies in the Netherlands: International Journal of Greenhouse Gas

Control, 3, 3, 322-332.

Det Norske Veritas [DNV] (2010), Recommended Practice DNV-RP-J202: Design and Operation of

CO2 Pipelines.

Det Norske Veritas [DNV] (2012a), Recommended Practice DNV-RP-J203: Geological Storage of

Carbon Dioxide.

Det Norske Veritas [DNV] (2012b), DNV Service Specification DNV-DSS-402: Qualification

Management for Geological Storage of CO2.

Eiken, O., Zumberge, M., Torkjell, T., Sasagawa, G., and Nooner, S. (2004), Gravimetric monitoring of

gas production from the Troll Field: Presented at SEG 74th

Annual Meeting.

Eiken, O., Ringrose, P., Hermanrud, C., Nazarian, B., Torp, T.A., and Loier, L. (2011), Lessons learned

from 14 years of CCS operations: Sleipner, In Salah and Snohvit: Energy Procedia, 4, 5541-5548.

European Parliament and the Council of the European Union [EU] (2001), Directive 2001/80/EC of

the European Parliament and of the Council of the European Union on the limitation of emissions of

certain pollutants into the air from large combustion plants: Official Journal of the European Union,

23 October 2001.

European Parliament and the Council of the European Union [EU] (2009), Directive 2009/31/EC of

the European Parliament and of the Council on the geological storage of carbon dioxide: Official

Journal of the European Union, 23 April 2009.

European Commission (2011), Special Eurobarometer 364: Public awareness and acceptance of CO2,

capture and storage: [TNS Opinion and Social], Directorate-General for Energy, Brussels, Belgium.

European Technology Platform for Zero Emission Fossil Fuel Power Plants [ZEP] (2011), The Costs

of CO2 Capture, Transport and Storage: Post-demonstration CCS in the EU: ZEP.

http://www.zeroemissionsplatform.eu/library/publication/165-zep-cost-report-summary.html

(accessed 1 March 2012).

Feron, P. H. and Jansen, A. E. (2002), CO2 separation with polyolefin membrane contactors and

dedicated absorption liquids: performances and prospects: Separation and Purification Technology,

27, 3, 231-242.

Geosequestration

113

Gale, J. and Davison, J. (2004), Transmission of CO2—safety and economic considerations: Energy 29,

9, 1319–1328.

Gasperikova, E., and Hoversten, G.M. (2006), A feasibility study of nonseismic geophysical methods for

monitoring geologic CO2 sequestration: The Leading Edge, 25, 10, 1282-1288.

Gilfillan, S.M.V., Lollar, B. S., Holland, G., Blagburn, D., Stevens, S., Schoell, M., Cassidy, M., Ding,

Z., Zhou, Z., Lacrampe-Couloume, G., and Chris J. Ballentine, C.J. (2009), Solubility trapping in

formation water as dominant CO2 sink in natural gas fields: Nature, 458, 7852, 614-618.

Global CCS Institute (2011a), The Global Status of CCS: 2010, Global CCS Institute, Canberra, ACT,

Australia.

Global CCS Institute (2011b), Communicating the risks of CCS, Report [Bradbury, J., Greenberg, S.,

and Wade, S.]: Global CCS Institute, Canberra, ACT, Australia.

Global CCS Institute (2011c), Global status of large-scale integrated CCS projects: June 2011 update:

Global CCS Institute, Canberra, ACT, Australia.

Global CCS Institute (2011d), The Costs of CCS and Other Low-Carbon Technologies: Issues Brief

2011, No. 2: [Abellera, C., and Short, C.]: Global CCS Institute, Canberra, ACT, Australia.

Gorecki, C.D., Hamling, J.A., Klapperich, R.J., Steadman, E.N., and Harju, J.A. (2012), Integrating CO2

EOR and CO2 Storage in the Bell Creek Oil Field: Presented at Carbon Management Technology

Conference 2012, Paper No. CMTC 151476.

Government of Western Australia (2011), Collie-South West CO2 Geosequestration Hub – Project and

Activity Progress Report for the Global Carbon Capture and Storage Institute: Government of

Western Australia, Department of Mines and Petroleum, Perth, Western Australia, Australia: Report

No. DMPNOV11_1578.

Gunter, W.D., Perkins, E.H. and McCann, T.J. (1993), Aquifer disposal of CO2-rich gases: reaction

design for added capacity: Energy Conversion and Management, 34, 9-11, 941-948.

Gunter, W. D., Wiwehar, B., and Perkins, E. H. (1997), Aquifer disposal of CO2-rich greenhouse gases:

extension of the time scale of experiment for CO2-sequestering reactions by geochemical modelling:

Mineralogy and Petrology, 59, 1-2, 121-140.

Haszeldine, S. (2011), Why the UK buried a world-first carbon-capture scheme: New Scientist: Opinion,

21 October 2011. http://www.newscientist.com/article/dn21080-why-the-uk-buried-a-worldfirst-

carboncapture-scheme.html?full=true (accessed 7 June 2012).

Geosequestration

114

Helgesen, O.K. (2010), CO2-problemer på Snøhvit : Statoil klarer ikke å lagre all CO2-en fra Snøhvit:

Teknisk Ukeblad, 10 March, 2010. http://www.tu.no/olje-gass/2010/03/10/co2-problemer-pa-snohvit

(accessed 28 February 2013).

Helseth, J. (2011), CCS in Germany – is it game over?: Global CCS Institute,

http://www.globalccsinstitute.com/community/blogs/authors/jonashelseth/2011/10/10/ccs-germany-it-

game-over (accessed 7 June 2012)

Hill, B., Hovorka, S., and Metzer, S. (2013), Geologic carbon storage through enhanced oil recovery: in

Proceedings of the 11th International Conference on Greenhouse Gas Control Technologies (GHGT-

11). http://www.fossiltransition.org/filebin/images/GHGT11_Hill_Hovorka_Melzer-2.pdf (accessed

1 April 2013) [in press].

Holloway, S. (2001), Storage of fossil fuel-derived carbon dioxide beneath the surface of the earth:

Annual Review of Energy and the Environment, 26, 1, 145-166.

Holtz, M.H. (2003), Optimization of CO2 sequestered as a residual phase in brine-saturated formations:

In Briefing book: Second Annual Conference on Carbon Sequestration: developing & validating the

technology base to reduce carbon intensity, Gulf Coast Carbon Center [GCCC] Digital Publication

Series 03-03, unpaginated.

Hooper, B. (2011), Transport: CO2CRC Education and Training: The 2011 CO2CRC CCS School, The

University of Melbourne, Victoria, Australia [unpublished].

Hosa, A., Esentia, M., Stewart, J., and Haszeldine, S. (2011), Injection of CO2 into saline formations:

Benchmarking worldwide projects: Chemical Engineering Research and Design, 89, 1855-1864.

Hovorka, S.D., Meckel, T.A., Trevino, R.H., Lu, J., Nicot, J.P., Choi, J.W., Freeman, D., Cook, P.,

Daley, T.M., Ajo-Franklin, J.B., Freifeild, B.M., Doughty, C., Carrigan, C.R. La Brecque, D.,

Kharaka, Y.K., Thordsen, J.J., Phelps, T.J. Yang, C. Romanak, K.D., Zhang, T., Holt, R.M., Lindler,

J.S., and Butsch, R. J. (2011), Monitoring a large volume CO2 injection: Year two results from

SECARB project at Denbury’s Cranfield, Mississippi, USA: Energy Procedia, 4, 3478-3485.

ICIS Heren (2011), Dutch CCS in disarray as ‘on land’ storage ruled out: ICIS Heren, 15 February 2011.

http://www.icis.com/heren/articles/2011/02/15/9435644/dutch-ccs-in-disarray-as-on-land-storage-

ruled-out.html (accessed 15 June 2012).

International Energy Agency [IEA] (2008a), CO2 capture and storage: A key carbon abatement option:

International Energy Agency, OECD, Paris, France.

Geosequestration

115

International Energy Agency [IEA] (2008b), World Energy Outlook 2008: International Energy

Agency, OECD, Paris, France.

International Energy Agency [IEA] (2009), Technology Roadmap – Carbon Capture and Storage:

International Energy Agency, OECD, Paris, France.

International Energy Agency [IEA] (2011), Carbon Capture and Storage Legal and Regulatory

Review: Edition 2: International Energy Agency, OECD [Garrett, J., Beck, B., Lipponen, J., McCoy,

S., Houssin, D., Volk, D., Yoshimura, T., Finkenrath, M. and Gaghen, R.]: Paris, France.

International Energy Agency [IEA] (2012), Energy Technology Perspectives 2012: International

Energy Agency, OECD, Paris, France.

International Energy Agency Greenhouse Gas R&D Programme [IEA GHG] (Link 1), Natural

Releases of CO2: IEA Greenhouse Gas R&D Programme [Mike Stenhouse, Monitor Scientific

(compiler)] Cheltenham, Gloucestershire, UK.

International Energy Agency Greenhouse Gas R&D Programme [IEA GHG] (2004), Overview of

Monitoring Requirements for Geologic Storage Projects: IEA Greenhouse Gas R&D Programme,

Report No. PH4/29.

International Energy Agency Greenhouse Gas R&D Programme [IEA GHG] (2005), Building the

Cost Curves for CO2 Storage: European Sector, Report 2005/02, IEA Greenhouse Gas R&D

Programme. http://www.ieaghg.org/docs/overviews/2005-2.pdf (accessed 19 July 2012).

International Energy Agency Greenhouse Gas R&D Programme [IEA GHG] (2007), Capturing

CO2, IEA Greenhouse Gas R&D Programme [Adams, D. and Davison, J.], Cheltenham,

Gloucestershire, UK.

International Energy Agency Greenhouse Gas R&D Programme [IEA GHG] (2008), Geologic

Storage of Carbon Dioxide: Staying Safely Underground, IEA Greenhouse Gas R&D Programme,

[Price, J., and Smith, B.], Virginia, USA.

International Energy Agency Greenhouse Gas R&D Programme [IEA GHG] (2009), Natural and

Industrial Analogues for Geological Storage of Carbon Dioxide. IEA Greenhouse Gas R&D

Programme [Stenhouse, M. (compiler)], Cheltenham, Gloucestershire, UK.

International Energy Agency Greenhouse Gas R&D Programme [IEA GHG] (2012), Capture,

Transport and Storage of CO2, IEA Greenhouse Gas R&D Programme, Cheltenham, Gloucestershire,

UK

Geosequestration

116

Intergovernmental Panel on Climate Change [IPCC] (2005), Intergovernmental Panel on Climate

Change Special Report on Carbon Dioxide Capture and Storage: Prepared by Working Group III of

the Intergovernmental Panel on Climate Change [Metz, B.O., Davidson, H., de Coninck, C., Loos, M.

and Meyer, L.A. (eds.)]: Cambridge University Press, Cambridge, United Kingdom and New York,

NY, USA.

Intergovernmental Panel on Climate Change [IPCC] (2007), Climate Change 2007: Mitigation:

Contribution of Working Group III to the Fourth Assessment Report of the Intergovernmental Panel

on Climate Change, [Metz, B., Davidson, O., Bosch, P., Dave, R. and Meyer, L. (eds.)], Cambridge

University Press, Cambridge, UK and New York, NY, USA.

Ishida, T., Aoyagi, K., Niwa, T., Chen, Y., Murata, S., Chen, Q., and Nakayama, Y. (2012), Acoustic

emission monitoring of hydraulic fracturing laboratory experiment with supercritical and liquid CO2:

Geophysical Research Letters, 39, 16.

Itaoka, K., Saito, A., and Akai, M. (2004), Public acceptance of CO2 capture and storage technology: A

survey of public opinion to explore influential factors: in Proceedings of the 7th International

Conference on Greenhouse Gas Control Technologies (GHGT-7).

Itaoka, K., Saito, A., Paukovic, M., de Best-Waldhober, M., Dowd, A-M., Jeanneret, T., Ashworth, P.

and James, M. (2012), Understanding how individuals perceive carbon dioxide: Implications for

acceptance of carbon dioxide capture and storage: CSIRO Report No. EP 118160, Australia.

Jenkins, C.R., Cook, P.J., Ennis-King, J., Undershultz, J., Boreham, C., Dance, T., de Caritat, P.,

Etheridge, D.M., Freifeld, B.M., Hortle, A., Kirste, D., Paterson, L., Pevzner, R., Schact, U., Sharma,

S., Stalker, L., and Urosevic, M. (2011), Safe storage and effective monitoring of CO2 in depleted gas

fields: in Proceedings of the National Academy of Sciences (PNAS) Early Edition.

http://www.pnas.org/content/early/2011/12/16/1107255108.full.pdf+html (accessed 9 October 2012).

Kaldi, J. (2011), Site Selection and Characterisation I: Capacity, trapping and injectivity: CO2CRC

Education and Training: The 2011 CO2CRC CCS School, The University of Melbourne, Victoria,

Australia [unpublished].

Kessler, R.A. (2009), Midwest group cancels Ohio CCS project despite DOE funding: Recharge, 21

August 2009. http://www.rechargenews.com/business_area/innovation/article186348.ece (accessed 30

May 2012).

Ketzer, J. M., Iglesias, R., Einloft, S., Dullius, J., Ligabue, R., and De Lima, V. (2009), Water–rock–

CO2 interactions in saline aquifers aimed for carbon dioxide storage: Experimental and numerical

Geosequestration

117

modeling studies of the Rio Bonito Formation (Permian), southern Brazil: Applied Geochemistry, 24,

5, 760-767.

Kopp, A., Class, H. and Helmig, R. (2009a), Investigations on CO2 Storage Capacity in Saline Aquifers

Part 1: Dimensional Analysis of Flow Processes and Reservoir Characteristics: International Journal

of Greenhouse Gas Control, 3, 263–276.

Kopp, A., Class, H., and Helmig, R. (2009b), Investigations on CO2 Storage Capacity in Saline Aquifers

Part 2: Estimation of Storage Capacity Coefficients: International Journal of Greenhouse Gas

Control, 3, 277–287.

Kuijper, M. (2011), Public acceptance challenges for onshore CO2 storage in Barendrecht: Energy

Procedia, 4, 6226-6233.

Kumar, A., Noh, M., Ozah, R., Pope, G., Bryant, S., Sepehrnoori, K., and Lake, L. (2005), Reservoir

Simulation of CO2 Storage in Aquifers: SPE Journal, 10, 3, 336-348.

Lake, L.W. (1989), Enhanced Oil Recovery. Prentice Hall: Englewood Cliffs, NJ, USA.

Lumley, D. (2001), Time-lapse seismic reservoir monitoring: Geophysics, 66, 50-53.

Lumley, D. (2010), 4D Seismic monitoring of CO2 sequestration: The Leading Edge, 29, 2, 150-155.

Lumley, D., Sherlock, D., Daley, T., Huang, L., Lawton, D., Masters, R., Verliac, M., and White, D.,

(2010), Highlights of the 2009 SEG summer research workshop on CO2 sequestration geophysics: The

Leading Edge, 29, 2, 138-144.

Lumley, D. (2011), 4D Seismic Monitoring of CO2 Sequestration – Theory and Practice: GeoProc2011,

Paper No. GP059.

MacDowell, N., Florin, N., Buchard, A., Hallett, J., Galindo, A., Jackson, G., Adjiman, C.S., Williams,

C.K., Shah, N. and Fennell, P. (2010), An overview of CO2 capture technologies: Energy and

Environmental Science, 3, 1645-1669.

McClure, M. (2012), Carbon capture and storage likely to cause earthquakes, say Stanford researchers:

Stanford Report, June 19, 2012. http://news.stanford.edu/news/2012/june/carbon-capture-earthquakes-

061912.html (accessed 20 June 2012).

McKenna, P. (2012), Fracking could be combined with carbon capture plans: New Scientist:

Environment, 31 August 2012. http://www.newscientist.com/article/dn22232-fracking-could-be-

combined-with-carbon-capture-plans.html (accessed 9 September 2012).

Geosequestration

118

Miller, E., Summerville, J., Buys, L., and Bell, L. (2008), Initial public perceptions of carbon

sequestration: implications for engagement and environmental risk communication strategies:

International Journal of Global Environmental Issues, 8, (1), 147-164.

Molde, A.I. (2011), Reservoaret klarer ikke mer Snøhvit-CO2: Petro.no: 20 May 2011.

http://www.petro.no/modules/module_123/proxy.asp?C=14&I=16697&D=2&mid=195 (accessed 26

May 2011).

North American Carbon Storage Atlas [NACSA] (2012), The North American Carbon Storage Atlas

2012: First Edition. http://www.netl.doe.gov/technologies/carbon_seq/refshelf/NACSA2012.pdf

(accessed 17 July 2012).

National Audit Office (2012), Department of Energy and Climate Change: Carbon Capture and

Storage: lessons from the competition for the first UK demonstration: Report by the Comptroller and

Auditor General, National Audit Office, UK, Report No. HC 1829 Session 2010-2012.

National Research Council (2012), Induced Seismicity Potential in Energy Technologies: The National

Academies Press, Washington, DC.

National Energy Technology Laboratory, Department of Energy [NETL DOE] (2010), Carbon

Sequestration Atlas of the United States and Canada – Third Edition (Atlas III): Appendix B -

Summary of the Methodology for Development of Geologic Storage Estimates for Carbon Dioxide:

National Energy Technology Laboratory (NETL), Department of Energy (DOE).

http://www.netl.doe.gov/technologies/carbon_seq/refshelf/atlasIII/2010AtlasIII_AppendixB.pdf

(accessed 19 July 2012).

National Energy Technology Laboratory, United States Department of Energy [NETL DOE]

(2011), Archer Daniels Midland Company: CO2 Capture from Biofuels Production and Sequestration

into the Mt. Simon Sandstone: [Project Fact Sheet]: National Energy Technology Laboratory, United

States Department of Energy,

http://sequestration.mit.edu/tools/projects/DOE%20projects/Industrial%20CCS%20projects/Archer-

Midland-Daniels-Tech-Update-2011.pdf (accessed 17 April 2013).

Obdam, A., L.G.H. Van der Meer, May, F., Bech, N., Kervevan, C. and Wildenberg, A., (2003),

Effective CO2 Storage Capacity in Aquifers and in Hydrocarbon Fields: in Proceedings of the 6th

International Conference on Greenhouse Gas Control Technologies (GHGT-6), [Gale J. and Kaya, Y.

(eds.)], 1; 339-344.

Pacala, S., and Socolow, R. (2004), Stabilization Wedges: Solving the Climate Problem for the Next 50

Years with Current Technologies: Science, 305, 5686, 968-972.

Geosequestration

119

Parliamentary Office of Science and Technology [POST] (2009), CO2 Capture, Transport and Storage:

Postnote, June 2009, No. 335, London, United Kingdom.

Perkins, E., Czernichowski-Lauriol, I., Azaroual, M., and Durst, P. (2004), Long term predictions of CO2

storage by mineral and solubility trapping in the Weyburn Midale Reservoir: in Proceedings of the 7th

International Conference on Greenhouse Gas Control Technologies (GHGT-7), 2093-2096.

Rackley, S.A. (2010a), Carbon capture from power generation. In Carbon capture and storage,

Butterworth-Heinemann, Elsevier Inc., Burlington, MA, USA, 4, 65-93.

Rackley, S.A. (2010b), Adsorption capture systems. In Carbon capture and storage, Butterworth-

Heinemann, Elsevier Inc., Burlington, MA, USA, 7, 133 - 157.

Rackley, S.A. (2010c), Membrane separation systems. In Carbon capture and storage, Butterworth-

Heinemann, Elsevier Inc., Burlington, MA, USA, 8, 159-194.

Rackley, S.A. (2010d), Geological storage. In Carbon capture and storage, Butterworth-Heinemann,

Elsevier Inc., Burlington, MA, USA, 11, 229-266.

Rackley, S.A. (2010e), Carbon dioxide transportation. In Carbon capture and storage, Butterworth-

Heinemann, Elsevier Inc., Burlington, MA, USA, 15, 329-344.

Redhead, P.A. (1962), Thermal desorption of gases: Vacuum, 12, 4, 203-211.

Reiner, D., Curry, T., de Figueiredo, M., Herzog, H., Ansolabehere, S., Itaoka, K., Akai M., Johnsson, F.,

and Odenbergerm M. (2006), An international comparison of public attitudes towards carbon capture

and storage technologies: Presented at 8th International Conference on Greenhouse Gas Control

Technologies (GHGT-8).

Reiner, D.M. (2008), A looming rhetorical gap: A survey of public communications activities for carbon

dioxide and storage technologies: Electricity Policy Research Group, Working Papers, No. 801.

Reuters (2011), Update 2 – Vattenfall drops carbon capture project in Germany: Reuters, 5 December

2011. http://www.reuters.com/article/2011/12/05/vattenfall-carbon-idUSL5E7N53PG20111205

(accessed 7 June 2012).

Roberts, T., and Mander, S. (2011), Assessing public perceptions of CCS: Benefits, challenges and

methods: Energy Procedia, 4, 6307–6314.

Salvi, S., Quattrocchi, F., Angelone, M., Brunori, C.A., Billi, A., Buongiorno, F. Doumaz, F., Funiciello,

R. Guerra, M., Lombardi, S., Mele, F., Pizzino, L. and Salvini, F. (2000), A multidisciplinary

approach to earthquake research: implementation of a Geochemical Geographic Information System

for the Gargano site, Southern Italy: Natural Hazard, 20, 1, 255–278.

Geosequestration

120

Schlumberger Business Consulting Energy Institute [SBC Energy Institute] (2012), Leading the

Energy Transition: Bringing Carbon Capture and Storage to Market, [Soupa, O., Lajoie, B., Long,

A., and Alvarez, H.]: Schlumberger Business Consulting Energy Institute.

Schlumberger Business Consulting Energy Institute [SBC Energy Institute] (2013), Leading the

Energy Transition: Carbon Capture and Storage: Bringing Carbon Capture and Storage to Market:

Factbook Edition, January 2013 Update: Schlumberger Business Consulting Energy Institute.

Schilling, F., Borm, G., Würdemann, H., Möller, F., and Kühn, M. (2009), Status report on the first

European on-shore CO2 storage site at Ketzin (Germany): Energy Procedia, 1, 1, 2029–2035.

Scott, K.C., Parker, D.J., Trupp, M., and Clulow, B. (2010), Setting New Environmental, Regulatory, and

Safety Benchmarks: The 2009 Gorgon CO2 3D Baseline Seismic Project, Barrow Island, Western

Australia: in Proceedings of the SPE Asia Pacific Oil and Gas Conference and Exhibition, edited,

Brisbane, Queensland, Australia.

Seiersten, M. (2001), Material selection for separation, transportation and disposal of CO2. In

Proceedings Corrosion, National Association of Corrosion Engineers, Paper No. 01042.

Seevam, P., Race, J.M., and Downie, M.J. (2008), CO2 Transport UKCCSC Progress Report:

[PowerPoint presentation], Newcastle University, Newcastle, UK.

http://www.findthatpowerpoint.com/search-5083712-hDOC/download-documents-newcastle08-

ppt.htm (accessed 6 January 2012).

Selley, R.C. (1998), Elements of Petroleum Geology 2nd

Edition: Academic Press, California, USA and

London, UK.

Shackley, S., Waterman, H., Godfroij, P., Reiner, D., Anderson, J., Draxlbauer, K., de Coninck, H.,

Groenenberg, H., Flach, T. and Flagstad, O. (2007), Stakeholder Perceptions of CO2 Capture and

Storage in Europe: Results from the EU-funded ACCSEPT Survey: Main Report: The ACCSEPT

project, Manchester, United Kingdom.

Shackley, S., Reiner, D., Upham, P., de Coninck, H., Sigurthorsson, G., and Anderson, J. (2009), The

acceptability of CO2 capture and storage (CCS) in Europe: An assessment of the key determining

factors: Part 2. The social acceptability of CCS and the wider impacts and repercussions of its

implementation: International Journal of Greenhouse Gas Control, 3, 3, 344–356.

Shao, R. and Stangeland, A. (2009), Amines Used in CO2 Capture- Health and Environmental Impacts:

Bellona Report 2009: The Bellona Foundation, Oslo, Norway.

Geosequestration

121

Sheriff, R.E. (2002), Encyclopedic Dictionary of Applied Geophysics (4th

Edition, reprinted 2011),

Society of Exploration Geophysicists, Tulsa, OK, USA.

Skriung, C.S. (2011), The amine technology is qualified: Zero Emissions Resource Organisation

(ZERO), 30 August 2011. http://www.zeroco2.no/the-amine-technology-is-qualified (accessed 11

February 2013).

Streit, J., A. Siggins and B. Evans, 2005: Predicting and monitoring geomechanical effects of CO2

injection, Carbon Dioxide Capture for Storage in Deep Geologic Formations—Results from the CO2

Capture Project, v. 2: Geologic Storage of Carbon Dioxide with Monitoring and Verification,

[Benson, S.M. (ed.)], Elsevier Science, London, 751–766.

Surovtseva, D., Amin, R. and Barifcani, A. (2011), Design and operation of pilot plant for CO2 capture

from IGCC flue gases by combined cryogenic and hydrate method: Chemical Engineering Research

and Design, 89, 9, 1752–1757.

The Commonwealth of Australia (2012), The Clean Energy Legislative Package: The Commonwealth

of Australia. http://www.climatechange.gov.au/en/government/clean-energy-future/legislation.aspx

(accessed 12 August 2012).

The House of Commons (2012), Energy Bill 135: [as amended in Public Bill Committee], The House of

Commons, London, UK.

The Norwegian Petroleum Directorate (2012), CO2 Storage Atlas Norwegian North Sea, [Halland,

E.K., Johansen, W.T. and Riis, F. (eds.)], The Norwegian Petroleum Directorate, Stavanger, Norway.

http://www.npd.no/Global/Norsk/3-Publikasjoner/Rapporter/PDF/CO2-ATLAS-lav.pdf (accessed 27

August 2012).

The Parliament of Western Australia (2011), Petroleum and Geothermal Energy Legislation

Amendment Bill 2011: Draft 3: Perth, Western Australia, Australia.

http://www.dmp.wa.gov.au/documents/Petroleum_and_Geothermal_Energy_Legislation_Amendment

_Bill_2011.pdf (accessed 1 April 2013).

Torp, T. and Brown, K. (2002), CO2 Underground Storage Costs as Experienced at Sleipner and

Weyburn: in Proceedings of the 7th International Conference on Greenhouse Gas Control

Technologies (GHGT-7).

United Kingdom Energy Research Centre [UKERC] (2012), UKERC Research Project: Carbon

Capture and Storage: Realising the potential? [Watson J. (ed.), Kern, F., Gross, M., Gross, R.,

Heptonstall, P., Jones, F., Haszeldine, S., Ascui, F., Chalmers, H., Ghaleigh, N., Gibbins, J.,

Geosequestration

122

Marusson, M., Marsen, W., Rossati, D., Russell, S., Winskel, M., Pearson, P., and Araposthathis, S.]:

UK Energy Research Centre. www.ukerc.ac.uk (accessed 2 July 2012).

United Nations Industrial Development Organisation [UNIDO] (2011), Global Technology Roadmap

for CCS in Industry: Sectoral Assessment CO2 Enhanced Oil Recovery: prepared by Advanced

Resources International for United Nations Industrial Development Organisation, Arlington, VA,

USA.

United States Department of Energy [US DOE] (2012), Carbon Storage Partner Completes First Year

of CO2 Injection Operations in Illinois: United States Department of Energy, Fossil Energy Office of

Communications, 19 November 2012. http://www.fossil.energy.gov/news/techlines/2012/12056-

Carbon_Storage_Partner_Completes_F.html (accessed 31 March 2013).

Vangkilde-Pedersen, T., Anthonsen, K.L., Smith, N., Kirk, K., Neele, F., van der Meer, B., Le Gallo, Y.,

Bossie-Codreanu, D., Wojcicki, A., Le Nindre, Y-M, Hendriks, C., Dalhoff, F., Christensen, N.P.

(2009), Assessing European capacity for geological storage of carbon dioxide–the EU GeoCapacity

project, Energy Procedia, 1, 1, 2663–2670.

Vanorio, T., Mavko, G., Vialle, S., and Spratt, K. (2010), The rock physics basis for 4D seismic

monitoring of CO2 fate: Are we there yet?: The Leading Edge, 29, 2, 156-162.

Vialle, S. and Vanorio, T. (2011), Laboratory measurements of elastic properties of carbonate

rocks during injection of reactive CO2 saturated water: Geophysical Research Letters, 38, 1.

Voser, P. (2012), Shell to construct world’s first oil sands CCS project: Royal Dutch Shell, [Press

Release], 5 September 2012. http://www.shell.ca/en/aboutshell/media-centre/news-and-media-

releases/archive/2012/0905quest.html (accessed 1 April 2013).

WA Energy Research Alliance [WA ERA] (2010), Geosequestration R&D: A Report to Australian

National Low Emission Coal R&D Program, March 2010.

Wald, M.L. and Broder, J.M. (2011), Utility shelves ambitious plan to limit carbon: NY Times, 13

February 2011. http://www.nytimes.com/2011/07/14/business/energy-environment/utility-shelves-

plan-to-capture-carbon-dioxide.html?_r=2& (accessed 31 July 2012).

Wallquist, L., Visschers, V.M., and Siegrist, M. (2010), Impact of Knowledge and

Misconceptions on Benefit and Risk Perception of CCS: Environmental Science & Technology, 44, 17,

6557-6562.

Watt, J. (2010), Carbon dioxide transport infrastructure key learning and critical issues: Journal of

Pipeline Engineering, 9, 4, 213-222.

Geosequestration

123

Whitmarsh, L., Seyfang, G., and O'Neill, S. (2011), Public engagement with carbon and climate change:

To what extent is the public 'carbon capable'? Global Environmental Change, 21, 1, 56-65.

Whittaker, S.G. (2004), Investigating geological storage of greenhouse gases in southeastern

Saskatchewan: The IEA Weyburn CO2 Monitoring and Storage Project: in Summary of Investigations:

Saskatchewan Geological Survey, Saskatchewan Industry Resources, Misc. Rep. 2004-4.1, Paper A-2.

Whittaker, S., Rostron, B., Hawkes, C., Gardner, C., White, D., Johnson, J., Chalaturnyk, R., and

Seeburger, D. (2011), A decade of CO2 injection into depleting oil fields: monitoring and research

activities of the Weyburn-Midale CO2 Monitoring and Storage Project:Energy Procedia, 4, 6069-

6076.

Wickham, C. (2012), Boost for carbon capture from new non-toxic absorber: Reuters, 23 September

2012. http://uk.reuters.com/article/2012/09/23/us-science-carbon-emissions-

idUKBRE88M0BM20120923 (accessed 30 September 2012).

Wilson, M., and Monea, M. (2004), IEA GHG Weyburn CO2 monitoring and storage project. Summary

report 2000-2004: in Proceedings of the 7th

International Conference on Greenhouse Gas Control

Technologies (GHGT-7).

Wong-Parodi, G., Ray, I., and Farrell, A.E. (2008), Environmental non-government organisations'

perception of geologic sequestration: Environment Research Letters 3, 1-4.

Young-Lorenz, J.D. and Lumley, D. (2013), Portfolio analysis of carbon sequestration technologies and

barriers to adoption: General methodology and application to geological storage: Energy Procedia, 37,

5063-5079.

Zoback, M.D. and Gorelick, S.M. (2012), Earthquake triggering and large-scale geologic storage of

carbon dioxide: in Proceedings of the National Academy of Sciences (PNAS) 2012.

Algae CCS

124

Algae CCS

125

CHAPTER 4

APPLICATION OF METHODOLOGY:

ALGAE CCS

OVERVIEW

The new assessment methodology and rating system are applied to Algae CCS. Algae CCS

provides a source of renewable energy that temporarily captures and stores CO2 through

artificially accelerated photosynthetic processes until the biomass is converted to bioenergy or

biofuels. Algae CCS offers a range of possible revenue streams and environmental benefits, but

the time scale for commercial implementation is in the medium to long term due to significant

R&D requirements. A lack of published case studies and generalised life cycle assessments for

this technology required the extrapolation of some evaluation criteria from analogues such as

renewables and bioenergy. Following an initial review of the applicability of the methodology

and an introduction to Algae CCS, I apply my new methodology, then summarise and discuss

the results.

Algae CCS

126

Algae CCS

127

4.1 APPLICATION OF METHODOLOGY: ALGAE CCS

First, I check that all evaluation criteria of the methodology as derived from the calibration

with Geosequestration are applicable, and recalibrate and update the Science and Technology

Sub-Matrix to define the technology-specific elements of Algae CCS. The aim of the update is

to stay as close as possible to the original Sub-Matrix technology elements, calibrated with

Geosequestration, to allow fair comparison while capturing the unique elements of the specific

technologies.

I identify the technology elements for Algae CCS as:

1) Site identification and characterisation

2) Capture

3) Transportation

4) Cultivation

5) Conversion

The Cultivation element substitutes the fourth technology element of Geosequestration

(Injection and Storage). The growing and harvesting of sufficient algae biomass is an essential

technology element for the Algae CCS carbon mitigation option. Likewise, the fifth technology

element of Geosequestration, Measurement, Monitoring and Verification (MMV), has been

replaced by Conversion of the algae biomass into renewable biofuels or bioenergy. This last

element is not assessed in detail for specific products (bioenergy, biodiesel, ethanol, etc.) as

such an analysis is beyond the scope of this carbon capture and sequestration feasibility and risk

assessment.

4.2 INTRODUCTION

In their Stabilisation Wedge Model, Pacala and Socolow (2004) identify current technologies

which can be scaled up to achieve goals needed to reduce greenhouse gas (GHG) emissions.

One such technology is the decarbonisation of electricity and fuel through the substitution of

fossil fuels with biofuels. The Intergovernmental Panel on Climate Change [IPCC] Special

Reports (2011, 2012) on Renewable Energy Sources and Climate Change Mitigation (SRREN)

include the use of algal biomass as a potential bioenergy or biofuels option offering

opportunities for emissions reductions through GHG avoidance and CO2 sequestration.

Algae CCS

128

Algae produce complex organic compounds through photosynthesis or chemosynthesis

(inorganic chemical reactions), with the majority of algae having a photosynthetic mechanism

(US DOE, 2010). The mechanism of photosynthesis uses solar energy to convert CO2 and water

into carbohydrates and oxygen (O2). Algae produce 50% of the earth’s O2 but comprise only 1%

of the plant biomass, and are highly energy efficient. Algae reach a solar energy efficiency, the

ability to convert solar energy into chemical energy, of up to 5% (Fernández et al., 2012; Fon

Sing et al., 2013), compared to most plants at 1% solar energy efficiency (McKendry, 2001).

The photosynthetic process of converting inorganic carbon into organic compounds (lipids =

triglycerides and fatty acids, proteins and carbohydrates) is referred to as carbon fixation. The

lipids and carbohydrates are fuel precursors (biodiesel, jet fuel and gasoline, etc.) while the

proteins can be used for by-products like animal feed and fertiliser.

Algae can be divided into two classes, microalgae (e.g. single cell organisms) and

macroalgae (e.g. seaweed). The focus of my analysis will be on microalgae as they can achieve

a doubling of biomass per day (Sheehan et al., 1998), with a lipid yield for biodiesel production,

or a starch/polysaccharide yield for ethanol production up to ten times higher than for the best

performing terrestrial crops (US DOE, 2010). Of the ca. 30,000 known species of microalgae,

the global research focus has been on high oil producing species for fresh water which can

produce up to 70% of the biomass weight as lipids or vegetable oil, and saline water of ~10% of

biomass as lipids (Price, 2011). Fernández et al. (2012) state that typical oil content of algae

biomass is between 10% and 50% of weight, with the upper end of the range not currently

achievable in mass production.

Algae CCS incorporates the addition of anthropogenic CO2 to a microalgae system to

accelerate CO2 uptake and the growth cycle (Zeiler et al., 1995). Dependent upon the algae

strain and conditions of cultivation, ~1.8 to 2 tonnes (t) of CO2 will be utilised per tonne of

dehydrated algal biomass produced (Global CCS Institute, 2011). The biomass can be converted

into liquid biofuels and bioenergy, substituting fossil fuels and representing a temporary form of

storage, as well as by-products like chemicals and biochar, representing a more long term form

of storage. The combustion of algae biomass to produce energy or fuels releases the stored CO2,

but as the process recycles atmospheric CO2, it still represents a decrease in CO2 through

displacement of the burning of fossil fuels (Zeiler et al., 1995; Price, 2011). An overview on the

derivative products of microalgae biomass is given in Figure 4.1. Some discussion of the

derivative products is presented, but the main focus of this analysis is on the temporary storage

in the form of biofuel. Algae-based biodiesel or ethanol is considered a third-generation biofuel,

with first-generation biofuels based on agricultural commodities (e.g. wheat or corn), and

second-generation biofuels on processing of lignocellulosic biomass (e.g. straw or wood).

Algae CCS

129

Microalgae are cultivated in open (raceway ponds) and closed (photobioreactor) systems.

Open systems use paddle wheels to increase exposure to sunlight, prevent sedimentation and

mix nutrients in the raceway ponds (Figure 4.2). CO2 is introduced in the system from a buffer

tank. While having a lower OPEX and CAPEX, open systems have large areal requirements,

lower productivity and the risk of contamination (Price, 2011). In closed systems, the CO2 is

introduced along a series of transparent pipes and tubes, achieving a higher exposure of the

algae to the CO2 (Figure 4.3). The higher OPEX and CAPEX of closed systems is offset by the

lowered risk of environmental contamination and increased productivity. Combination systems

try to combine the advantages of both systems, using an initial closed incubator, followed by the

release into an open system. A range of techniques and combinations thereof are available for

the harvesting of microalgae; flocculation refers to the binding of algae through addition of

chemicals, sedimentation to the removal of algae strains from the base of settlement tanks,

filtration to the removal by use of filters, centrifugation to a centrifugal separation, and foam

fractionation to the introduction of sparged air to obtain concentrated foam (Price, 2011; Molina

Grima et al., 2003; Brennan and Owende, 2010). Sparging is a technique that bubbles a

chemically inert gas, such as air, or pseudo-inert gas such as CO2, through the algae growth

medium/liquid to separate the heavier concentration of algae slurry (Price, 2011; Scott et al.,

2010).

Figure 4.1. Algal Biofuel and Bioenergy Pathways (adapted from NNFCC, 2012).

Algae CCS

130

Figure 4.2. Example of an open raceway pond in Israel for the cultivation of microalgae for biofuel

(Image source: Seambiotic, 2013).

Figure 4.3. Example of a closed column tubular photobioreactor for cultivation of microalgae in China

(Image source: Energy Algae Group, 2013).

Algae CCS

131

The concentrated algae biomass can be used for a range of applications including (Figure 4.1):

1) Generation of bioethanol through a fermentation process,

2) Aid in wastewater treatment by removing nutrients with biofilms, encouraging the

anaerobic digestion of microbes, and preventing eutrophication (NNFCC, 2012).

Eutrophication describes the oxygen depletion of water as a consequence of excess

algae growth and decay because of an overabundance of nutrients such as

phosphates and nitrates.

3) Providing feedstock for biodiesel production through extraction of the oil in

physical and chemical processes or the addition of chemicals, and

4) Useful by-products (nutriceuticals such as carotenoids and omega oils,

pharmaceuticals, health food, animal feed, chemicals, fertiliser, and biochar).

The commercial cultivation and processing of high-value by-products, such as nutraceuticals

and health food, is already in operation. There are currently no Algae CCS projects to sequester

CO2 at commercial scales or for biofuel or bioenergy production (Campbell et al., 2011).

However, the Green Crude Farm facility is under development in Columbus, New Mexico,

USA by Sapphire Energy, the US Department of Energy [US DOE] and the US Department of

Agriculture. The first phase of 100 acres was completed in August 2012 (Sapphire Energy,

2012). By 2018, the integrated algae biorefinery will be the world’s first commercial-scale algae

to energy production facility, consuming ~56Mt CO2 per day on ~300 cultivated acres using

non-potable groundwater resources to produce ~1.5 million gallons pa of crude oil for jet fuel

and biodiesel (Sapphire Energy, 2013; Lane, 2012). The CO2 will either be sourced by tapping

the national pipeline network (Cortez pipeline) or from co-location with an industrial CO2

emitter (Lane, 2012). Research into algae farming, or “algaculture,” and the use of algae for

bioenergy and biofuels has evolved since the 1970’s (Sheehan et al., 1998). China and Chile are

the leaders in algae cultivation and research activities are proceeding in Australia, North

America, Norway and the UK. The huge potential of algae as a renewable source of carbon-

neutral liquid biofuel and bioenergy is recognised by governments, academia and industry, e.g.

in 2009, Exxon Mobil and BP invested $600 million and $10 million respectively into research

about the economic sustainability of algae-based biofuels and energy production (NNFCC,

2012).

Algae CCS

132

Algae CCS

133

4.3 RESULTS

Public Acceptance

How does Algae CCS affect me?

The CCS benefits of renewable biofuels are used in the assessment of Algae CCS as I have

not located literature regarding Public Acceptance specifically for this method. The benefits of

renewable energy sources are generally accepted by the majority of the wider public, but there is

some not in my backyard (NIMBY) attitude because the required infrastructure leads to a

perceived loss in property values due to noise pollution and changes to the visual landscape

(Upreti and van der Horst, 2004; Savvanidou et al., 2010; van der Horst, 2007; European

Commission, Directorate General for Research [EU DGR], 2003). As a third generation biofuel,

microalgae do not compete for agricultural land, unlike a food crop-based biomass feedstock.

The use of marginal land, unsuitable for agriculture, such as remote inland deserts with access

to saline groundwater, or located next to emission sources, will limit the exposure of the public

to the associated infrastructure. Nevertheless, the scale of commercial-scale production could

result in some conflicts, with estimates to cover ~10% of Australia’s oil demand suggesting 650

hectares (ha) of ponds and 4 giga-litre per annum (pa) of water as a requirement (Fon Sing et

al., 2013). This lowers the how does it affect me sub-criterion to a neutral (2) score, despite the

reduced competition for agricultural land.

How natural is Algae CCS?

Algae CCS is an acceleration of a natural photosynthetic process, and its by-products

(bioenergy and biofuel) are considered renewable energy, and it therefore receives the

associated positive public attitude towards “green” energy (Itaoka et al., 2012; Savvinidou,

2010; Ek, 2005: EU DGR, 2003). This yields a score of 3 for the how natural is it sub-criterion.

Science and technology comprehension

The positive rating is further enhanced as the public has a basic understanding of the

scientific concepts behind photosynthesis (Itaoka et al., 2012) through basic science education,

and its presence in the context of reforestation and “save the rainforest” campaigns. This earns a

science and technology comprehension score of 3.

An overall rating of 2.7 for Public Acceptance suggests that a “public positive” attitude prevails

towards Algae CCS.

Algae CCS

134

Regulatory Frameworks

Enabling body of rules and regulations

Algae CCS is regulated under renewable biomass legislation in the EU and USA that

encourages the use of renewable energy sources. A number of incentives are offered, such as the

EU Renewable Energy Strategy (RES) Directive, mandating each member state to obtain 20%

of its energy from renewable sources by 2020. The United States’ Department of Energy [US

DOE] announced under the American Recovery and Reinvestment Act (ARRA) that funds were

available for algae and advanced biofuel research ($85 million in 2009) and for biofuel research

and fuel infrastructure development ($78 million in 2010) (US DOE, 2010). Therefore the

enabling regulation sub-criterion scores a 3.

Restrictive/Negative legislation?

There is no direct restrictive/negative legislation specific to Algae CCS. However,

restrictive/negative legislation does occur in the indirect form of restrictions on CO2 onshore

transport infrastructure, namely pipelines, as in the UK (WA Energy Research Alliance [WA

ERA], 2010). This affects Algae CCS projects due to the continuous and large volume supply of

CO2 which is required, especially if remote sites for the cultivation infrastructure are

considered. This leads to a neutral (2) score for restrictive/negative legislation.

Regulatory certainty?

By making it a cost of doing business, regulatory compliance can significantly speed

up the adoption process. The mandated use of biodiesel in some US states, the UK

Government proposal to introduce Feed-in Tariff with Contracts for Difference (CfD)

and the German Renewable Energy Sources Act, mandating the addition of biofuel to

gasoline, and electricity feed-in tariffs (EEG, 2012) are all such examples. Because of

the certainty for economic planning which these regulations provide, regulatory

certainty scores a 3.

The overall rating of Regulatory Frameworks is a “supportive” 2.6.

Algae CCS

135

Economics

Viable business case?

The CAPEX and OPEX for Algae CCS investments is significant (US DOE, 2010; Fon Sing

et al., 2013), particularly for cultivation and conversion systems. The US DOE’s National Algal

Biofuels Roadmap (2010) does state however that a combination of improved productivity and

fully integrated production systems could lower the price of biofuels to compete with petroleum

at ~US$100 per barrel. The absence of commercial scale projects makes it difficult to

financially model and estimate the required investment level. Algae CCS is specifically omitted

in a number of governmental reports on renewable energy assessments, while the available

economics literature estimates have large uncertainties and vary widely in both technical and

economic assumptions (US DOE, 2010). Amongst others, the conversion rate of algae biomass

to oil is highly variable depending on climatic conditions and algae species, resulting in a highly

variable income stream (Campbell et al., 2009). Although Algae CCS has the potential to bring

in multiple income-streams, such as small volume – high value commodity chemicals (e.g.

omega oils), the focus in this analysis is on biomass to liquid biofuels and bioenergy. The price

of crude oil determines the production of microalgae at commercial scales (Borowitzka et al.,

2012). The viable business case sub-criterion scores a 2 rating, with cost uncertainties

downgrading the score while the multiple revenue streams from co-production have a positive

impact.

Government funding/incentives?

While no commercial scale Algae CCS production exists, a comparison with other

renewables shows a high dependency on government financial incentives like R&D grants,

feed-in-tariffs and tax credits or exemptions (EEG, 2012; Department of Energy and Climate

Change [DECC], 2011), therefore government funding/incentives earns a score of 1.

Carbon market?

It is unclear under current CCS regulatory frameworks whether Algae CCS qualifies as a

carbon-credit generating activity because it represents a carbon absorption and re-use

technology rather than permanent sequestration. It is argued that carbon credits could be

generated from the displacement of fossil fuels through alternative fuels and energy (Campbell

et al., 2011), but specific carbon policies will need to be developed to establish regulatory

Algae CCS

136

certainty (US DOE, 2010). This, together with the low carbon price and immature carbon

market yield a rating of 1 for the carbon market sub-criterion.

The overall assessment for Economics is “unaffordable” with an averaged score of 1.3.

Storage Quality

Capacity?

Algae CCS has the potential to absorb large volumes of CO2 in both offshore and onshore

locations throughout the world. According to Fernández et al. (2012), the production of 1t of

algae biomass requires a minimum of 1.8 t CO2, indicating significant technical capacity if up-

scaling is performed. However, the large areal space and water requirements for cultivation

systems and logistical challenges of matching emissions source and sink result in a

differentiation between technical and economic storage capacity. Current algae biomass

producers supply markets with prices of 10 – 300 EUR/kg for specialty products, while a

significant up-scaling of production to utilise the technical capacity will require the entering of

bulk commodity markets as feedstock for bioenergy or biofuels, with an anticipated price of

0.01 – 0.50 EUR/kg (Fernández et al., 2012). The large technical storage volumes, although not

currently translating into economic capacity, still earn Algae CCS a capacity rating of 2.

Rate?

The CO2 absorption rate is site specific and dependent on a large range of environmental and

resource parameters (CO2 source, sunlight hours, humidity, climate, water source, pond area,

algae strain, land area nutrients). An example calculation by the Commonwealth Scientific and

Industrial Research Organisation [CSIRO] (Campbell et al., 2009) for a Algae farm in Australia

assumes an annual CO2 utilisation of 185t CO2 per hectare, suggesting that the rate of storage is

mainly a question of scalability to reach commercial scale capacity of >1Mtpa. At the same

time, they observed significant economical and space limitations for the sites in Australia which

they considered.

Suitable Australian locations of a single algae farm have been identified in [2] as being

Mackay, Rockhampton, Gladstone (all in Queensland), Dry Creek, Port Augusta, Whyalla

(all in South Australia), Port Hedland, Broome and Kununurra (all in Western Australia).

Algae CCS

137

None of the above sites have sufficient land nearby to build enough algae farms to utilise all

of the carbon dioxide outputs of the power stations in question without increasing the piping

distance to 50km or more.

CSIRO (Campbell et al., 2009)

Due to the large number of variables impacting the algae growth rate as well as the limitations

imposed from the land usage side, rate earns a score of 2.

Duration?

Storage of CO2 in algae is a temporary form of storage until the end product, such as energy

or fuels, is combusted, and thus not considered as sequestration (US DOE, 2010), although the

displacement of fossil fuels represents a decrease in GHG concentrations. A proposed scenario

life cycle assessment by Brune et al. (2009) compares a gas-fired power plant with microalgae

as feedstock for power generation, suggesting a GHG avoidance potential of 22.3% – 29.7%.

Some algae by-products like charcoal or biochar may qualify as more permanent storage, but

the principal utilisation of algae as a biofuel/energy feedstock earns the duration sub-criterion a

1.

Verification?

Due to the temporary nature of the storage in form of biofuels, verification is not applicable.

Verification of stored CO2 is only applicable to pyrolitic by-products like biochar or charcoal

and therefore earns a rating of 1.5.

Storage Quality scores a “medium” (1.6) rating, mainly based on the large potential volumes

and achievable growth rate and potential of Algae CCS to displace of fossil fuels as a carbon

neutral renewable energy resource.

Environmental Impact

Potential impact on eco- and water systems?

Algae CCS provides a range of environmental benefits ranging from the recycling of CO2

and production of O2, the decrease in reliance on fossil fuels by providing a renewable energy

source, to the manufacture of other useful products such as biochar. The cultivation of some

Algae CCS

138

species in marine environments can result in the release of halogenated hydrocarbons and

dimethyl sulphide, which can produce a climate cooling effect (NNFCC, 2012). Algae

cultivation has a high water volume requirement, but cultivation systems can use salty or

brackish water as well as fresh water. Algae cultivation can also be utilised for wastewater

treatment (Pittman et al., 2011). The digestion of macronutrients (nitrogen (N) and phosphorus

(P)), trace elements (copper) and the removal of noxious gases reduces the eutrophication of

water sources, thus increasing O2 levels in the water and decreasing negative effects on aquatic

and marine diversity (Mata et al., 2010).

The possibility of using a variety of water sources as alternatives to potable water,

significantly mitigates the pressure on local freshwater resources. Marine algae systems may

place demands on offshore nutrients and impact other marine organisms. Mitigation strategies to

manage water resources include the careful selection of sites and assessment of the baseline

ecosystem before establishing an offshore farm, followed by careful monitoring and

management. Containment procedures like biofilm systems exist to address the risk of escaping

algae from open systems contaminating eco- or water systems. The potential impact on eco- and

water systems sub-criterion scores very favourably at 3.

Environmental footprint?

Algae CCS cultivation infrastructure in the form of large land based ponds (Figure 4.4),

photobioreactors or open marine systems should ideally be sited in close proximity to CO2

emissions sources. Additional infrastructure for transportation, processing and overnight storage

adds to this environmental footprint (Global CCS Institute, 2011; Campbell et al., 2009), and

can interfere with commercial agriculture, aquaculture or public access. This results in a

negative environmental footprint score of 1.

Overall, Environmental Impact earns a “medium” rating of 2.

Algae CCS

139

Figure 4.4 Aerial view of Cyanotech microalgae farm and facilities in Keahole Point, Hawaii, USA which

produces high-value nutraceutical and food products, such as spirulina, from microalgae (Image source:

Cyanotech Corporation, 2013).

4.3.1 Science and Technology: Sub-Matrix Assessment

Technology Element 1: Site Identification and Characterisation

Site identification and characterisation for Algae CCS determines the economic and

technical feasibility of onshore or offshore locations to provide sufficient carrying capacity to

cultivate algae biomass at sufficient volumes. Factors to be assessed include the nutrient

availability for offshore marine locations to avoid excessive stripping of nutrients from the

water column and impacting marine life, access to water sources (fresh, saline or waste),

location of CO2 emissions sources, and climactic conditions and their impact on growth rates

(NNFCC, 2012; Borowitzka et al., 2012).

The site identification and characterisation process involves little public interaction and

awareness of the activities undertaken to assess offshore or terrestrial sites, until the later stages

when the public should be engaged/consulted, e.g. public hearings for offshore cultivation and

fishing industry concerns. These factors balance out to a Public Acceptance neutral (2) score.

The absence of consistent regulation regarding Algae CCS site specifications and standards for

cultivation and conversion activities (how and where they should be set up, environmental

Algae CCS

140

precautions and mitigation of contamination, pollution, etc.), negatively impacts the regulatory

framework score, even though regulations from analogue technologies may provide guidelines

(US DOE, 2010). While the ability to work within existing regulations which cover various

aspects of Algae CCS is positive, the lack of a technology-specific and consistent framework

leads to uncertainty and is a risk factor, thereby earning a Regulatory Frameworks score of 2 .

The studies (e.g. water and nutrient availability, land topography, area and availability,

climate) to define the best sites are affordable. The need for large and flat areas can be met

through the usage of marginal areas with low alternative economic values, e.g. deserts, and salt

or brackish water sources can also be used (Campbell et al., 2011; Borowitzka and Moheimani,

2010; Borowitzka et al., 2012) (Figure 4.5). Economics scores an “affordable” (3) rating. The

evaluation tests are well established; although the absence of commercial scale Algae CCS

operations, and associated lack of experience about productivity, technical design or economics,

makes it difficult to define the specifications, especially long term, for cultivation sites

(Campbell et al., 2009). This lowers the Demonstrated Technical Readiness score to 2. The

studies to identify and characterise suitable locations are not invasive and leave no

environmental footprint. Ideally, a cultivation facility should be sited close to a source of CO2

emissions or pipeline networks designed to provide a regulated CO2 stream (NNFCC, 2012).

Environmental Impact scores a “high” (3) rating.

The average rating for Site Identification and Characterisation is just under the cut-off point

between medium and high (2.4).

Figure 4.5.

The Green Crude Farm

facility being developed in

Columbus, New Mexico,

USA by Sapphire Energy,

the US Department of

Energy [US DOE] and the

US Department of

Agriculture (Image source:

Sapphire Energy, 2013).

Algae CCS

141

Technology Element 2: Capture

Capture for Algae CCS refers to the process of separating and removing CO2, and scrubbing

of gases like hydrogen (H2) and sulphur oxides (SOx) from the air stream at large emitter sites

(power plants, fertilizer plants, etc). This element receives a favourable (3) Public Acceptance

rating as there is no perceived risk and the benefits to the environment are widely

acknowledged. With a few exceptions (e.g. UK and EU), the absence of legislated industry

compliance for the utilisation of capture technologies has a strongly negative impact and earns a

Regulatory Framework score of 1.

In general, Capture represents the most expensive part of a CCS project as the majority of

CO2 emission sources, such as coal fired power plants or other industrial emitters, have a CO2

content of less than 15% in the flue gas (IPCC, 2005). As typical gas stream properties of

natural gas include CO2 concentrations ranging from 2–65% (IPCC, 2005), the cost of

separating the CO2 before conversion to liquid natural gas (LNG) is an inherent cost of normal

production operations for LNG operators, so additional CAPEX for capture is limited. To

address low concentrations of CO2, complex and expensive technical solutions for separation

have to be employed. This earns Economics an unfavourable rating (1). On the technical side,

mature technology solutions for the capture process exist, such as post-combustion capture. This

solution captures approximately 90% of CO2 from industrial operations with a system

dependent energy penalty between 16% and 30% (SBC Energy Institute, 2012). A

Demonstrated Technical Readiness score of 3 indicates capture is ready and/or available in the

short term. The prevention of CO2 release and the availability of mitigation strategies to manage

negative by-products like amine solvents (Shao and Stangeland, 2009) yield an Environmental

Impact score of 3.

Capture scores an overall score of 2.2, as unfavourable economics and an absence of legislated

compliance negatively impact the score.

Technology Element 3: Transportation

Transportation for Algae CCS encompasses the safe and secure transport of the CO2 from

emissions points to the cultivation facility, as well as transport of the biomass for processing

and distribution of the end products. The impact of CO2 pipeline infrastructure and increased

vehicle traffic to transport biomass and end products in densely populated or environmentally

sensitive areas could negatively affect public opinion. Appropriate transport solutions, e.g.

Algae CCS

142

trains, may mitigate some of the impact if feasible. Public Acceptance scores a neutral (2)

rating. The transport of biomass and other by-products is already covered by existing pipeline

and haulage legislation. The same is valid for the road and rail transportation of CO2, however

conveyance of CO2 by pipeline may require amendments to existing pipeline legislation. This

earns the Regulatory Framework a score of 3.

Pipelines represent the most effective and economic onshore transport solution for the CO2

(SBC Energy Institute, 2012) and existing road and rail infrastructure can be utilised for

biomass and end products. However, it is unknown if existing pipeline and tanker designs and

algal lipid intermediates are compatible (US DOE, 2010). This balances out in a neutral (2)

rating for Economics, as the potential upgrade of components of the infrastructure may be

necessary. Pipeline and truck technology for the transport of biofuels and volatile gases is

already mature (IPCC, 2005). Compatibility of existing technologies with the stability,

prevention of auto-oxidation, temperature, and materials needs of algal bio-crude transportation

are unknown. Currently, they are assumed to be similar to plant lipids (US DOE, 2010). This

earns Demonstrated Technical Readiness a neutral (2) score. There are logistical challenges

because of the distances involved between widely distributed emissions sources, cultivation

ponds and conversion facilities. Some environmental footprint exists from pipeline networks,

possible increased road and rail infrastructure, and higher utilisation of the infrastructure. The

latter is especially critical if the preferred transport solution is truck haulage for biomass to

conversion facility and distribution of end products. Therefore Environmental Impact rates a

“high negative impact” score (1).

The overall score for the Transportation element is a medium (2) rating.

Technology Element 4: Cultivation

Cultivation refers to the growth, production and harvesting of microalgae at commercial

scales using either open raceway ponds, closed systems (series of transparent pipes, tubes and

hoses) or a combination of both with the first being the most commonly used since the 1960’s.

Most products require the removal of water and extraction of oil. Worldwide research efforts

and pilot studies in Australia, the UK, USA (California), EU (Germany, Netherlands) and

Norway have focused on establishing the best strains of microalgae for high lipid productivity,

high density, rapid growth rates, harvestability and survivability (Hannon et al., 2010; Hu et al.,

2008; Borowitzka et al., 2012).

Algae CCS

143

One challenge is to produce sufficient biomass feedstock that is sustainable and provides

reliable productivity across seasons and in different types of environments. Research by Dr. Jian

Qin from Flinders University, Queensland, Australia found that for Botryococcus braunii, the

maximum biomass for optimum hydrocarbon production is achieved with 23°C, 30-60 W/m2

irradiance light intensity, a 12hr light – 12hr dark photo period and a salinity of 8.8%, i.e.

confirming that this strain is tolerant of brackish water (Davidson, 2005). High concentrated

CO2 must be supplied to the microalgae cultivation systems, with some challenges such as the

rise in pH and salinity of the algal culture, the large volumes of gas that must be processed, and

vast quantities of water are required to keep the CO2 at low solubility. Algae consume CO2

during the daytime and they respire and release CO2 in the night, therefore storage facilities are

needed for CO2 piped from emissions sources such as power plants. Technologies from existing

algae applications can be applied to the algae biofuels production process, for example,

ultrasonic cavitation for harvesting from the nutraceuticals industry, which uses ultrasound

waves to break down algae cell walls (Cravotto et al., 2008).

Renewable sources of energy are favoured by the public and algae can absorb atmospheric

and anthropogenic CO2 during the photosynthetic process. Some NIMBY attitude prevails with

respect to the large land and water requirements, although siting in remote locations can

mitigate this. The same balance can be found with competition for land where the large

environmental footprint for ponds or aquatic farms restricts land usage (e.g. agricultural) or

marine access. Balancing these factors, Public Acceptance earns a “neutral” (2) rating. The

cultivation and conversion of algae biomass to biofuel, bioenergy or specialty products are

regulated and enabled under existing agricultural and renewables legislation so Regulatory

Frameworks earns a “supportive” (3) rating.

The high CAPEX and OPEX (dewatering and separation of products, high energy needs) of

cultivation systems and the difficulties with scaling up integrated extraction systems to

commercial levels affects the economics very unfavourably at the current stage (Fon Sing et al.,

2013). As salt or brackish water sources can be used, the reliance on expensive treated or fresh

water can be minimised, and has the additional benefit of minimising bacteria in the algae

strains. But overall the Economics of Cultivation scores an “unaffordable” (1) rating. The

technical readiness of cultivation systems has not been demonstrated at industrial scale.

Increased productivity research is estimated to take between three to fifteen years to identify the

best algae strains and cultivation systems to ensure sufficient volumes of biomass (US DOE,

2010; Global CCS Institute, 2011; NNFCC, 2012). Demonstrated Technical Readiness scores a

“long-term” (1). Algae CCS benefits the environment by recycling anthropogenic CO2 and

Algae CCS

144

providing a renewable energy source for liquid fuels and power generation, but is offset by the

environmental footprint of cultivation infrastructure. The utilisation of available water

resources, even if not immediately of use (brackish water), still represents a significant

intervention in an ecosystem. A regulated CO2 stream, and the overnight storage facilities for

CO2 because of fluctuating production rates, adds to the infrastructure footprint. These factors

balance out to an Environmental Impact score of 2.

The overall assessment of Cultivation is “mid - long term” readiness (1.8).

Technology Element 5: Conversion

Conversion refers to the various industrial processes used to convert or refine the microalgae

biomass into liquid biofuels, bioenergy or other by-products. The processes involved depend on

the algae strain and desired end product, and include: extraction (pressing and separation),

dehydration, using mechanical (expression, ultrasonic, membranes), chemical, thermochemical

or biochemical means, or a combination of methods (Molina Grima et al., 2003). The

biotechnical process is similar to that used in conventional biodiesel production, i.e.

transesterification (Demirbas and Demirbas, 2011). More advanced conversion techniques such

as direct gasification and biomass to liquid have a longer lead time for R&D (European Biofuels

Technology Platform/Zero Emission Fossil Fuel Power Plants [EBTP/ZEP], 2012; SBC Energy

Institute, 2012; Jonker and Faaij, 2013). Figure 4.6 illustrates the biorefinery research facility at

Wageningen University and Research Centre, Netherlands.

Public Acceptance is based on the analogy of fuels and energy generated from renewable

biomass sources being preferred to refinery/power plants which combust fossil fuels (Itaoka et

al., 2012; EU DGR, 2003). However, concerns about infrastructure siting (Upreti and van der

Horst, 2004; Savvanidou et al., 2010; van der Horst, 2007) lower the Public Acceptance score

to 2. Regulatory Frameworks favour the use of renewables and incentivise them, e.g. supporting

the CAPEX and OPEX of conversion plants, and through feed-in tariffs (EEG, 2012; DECC,

2011). Existing regulation for refineries and pollution controls can be applied or modified,

therefore Regulatory Frameworks earns a favourable (3) rating.

The economics of conversion to biofuel depend on the end product, e.g. biodiesel, and are

defined by the bulk commodity character of high volume but low margin products. Some

conversion methods like thermal have a high CAPEX (NNFCC, 2012; US DOE, 2010). Studies

differ in their estimates of algal oil production costs and range from $8.52 per gallon for open

ponds to $18.10 for photobioreactors (Davis et al., 2011; Sun et al., 2011). The biomass could

Algae CCS

145

however be used to generate power through anaerobic digestion if an electricity plant has

biomass capabilities. All these factors, together with the high OPEX related to their high energy

requirements yield an “unaffordable” Economics score (1). Current conversion technologies

are proving difficult to scale up and lack commercial scale demonstration (SBC Energy

Institute, 2012; Davis et al., 2011), thus Demonstrated Technical Readiness rates a low score

(1). The end products of algae biomass conversion to biofuel or bioenergy are carbon

neutral/negative, but the infrastructure footprint and the energy intensity of current conversion

process technologies lower the Environmental Impact score to a medium (2).

The overall rating for Conversion is “mid-long” term (1.8).

Figure 4.6 Horizontal tubular photobioreactor microalgae cultivation system and biorefinery at

AlgaePARC Research Facility, Wageningen University and Research Centre, Netherlands (Image source:

AlgaePARC, Wageningen UR, 2013).

Summary of Science and Technology Sub-Matrix Assessment

With the exception of Site Identification and Characterisation, the technology elements

specific to Algae CCS score around the medium range (Table 4.1). The evaluation criteria of

Economics and Demonstrated Technical Readiness have the lowest scores. This is because of

the significant amount of further R&D needed (algae strains, cultivation systems and conversion

Algae CCS

146

technologies) to maximise production efficiencies and reliability for sustainable volumes, and to

lower the costs of Capture, Cultivation and Conversion (US DOE, 2010).

The Science and Technology Sub-Matrix receives a “mid-long term” (2.04) assessment rating

(Table 4.1) which is rounded down to 2 and entered into the Science and Technology column of

the Master Matrix (Table 4.2).

Table 4.1. Science and Technology Sub-Matrix Scorecard: Results of Application of Methodology to

Algae CCS.

4.3.2 Summary of Results: Overall Technology Rating

The overall assessment rating for Algae CCS at 2.0 (Table 4.2) suggests a medium

probability of success as a commercial-scale CO2 sequestration option. Certain evaluation

criteria assessed in this analysis of Algae CCS do not have public domain data available that

address this specific carbon mitigation technology. The lack of case studies and data about pilot,

demonstration, commercial or operational projects, means that some of the ratings are based on

my analysis which is an extrapolation of literature from closely related subjects, such as

renewables and bioenergy.

Table 4.2. Master Matrix Scorecard: Results of Application of Methodology to Algae CCS.

Algae CCS

147

4.4 DISCUSSION

My assessment of Algae CCS yielded a number of observations which should be taken into

account when planning a carbon management portfolio or project. My initial expectation of an

unfavourable (1) Science and Technology rating for Algae CCS due to the absence of large

scale demonstration projects, and the significant R&D efforts still required to identify the most

suitable microalgae strains and integrated conversion processes for high volume production of

biofuel, energy and biomass, was not upheld (IPCC, 2011; Global CCS Institute, 2011;

Borowitzka and Moheimani, 2010). The main contributing factors to a more favourable rating

are the generally high levels of public and political acceptance of renewable energy, and the

ability to refer to existing supportive regulatory frameworks and associated incentives for nearly

all of the sub-criteria. The lack of literature, data, case studies and demonstration projects means

some of my analysis is limited to extrapolation from literature from the related topics of

bioenergy and renewables.

Suggestions to Improve Uptake

I suggest the following as areas of priority to improve and accelerate the wide-spread

commercial deployment of Algae CCS:

An initial focus on demonstrating the combination of Algae CCS with municipal

waste water treatment as the first phase/application where CO2 from power plant

flue gas is fed into cultivation ponds using municipal waste water as both water and

fertiliser source. The microalgae utilise the CO2, nitrogen and phosphorus from the

nutrient-rich waste water to increase production of biomass and aid in anaerobic

digestion of microbes to form biogas (methane), both of which can be used to

generate electricity (Mitchell, 2008; Mata et al., 2010). A side benefit is the

mitigation of waterway eutrophication. A large-scale pilot project demonstrating this

combination of techniques would provide valuable learning and experience. It is a

simpler and faster means of up-scaling production which is solely reliant on

maximisation of biomass feedstock production, rather than having to wait for the

projected three to fifteen years worth of R&D activities to optimise biofuel lipids

and genetics of microalgae strains. (Pittman et al., 2011)

The development of biorefineries similar to traditional petroleum refineries. These

integrated biomass conversion facilities would co-produce a range of sustainable

fuels, power, heat and value added chemicals (low volume – high value products,

Algae CCS

148

e.g. beta-carotene), for multiple revenue streams, while maximising the utilisation

and improving the economics of biomass conversion processes (Brennan and

Owende, 2010; International Energy Agency Bioenergy [IEA Bioenergy], 2009; US

DOE, 2010).

An integrated technology that offers potential efficiencies is the engineered capture

of CO2 by photosynthesis, or algae (biologic) fixation. This process aims to convert

the algae directly into liquid fuels by feeding it a pure stream of CO2, i.e. CO2

capture with biofuel production. Investment in this biofuel technology could

improve efficiencies, but it is an early stage technology (demonstration phase) with

significant challenges to achieve technical maturity as the process is proving difficult

to industrialise and upscale, with logistical challenges, high biomass production

costs and heavy energy penalty (SBC Energy Institute, 2012; Davis et al., 2011;

Jonker and Faaij, 2013; Borowitzka et al., 2012).

4.5 CONCLUSIONS

Algae CCS is a third generation biofuel technology that not only lowers atmospheric and

anthropogenic levels of CO2, but also has the potential to provide renewable liquid biofuels and

bioenergy. As such, it has high levels of Public Acceptance and supportive Regulatory

Frameworks because it is a carbon neutral/negative renewable energy source. The most

important evaluation criterion affecting the commercial adoption of Algae CCS as a carbon

mitigation option is Science and Technology, which in turn affects Economics. The low (1.6)

score for Storage Quality, because of its temporary duration rather than permanent

sequestration, affects the Economics carbon market sub-criterion, but is offset by the ability to

offset GHG as a renewable energy for both liquid fuel and power generation.

Algae CCS represents significant potential, with multiple income streams and a number of

associated environmental benefits. At this point, all the commercialisation pathways (biofuels,

bioenergy and commodity chemicals) still need to prove their potential for sustainability, up-

scalability and industrialisation (SBC Energy Institute, 2012; Davis et al., 2011; Jonker and

Faaij, 2013; Fon Sing et al., 2013). Some technology elements have been demonstrated but are

not yet commercial because of financial viability or up-scaling issues. R&D efforts focus on the

identification of suitable microalgae strains, improvements in productivity, yield and the

development of integrated cultivation systems and energy-efficient conversion techniques, with

Algae CCS

149

reports estimating technical and commercial maturity timeframes between a wide ranging three

to fifteen years (SBC Energy Institute, 2012; US DOE, 2010; Global CCS Institute, 2011;

NNFCC, 2012). In time, Algae CCS promises to be an effective carbon mitigation solution

because of: fast microalgae growth rates, high yields of lipids and biomass for renewable liquid

biofuels and bioenergy, global applicability in both terrestrial and marine environments, little/no

competition with agricultural land, high quality and flexibility of by-products, and fast uptake of

and efficiency in utilising CO2.

Algae CCS

150

Algae CCS

151

REFERENCES

Borowitzka, M.A., and Moheimani, N.R. (2010), Sustainable Biofuels from Algae: Mitigation and

Adaptation Strategies for Global Change, 1-13.

Borowitzka, M.A., Boruff, B. J., Moheimani, N. R., Pauli, N., Cao, Y., and Smith, H. (2012),

Identification of the optimum sites for industrial-scale microalgae biofuelpProduction in WA using a

GIS model: Final Report prepared for The Centre for Research into Energy for Sustainable Transport

(CREST).

Brennan, L., and Owende, P. (2010), Biofuels from microalgae—a review of technologies for

production, processing, and extractions of biofuels and co-products: Renewable and Sustainable

Energy Reviews, 14, 2, 557-577.

Brune, D.E., Lundquist, T.J., and Benemann, J.R. (2009), Microalgal Biomass for Greenhouse Gas

Reductions; Potential for Replacement of Fossil Fuels and Animal Feeds; Journal of Environmental

Engineering, 135, 11, 1136-1144.

Campbell, P.K., Beer, T. and Batten, D. (2009). Greenhouse gas sequestration by algae – energy and

greenhouse gas life cycle studies: Presented at the 6th Australian Conference on Life Cycle

Assessment, Australian Life Cycle Assessment Society (ALCAS).

Campbell, P.K., Beer, T. and Batten, D. (2011), Life cycle assessment of biodiesel production from

microalgae in ponds: Bioresource Technology, 102, 2, 50-56.

Cravotto, G., Boffa, L., Mantegna, S., Perego, P., Avogadro, M., and Cintas, P. (2008), Improved

extraction of vegetable oils under high-intensity ultrasound and/or microwaves: Ultrasonics

Sonochemistry, 15, 5, 898-902.

Davidson, S. (2005), Growing fuel from algae: Ecos, 124, 34-34.

http://www.ecosmagazine.com/?act=view_file&file_id=EC124p34.pdf (accessed 10 December 2012 )

Davis, R., Aden, A., and Pienkos, P.T. (2011), Techno-economic analysis of autotrophic microalgae for

fuel production: Applied Energy, 88, 10, 3524–3531.

Demirbas, A., and Demirbas, M.F. (2011), Importance of algae oil as a source of biodiesel: Energy

Conversion and Management, 52, 1, 163-170.

Department of Energy and Climate Change [DECC] (2011), Planning our electric future: a White

Paper for secure, affordable and low-carbon electricity: Presented to Parliament by the Secretary of

Algae CCS

152

State for Energy and Climate Change by Command of Her Majesty, July 2011.

https://www.gov.uk/government/uploads/system/uploads/attachment_data/file/48129/2176-emr-

white-paper.pdf (accessed 28 July 2012).

EEG (2012), Gesetz für den Vorrang Erneuerbarer Energien, 25 October 2008 (BGBl. I S. 2074),

amended on 17 August 2012 (BGBl. I S. 1754).

Ek, K. (2005), Public and private attitudes towards ‘‘green’’ electricity: the case of Swedish wind power:

Energy Policy, 33, 1677–1689.

European Biofuels Technology Platform/Zero Emission Fossil Fuel Power Plants [EBTP/ZEP]

(2012), Biomass with CO2 capture and storage (Bio-CCS): The way forward for Europe: The

Advisory Council of the European Technology Platform for Zero Emission Fossil Fuel Power Plants

(ZEP) and the Steering Committee of the European Biofuels Technology Platform (EBTP).

http://www.biofuelstp.eu/bio-ccs.html (accessed 2 July 2012).

European Commission. Directorate General for Research [EU DGR] (2003), Energy: issues, options

and technologies, science and society, 20624 [European Opinion Research Group (Ed.)]: European

Commission.

European Parliament and the Council of the European Union [EP CEU] (2009). Directive

2009/28/EC of the European Parliament and of the Council, on the Promotion of the Use of Energy

From Renewable Sources and Amending and Subsequently Repealing Directives, 2001/77/EC and

2003/30/EC, 23 April 2009.

Fernández, F.G.A., González-López, C.V., Sevilla, J.F., and Grima, E.M. (2012), Conversion of CO2

into biomass by microalgae: how realistic a contribution may it be to significant CO2 removal?:

Applied microbiology and biotechnology, 96, 3, 577-586.

Fon Sing, S., Isdepsky, A., Borowitzka, M.A., and Moheimani, N.R. (2013), Production of biofuels from

microalgae: Mitigation and Adaptation Strategies for Global Change, 18, 1, 47–72.

Global CCS Institute (2011), Accelerating the uptake of CCS: Industrial use of captured carbon dioxide,

[Parsons Brinckerhoff], Canberra, Australia.

Hannon, M., Gimpel, J., Tran, M., Rasala, B. and Mayfield, S. (2010), Biofuels from algae: challenges

and potential: Biofuels, 1, 5, 763–784.

Algae CCS

153

Hu, Q., Sommerfeld, M., Jarvis, E., Ghirardi, M., Posewitz, M., Seibert, M., and Darzins, A. (2008),

Microalgal triacylglycerols as feedstocks for biofuel production: perspectives and advances: The Plant

Journal, 54, 4, 621-639.

International Energy Agency Bioenergy [IEA Bioenergy] (2009), IEA Bioenergy Task 42 Biorefinery.

[de Jong, E., Langeveld, H. and van Ree, R. (eds.)]: Wageningen University and Research Centre, The

Netherlands.

Intergovernmental Panel on Climate Change [IPCC] (2005), Intergovernmental Panel on Climate

Change Special Report on Carbon Dioxide Capture and Storage. Prepared by Working Group III of

the Intergovernmental Panel on Climate Change [Metz, B.O., Davidson, H., de Coninck, C., Loos, M.

and Meyer, L.A. (eds.)]: Cambridge University Press, Cambridge, United Kingdom and New York,

NY, USA.

Intergovernmental Panel on Climate Change [IPCC] (2011), Summary for policymakers. In: IPCC

Special Report on Renewable Energy Sources and Climate Change Mitigation (SRREN) [Edenhofer,

O., Pichs‐Madruga, R., Sokona, Y., Seyboth, K., Matschoss, P., Kadner, S., Zwickel, T., Eickemeier,

P., Hansen, G., Schlömer, S., von Stechow, C. (eds)]: Cambridge University Press, Cambridge, United

Kingdom and New York, NY, USA.

Intergovernmental Panel on Climate Change [IPCC] (2012), Intergovernmental Panel on Climate

Change Special Report on Renewable Energy Sources and Climate Change Mitigation (SRREN) [O.

Edenhofer, R. Pichs‐Madruga, Y. Sokona, K. Seyboth, P. Matschoss, S. Kadner, T. Zwickel, P.

Eickemeier, G. Hansen, S. Schlömer, C. von Stechow (eds)], Cambridge University Press, New York,

NY, USA.

Itaoka, K., Saito, A., Paukovic, M., de Best-Waldhober, M., Dowd, A-M., Jeanneret, T., Ashworth, P.

and James, M. (2012), Understanding how individuals perceive carbon dioxide: Implications for

acceptance of carbon dioxide capture and storage: CSIRO Report EP 118160, Australia.

Jonker, J.G.G., and Faaij, A.P.C. (2013), Techno-economic assessment of micro-algae as feedstock for

renewable bio-energy production. Applied Energy, 102, 461–475.

Lane, J. (2012), Change the world: Sapphire Energy’s Green Crude Farm, illustrated: Biofuels Digest, 2

October 2012. http://www.biofuelsdigest.com/bdigest/2012/10/02/change-the-world-sapphire-

energys-green-crude-farm-illustrated/ (accessed 24 April 2013).

Mata, T.M., Martins, A.A., and Caetano, N.S. (2010), Microalgae for biodiesel production and other

applications: A review: Renewable and Sustainable Energy Reviews, 14, 1, 217-232.

Algae CCS

154

McKendry, P. (2002), Energy production from biomass (part 1): Overview of biomass: Bioresource

Technology, 83, 37-46.

Mitchell, B. G., (2008), UCSD Consortium to Develop Biofuels from Microalgae: A concept; progress

one project at a time: Presented at the Algal Biotechnology Seminar Series.

http://www.spg.ucsd.edu/algae/pdf/080129_SDCMA_biofuel_symp_bgm.pdf (accessed 5 December

2012)

Molina Grima, E., Belarbi, E.H., Acien Fernandez, F.G., Robles Medina, A., and Chisti, Y. (2003),

Recovery of microalgal biomass and metabolites: process options and economics: Biotechnology

Advances, 20, 7, 491-515.

NNFCC (2012), Research Needs in Ecosystem Services to Support Algal Biofuels,Bioenergy and

Commodity Chemicals Production in the UK: A project for the Algal Bioenergy Special Interest

Group [Smith, N. and Higson, A.]: NNFCC Project Number: 12-008.

Pacala, S., and Socolow, R. (2004), Stabilization Wedges: Solving the climate problem for the next 50

years with current technologies: Science, 305, 5686, 968-972.

Pittman, J. K., Dean, A. P., & Osundeko, O. (2011), The potential of sustainable algal biofuel production

using wastewater resources: Bioresource technology, 102,1, 17-25.

Price, A. (2011), Sequestration of carbon dioxide using microalgae: Investigating Climate Change:

Parsons Brinkley (PB) Network, 72, 22-24.

Sapphire Energy (2012), Sapphire Energy’s commercial demonstration algae-to-energy facility now

operational: Initial phase of world’s first green crude farm completed on time and on budget: [press

release], 27 August 2012: Sapphire Energy, Columbus, NM, USA.

Sapphire Energy (2013), Green Crude Farm. http://www.sapphireenergy.com/locations/green-crude-

farm.html (accessed 24 April 2013).

Savvanidou, E., Zervas, E., Tsagarakis, K.P. (2010), Public acceptance of biofuels: Energy Policy, 38,

3482-3488.

SBC Energy Institute (2012), Leading the Energy Transition: Bringing Carbon Capture and Storage to

Market, [Soupa, O., Lajoie, B., Long, A., and Alvarez, H.]: Schlumberger Business Consulting [SBC]

Energy Institute.

Scott, S.A., Davey, M.P., Dennis, J.S., Horst, I., Howe, C.J., Lea-Smith, D.J. and Smith, A.G. (2010),

Biodiesel from algae: challenges and prospects: Current Opinion in Biotechnology, 21, 3, 277-286.

Algae CCS

155

Shao, R. and Stangeland, A. (2009), Amines Used in CO2 Capture- Health and Environmental Impacts:

Bellona Report 2009: The Bellona Foundation, Oslo, Norway.

Sheehan, J., Dunahay, T., Benemann, J. and Roessler, P. (1998), A Look Back at the U.S. Department of

Energy’s Aquatic Species Program: Biodiesel from Algae, 328, Golden: National Renewable Energy

Laboratory.

Sun, A., Davis, R., Starbuck, M., Ben-Amotz, A., Pate, R. and Pienkos, P. T. (2011). Comparative cost

analysis of algal oil production for biofuels: Energy, 36, 8, 5169-5179.

Upreti, B. and van der Horst, D. (2004), National renewable energy policy and local opposition in the

UK; the failed development of a biomass electricity plant: Biomass and Bioenergy, 26, 60-69.

US Department of Energy [US DOE] (2010), National Algal Biofuels Roadmap: US Department of

Energy, Office of Energy Efficiency and Renewable Energy, Biomass Program.

van der Horst, D. (2007), NIMBYor not? Exploring the relevance of location and the politics of voiced

opinions in renewable energy siting controversies: Energy Policy, 35, 2705-2714.

WA Energy Research Alliance [WA ERA] (2010), Geosequestration R&D: A Report to Australian

National Low Emission Coal R&D Program, March 2010.

Zeiler, K.G., Heacox, D.A., Toon, S.T., Kadam, K.L., and Brown, L. M. (1995), The use of microalgae

for assimilation and utilization of carbon dioxide from fossil fuel-fired power plant flue gas: Energy

Conversion and Management, 36, 6, 707-712.

Biochar CCS

157

CHAPTER 5

APPLICATION OF METHODOLOGY:

BIOCHAR CCS

OVERVIEW

The new assessment methodology and rating system are applied to Biochar CCS. Biochar

CCS involves the capture of CO2 through photosynthesis and sequestration occurs when the

biomass is converted into solid carbon. After an initial review of the applicability of the

methodology, I provide an initial introduction to Biochar CCS. This is followed by the

application of the methodology. The results of the analysis will be discussed and some

recommendations to accelerate the uptake made. It proved necessary to extrapolate some

evaluation criteria from analogues such as renewables and bioenergy due to a lack of

published literature or case studies in the public domain. The overall evaluation of Biochar

CCS predicts a high probability of success due to its immediate availability at different scales,

as well as a large range of economic, environmental and agricultural benefits.

Biochar CCS

158

Biochar CCS

159

5.1 APPLICATION OF METHODOLOGY: BIOCHAR CCS

Firstly, I check that all evaluation criteria of the methodology as derived from the

calibration with Geosequestration are applicable, and recalibrate and update the Science and

Technology Sub-Matrix to define the technology-specific elements of Biochar CCS,

mimicking the technology elements of Geosequestration as closely as possible.

I identify the technology elements of Biochar CCS as:

1) Site identification and characterisation

2) Feedstock

3) Transportation

4) Conversion (pyrolysis/gasification) to carbon products

5) MMV

The second technology element of Feedstock replaces the Capture element of

Geosequestration, as it is at this stage that CO2 is naturally “captured” from the atmosphere

during photosynthesis. Similarly to Algae CCS, the fourth element is substituted with

Conversion, albeit using different types of engineering and processes from those used in

Algae CCS. Biochar CCS includes the application of biochar to soil, however the end

applications of biochar, carbon materials, carbon products or energy are not assessed in detail

as such an analysis would go beyond the scope of this CCS feasibility and risk assessment.

5.2 INTRODUCTION

The management of forest and agricultural soils, under the category of natural carbon

sinks, is included by Pacala and Socolow (2004) in their Stabilisation Wedge Model as a

current technology which can be scaled up to achieve the goals needed to reduce greenhouse

gas (GHG) emissions. The use of biochar for sequestration purposes or as a potential

bioenergy or biofuel option offers opportunities for both emissions reductions and CO2

sequestration, and is included in the Intergovernmental Panel on Climate Change [IPCC]

Special Reports (2011, 2012) on Renewable Energy Sources and Climate Change Mitigation

(SRREN). A range of authors have discussed Biochar CCS, or pyrolysis of biomass with the

biochar returned to the soil, as one method of lowering GHG levels (Lehman et al., 2006;

Glaser et al., 2001; Laird, 2008; Lal, 2011; Roberts et al., 2009). The IPCC’s Special Report

Biochar CCS

160

on Land Use, Land Use Change, and Forestry (2000) including Biochar CCS as a key

technology to reach CO2 atmospheric concentration targets.

Biochar is one of three carbon-rich materials produced by pyrolysis, the thermo-chemical

decomposition of biomass (organic materials) in the absence of (or limited) oxygen (Demirbas

and Arin, 2010); the other two carbon products are tars or oils, and syngas. The definition of

biochar for the purposes of this thesis, follows that of the International Biochar Initiative [IBI]

(IBI, 2012), and incorporates the terms black carbon, char, charcoal, agrichar, and activated

carbon. The 2010 study by Lehmann and his research team at Cornell University estimates a

potential offset of 12% GHG emissions (CO2, methane (CH4) and nitrous oxide (N2O)). This

equates to 1.8 Gtpa CO2-equivalent (CO2-e) net emissions, with biochar produced from

sustainable biomass, and yields a larger GHG offset benefit than combustion of the same

biomass for bioenergy at 10% GHG emissions (Woolf et al., 2010). In its Special Report of

2000, the IPCC states that over 80% of organic carbon in terrestrial ecosystems is held in soil

rather than biomass. Plant matter absorbs CO2 through photosynthesis, trapping it in solid

stable carbon state when pyrolised (Figure 5.1).

Figure 5.1. Example of biochar (Image source: Australia and New Zealand Biochar Researchers

Network, 2013).

When added to soil at 0.5 m to 1 m depth, the biochar becomes sequestered through

humification, a state of physical and chemical stability for organic matter without further

break down of the constituents. Biochar is highly recalcitrant, i.e. stable and not easily

Biochar CCS

161

released into soluble form, thus slowing the rate at which carbon fixed by photosynthesis is

returned to the atmosphere. Biochar materials have a large surface area which are highly

porous (Figure 5.2), thereby supporting biochemical (micro organisms = microbiota,

aromaticity) and physiochemical reactions, facilitating nutrient uptake through increased

water retention and cation exchange capacity (CEC) (Sombroek et al., 2003; Liang et al.,

2006).

Biochar also supports surface sorption to adsorb nutrients and agricultural chemicals

(phosphates, nitrates), reducing leaching to soil and ground water. Many biochars have a

neutral to alkaline pH value and can act as a liming agent to neutralise acidic soils, adsorbing

ammonia (NH3) and decreasing its volitisation (Fowles, 2007; Waters et al., 2011; Lehmann

et al., 2006). It significantly reduces the release of CH4 and N2O from soil (Rondon et al.,

2005). The microbiota catalyse processes to increase carbon to nitrogen ratios, reducing

nitrogen loss and NO2 emissions by preventing/limiting anaerobic production of NO2 and

increasing nitrogen availability/efficiency for plant growth (Glaser et al., 2001). As a

fertiliser, the effects of biochar on soils, crop yield and soil carbon are determined by the

biochar quality (nutrient composition), as well as the properties of soil and ecosystem where

the biochar is applied (Amonette and Joseph, 2009; Lehmann et al., 2006).

Figure 5.2. Electron microscopic image of biochar showing micropore structure (Image source:

Biochar Project (Jocelyn), 2013).

Biochar CCS

162

Biochar quality, such as its water retention ability, and physical and chemical properties

and stability, and the proportion of products produced by pyrolysis are dependent upon the

nature of the feedstock and pyrolysis conversion conditions of temperature, heating rate,

pressure, water content, and residence time (Table 5.1) (Amonette and Joseph, 2009;

Demirbas and Arin, 2002; International Energy Agency Bioenergy [IEA Bioenergy], 2007).

Suitable feedstocks include organic matter such as crops, or municipal and agricultural wastes

(Figures 5.3 and 5.4). The use of waste biomaterials such as: wood, leaves, food waste, straw

or manure, can help mitigate landfill and waste management issues, and avoid methane gas

(CH4) emissions from decomposition (Malkow, 2004). Sustainable biomass is organic waste

that does not affect food product, land-use change or soil conservation. Pyrolysis converts

approximately 50% of the biomass into biochar.

Pyrolysis processes are split into two basic classifications of fast and slow, with slow

pyrolysis advocated as best for the purposes of carbon sequestration as it maximises

production of biochar and syngas over bio-oil (Table 5.1)(IEA Bioenergy, 2007; Brown,

2009; Demirbas and Arin, 2002; Jahirul et al., 2012; Laird et al., 2009). The syngas and bio-

oil can be used to generate power to run the pyrolysis conversion process facilities. If the

major objective of the plant is syngas production for electricity generation, gasification plants

can be used, but will not be discussed in detail in this thesis as they are not optimised for

biochar production in the context of a CCS scheme (IEA Bioenergy, 2007; Laird et al., 2009;

Brown, 2009).

Table 5.1. Typical product yields (dry wood basis) obtained by different modes of wood pyrolysis

(adapted from International Energy Agency Bioenergy [IEA Bioenergy], 2007).

Mode

Conditions

Bio-Oil

Biochar

Syngas

Fast Moderate temperatures (500°C) for 1 second 75% 12% 13%

Intermediate Moderate temperatures (500°C) for 10–20

seconds

50% 20% 30%

Slow

(carbonisation)

Low temperature, (400°C), very long solids

residence time

30% 35% 35%

Gasification

High temperature, 800°C, long vapour

residency time

5% 10% 85%

Biochar CCS

163

Figure 5.3. Potential ecosystem benefits from biochar production and application systems (Figure

adapted from Waters et al., 2011).

Figure 5.4. Overview of the sustainable biochar concept (Image Source: Woolf et al., 2010).

Biochar CCS

164

Pyrolysis systems are already in operation around the world at a range of scales; from

household stoves and kilns to industrial facilities producing products for specialty markets,

such as bio-oil derived acetic acid, resins, food flavourings, agrichemicals, fertilisers, and

emission-control agents (URS Australia, 2010; Winsley, 2007). Pyrolysis can involve a wide

range of reactor types, e.g. fixed and circulating fluid beds, bubbling fluidised bed, rotating

cone, ablative, and mechanical or centrifugal ablative processes (Winsley, 2007; Demirbas

and Arin, 2002; Jahirul et al., 2012; IEA Bioenergy, 2007). Global research activities (e.g.

New Zealand, Australia, Japan, Scandanavia, Europe, United Kingdom, Northern Africa,

North America and Brazil) are focused on more affordable, environmentally-friendly and

production-efficient pyrolysis systems, as well as on the agricultural and environmental

benefits of biochar (Jahirul et al., 2012; Winsley, 2007).

To date, there are no commercially operating biochar production systems for the primary

purpose of carbon sequestration on which to base an assessment of possible combinations of

feedstocks, conversion technologies, and application systems (National Resources Defense

Council [NRDC], 2010). Since 2006, Pacific Pyrolysis has operated and scaled up a pilot

demonstration facility in Somersby, north of Sydney, New South Wales, Australia, and is

constructing two additional slow pyrolysis plants. The Ballina plant project in NSW, owned

and operated by the Ballina Shire Council, will operate from a waste management centre, and

has the capacity to divert 29,000 t of organic waste from landfills, remove ~50,000 tpa of

carbon, and generate 6000 MWh of carbon negative renewable energy. The Carbon Negative

Electricity Project in Melbourne, owned and operated by Pacific Pyrolysis, will process solid

organic waste at the rate of 2 t/hr to generate 1MW of electricity and up to 5,000 t pa of high

grade biochar (Pacific Pyrolysis, 2013; BEN 2011, 2012; Ballina Information, 2012).

Biochar provides a number of complementary environmental, economic and societal

benefits that can be simultaneously achieved, and covers the mitigation of climate change by

sequestering CO2 in solid carbon state, waste and land management, renewable energy

production, and soil and water improvement (Lehmann and Joseph, 2009a; Laird et al., 2009).

Figure 5.4 provides an overview of the sustainable biochar concept proposed by Woolf et al.

(2010) illustrating the inputs, process, outputs, applications and impacts on the global climate,

with the relative proportions of the components under each category. De Gryze et al. (2010)

provide examples from different slow pyrolysis technology studies, examining the GHG

emissions avoided with different organic inputs, for example, corn stover ranging from 1.1 –

10.7 t CO2-e per t of dry feedstock, and yard waste ranging from 3.0 – 12.5 t CO2-e.

Biochar CCS

165

The potential benefits of Biochar CCS to plant-soil systems are shown in Figure 5.5. The

dark earth Terra Preta de Indio soils of the Amazon region are characterised by high content

of pyrogenic carbon (similar to biochar) and demonstrate the ability to maintain high fertility,

retain nutrients (reduce leaching), promote liming (overcome soil acidity) and increased

fertiliser efficiency and increased plant yield. Biochar CCS can address the growing problem

of food security, by producing more crop yield on less land (Glaser et al., 2001; Neves et al.,

2003; Lal, 2004). Radiocarbon dating of the Terra Preta soils measured the mean residence

time (MRT) as ranging from 500-7000 years illustrating that this is a long term sequestration

option (Neves et al., 2003). Whitman et al. (2010) suggest that for the purposes of carbon

accounting, although the MRT of biochars is >1000 years, more effective accounting could be

achieved by using a recalicitrant fraction, which is highly stable, and a labile fraction, which

describes the portion which is more quickly decomposed. The MRT and soil benefits depend

on the type of feedstock, conversion method and soil where the biochar is applied, and length

of time the soil is left undisturbed to minimise decay and potential release of CO2 back into

the atmosphere.

Figure 5.5. Potential impacts of Biochar application to plant-soil systems (Figure adapted from Waters

et al., 2011)

Biochar CCS

166

Biochar CCS

167

5.3 RESULTS

Public Acceptance

How does Biochar CCS affect me?

Application of Biochar CCS could require land-use changes to grow biomass and cause

potential conflicts and concern about agricultural land use switch to produce crops for

biomass and/or using food crops for biofuels (Ipsos-MORI, 2010; Müller et al., 2008). On the

other hand, there are multiple benefits to communities through improved soil health, water

retention, agricultural yield, lower fertiliser requirement, recycling of waste biomass and

reduction of landfill requirements, etc. Offsetting these against each other means the how does

it affect me sub-criterion scores a 2.

How natural is Biochar CCS?

In the Ipsos-MORI Experiment Earth Report (2010), the perceived naturalness of a

technique was a powerful stimulus for producing support, with Biochar CCS being perceived

as a more natural approach and more feasible in comparison to stratospheric aerosols, solar

radiation management or ocean fertilisation techniques. Biochar CCS enhances the natural

carbon cycle that presents a low/no risk. The other by-products of pyrolysis are renewable

“green” energies, and have an associated positive public attitude (Savvinidou, 2010; Ek,

2005). This yields a score of 3 for the how natural is it sub-criterion.

Science and technology comprehension

The benefits to the environment are easily understood by the public which has a basic

understanding of the scientific concepts behind photosynthesis (Itaoka et al., 2012; Ipsos-

MORI, 2010). Biochar CCS sequesters the CO2 captured through natural photosynthetic

processes into a stable form before decay and release of the CO2 or CH4 into the atmosphere

(Brown, 2009; Harris, 1999). The general public is also familiar with the basic concepts about

the technology used to produce charcoal as it has been utilised globally for thousands of years

(Food and Agriculture Organisation [FAO], 1983; Moscowitz, 1978). This earns a science and

technology comprehension score of 3.

An overall rating of 2.7 suggests that a “public positive” attitude prevails towards Biochar

CCS and this is entered into the Public Acceptance column of the Master Matrix.

Biochar CCS

168

Regulatory Frameworks

Enabling body of law and regulations?

The activities to implement Biochar CCS are implicitly or directly covered under a diverse

range of existing global regulation from a myriad of regulators including the United Nations

[UN], local, state, federal and regional governments as well as authorities (Cowie et al., 2012;

Driver and Gaunt, 2010). This includes Article 3.3 of the UN Framework for Climate Change

[UNFCC] (UNFCC, 1992), the Kyoto Protocol (UNFCCC, 1998a), and legislation for

renewable biofuels, sustainable feedstocks, production and environmental compliance policy

for industrial processors, etc. (Read, 2009; Cowie et al., 2012; Downie et al., 2012). Therefore

the enabling regulation sub-criterion scores a 3.

Restrictive/negative regulation?

Restrictive regulations that apply to Biochar CCS relate to land-use change, and the

production, use and conversion of sustainable biomass feedstocks. Suitable locations include

remote or under-utilised areas, or those which can benefit from soil amendment while

continuing with the existing land-use. Sufficient stocks of sustainable biomass exist in the

form of municipal, agricultural and forestry waste. The restrictive/negative regulation that

does exist is not a large barrier to technology uptake, therefore the score for

restrictive/negative regulation balances out to a neutral (2) score.

Regulatory certainty?

Industrial standards for pollution limits (e.g. emissions from contaminated waste streams,

dioxins, heavy metal and particulate emissions, air quality standards) depend on local

jurisdictions. Some regulatory bodies aim to develop a formal position on biochar regulation

and sustainable practice guidelines, e.g. the Scottish Environment Protection Agency (SEPA)

is the first regulatory agency in the UK, to authorise the manufacture and use of biochar from

waste (untreated wood waste from agriculture, horticulture and forestry) (SEPA, 2010). The

Council of Standards Australia has published an updated Standard, or quality framework,

defining processing requirements and chemical, biological and physical properties of

composts, soil conditioners and mulches to protect environment and public health (Standards

Australia, 2012). The positive moves by various regulatory bodies to provide more certainty

Biochar CCS

169

combined with the current overlapping laws and regulation that cover all stages of the Biochar

CCS lifecycle, earn regulatory certainty a “supportive” (3) rating.

An overall “supportive” (2.7) rating is entered into the Master Matrix in the Regulatory

Frameworks column.

Economics

Viable business case?

The economics of Biochar CCS are feasible in many places and at different scales using

current production technology (Winsley, 2007; URS Australia, 2010; Measurable Offsets,

2010; WorldStove, 2009; Phoenix Energy, 2013). At the moment, only small Biochar CCS

projects are operational, such as the WorldStove project which supplies modern pyrolysis

household stoves and stove hubs to developing nations (Surridge, 2011; WorldStove, 2009),

and Phoenix Energy in the US, which converts biomass to energy with biochar as a retail by-

product (Phoenix Energy, 2013; IBI, 2012).

Some experience for up-scaling may be gained from operators of biomass to electricity

conversion plants. The up-scaling of biochar plants can either be performed through increased

volume throughput, larger facilities, or adaptation of the distributed small scale operator

network business model like that employed by Phoenix Energy. Phoenix Energy constructs

small (0.5MW to 2MW) on-site biomass gasification plants in partnership with local

agricultural, forestry and waste management, thereby reducing transport costs, and increasing

potential revenue streams by commercialising the excess heat, electricity and biochar

produced (~1t biochar produced per day from a 0.5MW plant) (Phoenix Energy, 2013; IBI,

2012). Pacific Pyrolysis offers a commercial small-scale design which is optimised for

biochar production instead of energy production (Ballina, 2012; BEN, 2011, 2012; IBI, 2012).

De Gryze et al. (2010) outline in their compilation of theoretical economics analyses that one

of the challenges of larger plants may be the anticipated negative net return, partially

attributed to the fact that the biomass has to be sourced from a larger area, while illustrating

that existing net return calculations for smaller scale slow pyrolysis units, based on

agricultural waste, should yield positive net returns. At the same time, it is cautioned that

these studies are based on theoretical assumptions and not on actual production figures.

The estimated CAPEX of a slow pyrolysis biorefinery producing 2000 metric tonnes of

biochar per day is US$132 million, and US$200 million for a fast pyrolysis facility (Brown et

Biochar CCS

170

al., 2011), while de Gryze et al. estimate the establishment costs for a plant consuming 50kpa

to 70kpa short tons of feedstock at US$20 million to US$45 million, with an annual OPEX of

3% - 5% of the initial investment cost and biomass sourced from within a 60km radius. One of

the limitations to upscaling a project may not be the size of the conversion unit, but the ability

to source the necessary sustainable feedstocks efficiently. A distributed network of smaller

scale units may be a more efficient and profitable solution (de Gryze et al., 2010).

The average production cost of one short ton of biochar is estimated to be between US$150

and US$350 (de Gryze et al., 2010). This would place the production price within the range of

market prices which can be obtained. Biochar for specialty application like research purposes

and niche horticultural products (Sohi, 2012) has sold for US$200/ton (Miles, 2009) to

US$500 per metric tonne (The Climate Trust, 2010). Additional revenue streams through the

sale of bio oils and energy and cost savings through the use of excess heat may be achievable

(de Gryze et al., 2010). Baum and Weitner (2006) state that the production and application

costs of Biochar CCS may be fully recoverable based on increased agricultural production

yield and nitrogen fertiliser savings, making biochar affordable.

Roberts et al. (2009) state that the main costs are feedstock and pyrolysis, with the volume

and cost of delivered biomass feedstock and market prices for biochar and energy products the

most important economic constraints. Most cost savings and revenue can come through the

use of low cost waste streams as feedstock, i.e. wood, leaves, food waste, straw or manure

(Roberts et al., 2009). Agricultural residues such as corn stover (husks, stalks, leaves and cob

left after harvesting) (Figure 5.6) have high yields of energy generation and GHG reductions,

and have moderate potential to be profitable, whereas waste biomass streams such as

municipal yard waste have the greatest potential to be economically viable while still being

net energy positive and reducing GHG emissions. Other financial benefits such as tax credits

and renewable energy certificates are also available for renewable energy projects. In 2009,

The Royal Society estimated the negative emissions of bioenergy combined with CCS in

geological formations (BECCS) as an estimated 50 – 150 ppm decrease in CO2-e. A

combination of BECCS and Biochar CCS could generate numerous value streams (Read,

2009; Lehmann et al., 2006; Brown et al., 2011; Read and Lermit, 2005). The introduction of

further economic and environmental benefits could be extended into developing nations

through pyrolysis stove technology (Driver and Gaunt, 2010). For all these reasons, viable

business case earns a 3.

Biochar CCS

171

Figure 5.6. Bales of corn stover in Nebraska, USA (Image source: Science Daily, 2007).

Government funding/incentives?

Biochar projects are already operational globally, and there has been little reliance on

government incentives to date. However, to implement commercial-scale biochar facilities

with the principal purpose of carbon sequestration, will require more governmental support of

R&D activities to upscale and improve production systems, optimise and validate agronomic

performance and better understand carbon-recalcitrance (Winsley, 2007; NRDC, 2010). The

two biochar “waste to energy” projects in Australia have received government funding; AUD

4.5 million for the Melbourne project from the Victorian State Government, and AUD 4.3

from the Commonwealth of Australia’s Regional Development Fund matched by the Ballina

Shire Council for the Ballina Plant (Business Environment Network [BEN], 2011, 2012). A

Report by the US Natural Resources Defense Council proposes a research agenda and funding

of up to $150 million over eight years to start commercial-scale pilot projects (NRDC, 2010)

with the goal of supporting US Biochar policy development. On the other hand,

implementation of Biochar CCS at household scales is very affordable. The government

funding/incentives sub-criterion balances out to a score of 2.

Biochar CCS

172

Carbon market?

Articles 3.3 and 3.4 of the Kyoto Protocol, and Article 3.3 of the UNFCCC specifically

refer to carbon stock change in soil and biomass. The recognition of enhanced soil carbon

sequestration suggests that Biochar CCS projects should be considered eligible to generate

carbon offset credits. The additionality criteria, a measure of whether a carbon mitigation

project is additional/ incremental to a business as usual case, could be based on the type of

feedstock, location, or regulatory environment (The Climate Trust, 2010). Biochar projects

could also participate under alternative carbon credit programmes (URS Australia, 2010). For

example, the Verified Carbon Standard (VCS) (formerly known as the Voluntary Carbon

Standard), a voluntary reserve that accepts global projects and is a quality standard for the

voluntary carbon offset industry. VCS establishes auditing and MMV criteria based on Kyoto

Protocol’s Clean Development Mechanism (CDM) and ISO standards. The carbon price is

low and the market immature, but as there is currently no dependency on it, the carbon market

sub-criterion merits a rating of 3.

The overall assessment of Biochar CCS is a very positive “profitable” with an average score

of 2.7, which is entered into the Economics column of the Master Matrix.

Storage Quality

Capacity?

There are varying estimates of the potential technical global storage capacity of Biochar

CCS: e.g. 0.7 to 2.6Gtpa (Laird et al., 2009), 1.8 Gtpa CO2-e (Woolf et al., 2010); and 1.2 -

3.1Gtpa (Lal, 2011). 2.35 mt CO2-e is sequestered per metric tonne of biochar, or ~2.18mt

CO2-e after subtracting emissions from the energy penalty to produce the biochar (The

Climate Trust, 2010). McCarl et al., (2009) estimate the net balance of GHG offsets per tonne

of feedstock for fast pyrolysis at 0.823CO2-e and 1.113 CO2-e for slow pyrolysis. While

technical potential does not equate to relative utility (Levitan, 2010) with logistics and

feedstock sustainability issues, the ability to utilise feedstocks from a wide range of sources

and locations globally, including purpose-grown, landfill, and under-utilised waste biomass

(Winsley, 2007), earns a capacity rating of 3.

Biochar CCS

173

Rate?

The rate at which CO2 can be captured during photosynthetic processes depends upon

individual plant growth rates, climate, rainfall, sunlight hours, location, soil, etc. The time

required to sequester the biomass through pyrolysis ranges from one second to hours

depending on the feedstock and desired ratio of C-products (Table 4.3) (IEA Bioenergy,

2007). If sustainable waste feedstocks are used, then there is no direct dependency on the

photosynthetic cycle, and the rate for this does not have to be taken into account. Therefore, in

considering the pyrolysis rate, Biochar CCS earns a score of 3 for rate.

Duration?

Under the Kyoto Protocol, projects for Land Use, Land Use Change and Forestry

(LULUCF) must demonstrate “long-term” changes to terrestrial carbon storage and

atmospheric GHG concentrations (UNFCCC, 1998a). However, there is no consistent

definition of long-term or consensus about time frames for LULUCF projects. The 100 year

approach requires that the GHG benefits of a project, that is, the carbon stored, must be

maintained for a period of 100 years to be consistent with the Kyoto Protocol’s 100 year

reference time frame (Addendum to the Protocol, Decision 2/CP.3, paragraph 3), and its

adoption of the IPCC’s Global Warming Potentials (GWPs) (Article 5.3) for calculation of the

Absolute Global Warming Potential (AGWP) of CO2 (UNFCCC, 1998b, 2000).

Stability of the biochar depends on the pyrolysis procedure utilised, but biochar is

relatively inert, highly stable and resistant to biochemical breakdown (Lehmann et al., 2006;

Sombroek et al., 2003). It is estimated that the MRT of the stable fraction of biochar ranges

from several hundreds to a few thousand years (Lehmann et al., 2006; 2009c; The Climate

Trust, 2010; Roberts et al., 2009; de Gryze et al., 2010). The average age of black carbon

buried in deep sea sediments was found to be up to 13,900 years greater than the age of other

organic carbon (Masiello and Druffel, 1998). Further research is needed to validate and

optimise stabilisation and decomposition mechanisms of biochar in soil, but does not impact

the Duration sub-criterion which earns a 3 rating.

Biochar CCS

174

Verification?

Lehmann et al. (2006) state that the sequestered carbon amounts can be easily accounted

for and verified. Comparisons can be made between baseline measurements of water, soil and

air (NO4, CO2, CH4) with repeated measurements over time. In 2012, the IBI released the first

international Biochar Standards and Testing Guidelines. These guidelines include three test

categories; utility, toxicity, and advanced analysis and soil enrichment. The utility tests assess

the physical properties (particle size and moisture) and chemical properties of elemental

proportions. Lower values of the molar ratio of hydrogen to organic carbon indicate greater

carbon stability (IBI, 2012). The IBI is also developing a certification programme for

sustainable biochar as having met the requirements of the Biochar Standards. Verification

earns a 3.

The overall rating of Storage Quality is “high” (3) and this is entered into the Master Matrix.

Environmental Impact

Potential impact on ecosystems and/or water systems?

Biochar CCS can reduce both fertiliser manufacture and use. The microbial processes of

nitrification and denitrification in the soil are significant sources of N2O, a GHG with 298

times greater global warming potential than the equivalent mass of CO2 in the atmosphere

(IPCC, 2007; Forster et al, 2007). The manufacture of nitrogen fertiliser releases >3t CO2-e

per ton of nitrogen fertiliser produced (West and Marland, 2002). The application of biochar

to soils provides GHG mitigation benefits to N2O and CH4 as well as long-term stable soil

carbon sequestration value (Van Zweiten et al., 2009; Rondon et al., 2005). In 2008, Lehmann

et al. suggest that the carbon-cycle feedback could be reduced by the application of biochar to

soils.

Contaminants may occur from contaminated feedstocks (polychlorinated biphenyls (PCBs)

and heavy metals) or form during the pyrolysis process (dioxins and polycyclic aromatic

hydrocarbons (PAHs); however, strict sorting and control of feedstock and temperatures of

<500 °C can minimise the risk of potential soil contamination or contaminant formation

(Verheijen et al., 2009; IBI, 2012). Dioxin concentrations found in biochars have been very

low, are strongly bound and are not readily bio-available (Wilson and Reed, 2012). The

reduction of GHG emissions together with the multiple benefits of increased net primary

production from soil amendment, fertility and resilience, water retention, decreased runoff,

Biochar CCS

175

decreased fertiliser and lime use, the displacement of fossil fuels, etc. all positively impact

ecosystems and water systems, therefore the potential impact on eco- and water systems

earns a very favourable (3) rating.

Environmental footprint?

Some infrastructure for commercial scale

pyrolysis plants and storage facilities for the biomass

and biochar will be necessary at centralised

locations, such as municipal landfills. The

transportation systems and road, rail and shipping

infrastructure already exist, and mobile pyrolysis

units can be utilised for remote forestry activity for

uncommercial biomass or seasonal harvesting of

sustainable biomass on agricultural land (Winsley,

2007). The environmental footprint of one operator’s

gasification facilities converting biomass to energy

plants capable of generating 0.5MW to 2 MW is just

1250 square feet (ft2) (Phoenix Energy, 2013). One

study suggests that four typical households using 5kg

of fuel per day could sequester 1t CO2 pa by using

modern pyrolysis stoves (Figure 5.7), with the

implications of potential scale up hugely significant

as an estimated two billion people still use primitive

stoves or open fires for cooking and heating (Figure 5.8) (Anderson et al., 2010). In addition,

if municipal bio-wastes are utilised as feedstock, landfill requirements will decrease, therefore

the environmental footprint for Biochar CCS balances out to score of 2.

The overall rating for Environmental Impact is on the cut-off point between “medium” and

“high” (2.5).

.

Figure 5.7.

Example of a modern pyrolysis stove

for household use in developing non-

OECD nations (Image source:

SHALIN Finland in collaboration

with Aalto University, link).

Biochar CCS

176

Figure 5.8. Typical three-stone stove used in Kenya (Image source: SHALIN Finland in collaboration

with Aalto University, link)

5.3.1 Science and Technology: Sub-Matrix Assessment

Technology Element 1: Site Identification and Characterisation

Site identification and characterisation for Biochar CCS encompasses the logistical, socio-

economic and technical feasibility of sites for pyrolysis processing plants and to a lesser

extent the application of biochar. Ideal locations to site pyrolysis facilities would be low-value

land or co-sited with collective points with existing infrastructure of municipal or forestry and

agricultural waste products, and near to electricity lines to distribute power generated from

bioenergy. In general, public acceptance of renewable energy technologies coexists with

resistance to specific development projects in the siting of infrastructure (Owens and Driffill,

2008; van der Horst, 2007; EU DGR, 2003). Identification of suitable sites for Biochar CCS

require the identification of locations such as urban or rural collection points (e.g. municipal

landfills), or low-value land suitable for dedicated feedstock cultivation with nutrient poor

soils (Winsley, 2007). The general public, with the exception of farmers and foresters who

grant land access, will not be aware of the activities to test soils, air or water. As these groups

are the main beneficiaries of Biochar CCS through waste management of crop and forestry

residues, electricity generation and application of the biochar to enrich local soils and increase

productivity yield, limited resistance is expected. Therefore Public Acceptance earns a 3.

Biochar CCS

177

Biochar CCS activities, such as land surveying, environmental protection and baseline

testing of soil, air and water, are covered under existing rules and legislation from local

authorities, state, federal and even international bodies, therefore Regulatory Frameworks gets

a high (3) rating. The minimisation of transport costs of both the biomass as well as the end

products is desired. Land with nutrition-poor soils can be utilised so both the pyrolysis process

and end-use application of the biochar for soil amendment are in close proximity. For the

distribution and sale of excess electricity generated from the pyrolysis plant, access to nearby

power grids is required. All the tests and evaluations for these factors are very affordable and

can to a certain extent be based on public domain information, therefore Economics earns a 3

rating. The tests to measure air, water, and soil to determine suitable locations for pyrolysis

processing, and/or application of the biochar and surveys to identify suitable locations are

well-established and non-invasive (Coomes et al., 2002; Post et al., 2000; MacDicken, 1997),

therefore both Demonstrated Technical Readiness and Environmental Impact earn high (3)

scores.

The overall score of Site Identification and Characterisation is a very positive 3.

Technology Element 2: Feedstock

There are two basic types of feedstock, primary and waste. Primary feedstock refers to

purpose/dedicated crops specifically produced for conversion to biochar, and can utilise

marginal or degraded land. Waste feedstock refers to unutilised biomass, or organic matter

which would have been burnt, dumped in a landfill or left to decompose, e.g. agricultural/crop

residues, animal manures, organic municipal solid waste, forestry leavings, yard and other

urban wood waste. Waste feedstocks must be sorted to remove potential biochar

contaminants, e.g. metal, glass, sewage sludge or tyres. Sustainable feedstocks that do not

endanger food security, habitat or soil conservation are preferred (Woolf et al., 2010). There

may be competition for feedstocks as some of the wastes may have been used for other

purposes.

Waste feedstocks are preferred to primary by both the public and governments because

they do not incur emissions from land use change or energy penalty from production (NRDC,

2010). The concerns about changes in land use (e.g. pastoral to forest), and/or food crops

being replaced or cultivated as biomass for fuel, balance out to a Public Acceptance rating of

2. The cultivation of feedstocks is covered under current legislation. It is proposed that

existing schemes, such as the sustainability standards under Article 17 (6) of the EU

Biochar CCS

178

Renewable Energy Directive that recognises biofuels certified under a range of schemes,

could be utilised for sustainable biochar certification of agricultural practices to produce food

and fuel (European Parliament and the Council of the European Union [EP CEU], 2009;

Cowie et al., 2012; O’Connell et al., 2009). A sustainability framework and certification

programme, adapted from existing frameworks for bioenergy, has been proposed as one

solution to provide stakeholder confidence about sustainability practices, calculation and

verification of abatement and enable claims of legitimate offset for emissions trading schemes

(Cowie et al., 2012). Regulatory Frameworks merits a “3” score.

Costs are dependent on feedstock source and volumes, but the use of waste biomass can

keep these expenses low (Brown et al., 2011). It is proposed that only biochar projects that

use feedstocks that are unlikely to cause land-use change and are not purpose-grown should

qualify for carbon market offset credits. Urban waste aggregation at centralised collection

points offer potential revenue from tipping or disposal fees, therefore Economics scores a 3.

Feedstock source is the most important factor determining recovery rates of the initial carbon

from the biomass, ranging from 39% to 64% with the typical carbon-recovery around 50%

(Lehmann et al., 2006). The moisture content of the feedstock should be kept below 10% for

production efficiencies (Lehmann and Joseph, 2009b). Current agricultural practices and

technology are well established but further R&D efforts have to be made to optimise

feedstock production, harvesting, drying and grinding (Winsley 2007; Woolf et al., 2010). A

variety of feedstocks from different locations and climates have already been the subject of

pyrolysis production (Lehmann et al., 2006); for these reasons Demonstrated Readiness earns

a 3. Some infrastructure will be necessary for storage and handling of feedstock and biochar

storage and handling. If a centralised collection point is used, e.g. municipal landfill, these

buildings will likely be within the same facility. The landfill requirements and decomposition

of bio matter causing CH4 will decrease, therefore Environmental Impact earns a 3.

The overall score for Feedstock is a “high” 2.8.

Technology Element 3: Transportation

Transportation for Biochar CCS encompasses the collection and transfer of biomass

feedstock from various dispersed points to pyrolysis plants as well as of the carbon products

for further processing, application to soil, or distribution and sale. Biochar production

facilities are unlikely to be in close proximity to residential or environmentally sensitive areas.

Existing road and rail networks and transportation systems will be utilised, but increased

Biochar CCS

179

infrastructure usage is likely. Public Acceptance of transportation scores a “3” rating. The

transport of biomass as well as the carbon-products is already covered by existing haulage

legislation, including safe handling regulations. In Australia, charcoal is classified as

spontaneously combustible (United Nations Hazardous Goods Class 4 Division 4.2) and a

minor danger (packing group III), therefore safe handling and transportation practices are

regulated to avoid potential fire hazards (UN, 2009; Lehmann and Joseph, 2009b). This earns

the Regulatory Framework a high (3) rating.

Transportation distance differs between biochar systems, affecting cost and energy

consumption Roberts et al., 2009). Aggregation of feedstocks from multiple collection points

in urban environments can take advantage of energy sources and existing local government or

private services that aggregate, sort and pre-process the biomass. Pyrolysis plants could be

located next to waste processing plant sites, e.g. landfill/municipal dumps, to minimise

logistics and transportation costs. Aggregation of distant or widely dispersed/distributed

sources, e.g. rural farms, forest, will be more expensive. If the bioenergy is not to be sold to an

energy carrier, and only used to power the pyrolysis system, siting near power lines is not

important. The pyrolisers can be scaled with size to match locally distributed sources of

biomass and minimising transport costs, i.e. on a local scale for rural/distant locations with

processing on site or use of mobile units that rotate between sites (Badger and Fransham,

2006). Lowering the moisture content of the biomass lowers the cost of transportation per unit

of biochar and energy produced (Lehmann and Joseph, 2009b). These factors balance out in a

(2) rating for Economics.

The technology for truck transportation of biofuels, hazardous materials, and more volatile

gases is already mature. This earns Demonstrated Technical Readiness a (3) score. Very fine,

light black carbon (sub micron) dust fraction can occur during transportation or utilisation

causing biochar loss and pollution, but can be minimised or managed with storage or

application practices (NRDC, 2010; Lehmann and Joseph, 2009b). There is little or no

increased environmental footprint for transportation infrastructure but possible higher

utilisation of truck haulage and road infrastructure, therefore Environmental Impact rates a

neutral (2).

The overall score for Transportation is 2.6.

Biochar CCS

180

Technology Element 4: Conversion

Conversion refers to the transformation of biomass through pyrolysis or gasification into

biochar, synfuel or liquid biofuel. The ratio between the three depends on whether fast (high

temps >500⁰C) or slow pyrolysis (300-400 ⁰C) is used, as well as pressure, water content, and

residence time (Table 5.1). In some systems, the syngas and oil can be used as a fuel to run the

pyrolysis reaction, meaning no external energy source is required beyond the organic waste

itself. Public Acceptance is based on the analogy of biofuels as renewable energies which are

preferred by both the public and government. The pyrolysis plants can be sited at existing

collection points or in remote locations away from built up areas, thereby minimising the

NIMBY affect. In addition, the making of charcoal is familiar and one of mankind’s earliest

industrial processes (Brown, 2009; Harris, 1999), therefore, Public Acceptance earns a high

(3) score. Regulatory Frameworks scores a 3 as existing regulation covers the production of

biochar and carbon-materials through pyrolysis or gasification technologies and pollution. The

Economics of conversion scores a 3 as the current technologies are affordable and range in

scale from simple ovens/kilns to dedicated production facilities to produce specialised carbon-

products for sale or use (Brown et al., 2011; Sohi, 2012; Miles, 2009; The Climate Trust,

2010).

Conversion technologies are well-established (IEA Bioenergy, 2007) but the equipment

and processes could be optimised for low-emissions and high-yield quality biochar to

maximise soil benefits (Winsley, 2007; Brown, 2009; Demirbas and Arin, 2002; Jahirul et al.,

2012). Biochar CCS R&D efforts are based on slow pyrolysis reactor technology to optimise

production of biochar and syngas rather than bio-oils. Slow pyrolysis controlled continuous

kilns designed for the purpose of Biochar CCS are now commercially available and a pilot

demonstration has been in operation since 2006 in Sydney, New South Wales, Australia (BEN

2011, 2012; Pacific Pyrolysis, 2013). Phoenix Energy assembles its 0.5MW to 2MW biomass

gasification power plants using proven technology from well-known equipment suppliers

(Phoenix Energy, 2013)(Figure 5.9). According to de Gryze et al. (2010), the largest

pyrolysers in North America are capable of processing between 200 – 250 dry short tons of

biomass per day, with several companies developing and marketing industrial-scale

pyrolysers. Modular pyrolysis plants are designed for mobility and flexibility. They not only

convert biomass into biochar, syngas and liquid bio-oils, but also use the energy from the

syngas to power the pyrolysis reaction process (Badger and Fransham, 2006). The bio-oil can

also be used as process heat or for electricity generation. Demonstrated Readiness earns a

rating of 3.

Biochar CCS

181

Figure 5.9. 500kW biomass gasification plant in Merced, California, USA (Image source: IBI, courtesy

of Phoenix Energy, 2012).

The use of temporary mobile units reduces the environmental footprint from permanent

infrastructure and can be used in widely dispersed forest or agricultural locations with no

energy provider requirement (Badge and Fransham, 2006). Conversion of the biomass using

pyrolysis does include the potential release of GHG emissions or air pollutants, with potential

harmful emissions, e.g. dioxins causing air quality/pollutant issues if contaminated waste

feedstocks are used. This can be minimised through good management practices and is

covered under existing regulations. The balancing of these factors gives a score of 2 for

Environmental Impact.

The overall score for Conversion is “ready/short-term” (2.8).

Technology Element 5: MMV

MMV refers to the use of tools and tests to measure and monitor carbon in soils, biomass

and air, to obtain data/geochemical data (inSAR, fluid, mineralisation, air, soil sampling, etc.)

to build an accurate accounting of the stored volume of carbon and its permanence for GHG

accounting and economic purposes. Key methods to monitor carbon in soils include

biomarker analysis, chemo-thermal oxidation, chemical oxidation, thermal/optical

transmittance and reflectance, UV oxidation, thermal analysis, and infrared spectroscopy

(Manning and Lopez-Capel, 2009; de Gryze et al., 2010). Reversal of carbon sequestration

could occur through decomposition, erosion, burning, tilling, soil removal or land

Biochar CCS

182

development. If carbon offsets are claimed or sold, it is proposed that project developers must

demonstrate the carbon is still present 100 years after biochar application, as is the case for

forestry projects under the Climate Action Reserve (The Climate Trust, 2010). Two

challenges for MMV are that soil types could vary over small distances and feedstocks change

(NRDC, 2010). De Gryze et al. (2010) suggest repeat measurements at years 1, 5, 10, 20, and

50 years after the biochar is incorporated with the soil to quantify the amount of biochar

remaining in the soil. The monitoring systems could be based on the Biochar CCS life-cycle

(suitable for CDM) or more narrow definitions. The monitoring systems will be tailored for

projects based on the product boundaries and baselines: as projects will differ widely in terms

of biochar production, end-use, and impact on agricultural emissions (Gaunt and Cowie,

2009).

Public acceptance of MMV is rated a 3, as stakeholders want confidence that the carbon

remains sequestered by comparing repeated measurements of soil properties with baseline

measurements. In addition, the public wants confirmation of the associated environmental and

agricultural benefits, such as better nutrient and H2O holding capacity, less use of fertilisers,

increased crop yield, and avoided emissions of CH4, NO2, and CO2. Current legislation covers

the application and testing of biochar (NETL DOE, 2010), so Regulatory Frameworks scores

a 3. The project scale influences logistical and economic support of MMV activities. Large

projects (lifetime reductions - minimum 50,000Mt CO2-e, market-preferred >200,000Mt CO2-

e) require economies of scale to justify verification costs (The Climate Trust, 2010). Vertical

integration, where the biochar producer is also the user, is the most economically feasible

solution to account for, track and monitor where it is incorporated into soils. Smaller projects

are challenging (NRDC, 2010) but could achieve economies of scale through aggregation

(Driver and Gaunt, 2010). Economics scores an “affordable” (2) rating. The remote sensing,

laboratory and field work tools to measure and monitor carbon in biomass, atmosphere, and

soil are well established and commercially available (Coomes et al., 2002; Post et al., 2000;

MacDicken, 1997; NETL DOE, 2010), thereby Demonstrated Readiness earns as 3.

Environmental Impact scores a 3 as the tests do not create a permanent footprint.

MMV scores a “ready” (2.8) rating.

Biochar CCS

183

Summary of Science and Technology Sub-Matrix Assessment

All of the technology elements specific to Biochar CCS rate either medium (2) or high (3)

(Table 5.2) because the basic scientific and technical foundations for biochar and bioenergy

co-production have been demonstrated or are in operation in locations worldwide. R&D

activities focus on verification of the economic and agronomic benefits. The first is the

scientific validation and maximisation of the environmental and agricultural productivity

benefits and evaluation of the risks of the Biochar CCS life cycle. The second is the

optimisation of pyrolysis systems for low energy requirements, and improved efficiencies for

high yield of quality biochar, with gaseous output limited to H2O and CO2.

Table 5.2. Science and Technology Sub-Matrix Scorecard: Results of Application of Methodology to

Biochar CCS.

5.3.2 Summary of Results: Overall Technology Rating

The overall rating for Biochar CCS at 2.7 (Table 5.3) predicts a high probability of success

as a CO2 sequestration option. A significant benefit of this carbon mitigation option is the

immediate availability to apply it at scales ranging from the individual household to industrial

scale, in urban and rural regions, and in all countries.

Table 5.3. Master Matrix Scorecard: Results of Application of Methodology to Biochar CCS showing

overall technology rating.

Biochar CCS

184

Biochar CCS

185

5.4 DISCUSSION

My initial expectation of a favourable rating for Biochar CCS was upheld. The main

contributing factors to such a favourable rating are the generally high acceptance of

renewables, the multiple environmental and agricultural benefits, the relatively modest

CAPEX and OPEX, multiple income streams, the ability to be applied immediately at

different scales, and in urban and rural locations all over the world. A capacity goal of

1.8GtC-e (Woolf et al., 2010) could be achieved by not relying solely on large commercial-

scale projects, but by simultaneous deployment of projects at different scales. For example,

volumetric economies of scale by replacing inefficient traditional stoves in developing nations

with modern pyrolysis stoves (WorldStove, 2009; Surridge, 2011; Measurable Offsets, 2010),

biochar production facilities tied with energy generation (Phoenix Energy, 2013), as well as

commercial-scale operations tied with waste and landfill management (Ballina Information,

2012; BEN 2011, 2012; Pacific Pyrolysis, 2013; de Gryze et al., 2010). This score is bolstered

by the ability to refer to existing regulatory frameworks for nearly all of the sub-criteria.

Suggestions to improve uptake:

A multi-scale deployment approach:

Implementation on a household-scale for developing nations, i.e. cooking stoves

and community kilns (Ewing and Msangi, 2009; Bailis, 2009) could aid with faster

uptake while providing multiple environmental and soil amendment benefits, e.g.

increased agricultural productivity yields and soil water retention. Biochar CCS

projects utilising household stoves and stove hubs are operational in developing

nations such as Haiti, Kenya, and Indonesia. The goals of the WorldStove Five-

Step Program are not only to sequester CO2, but also to offer health, soil

conservation, and agricultural and socio-economic benefits, such as the setting up

of businesses, job creation, and certified and measurable carbon offsets (Surridge,

2011; WorldStove, 2009). There could be MMV logistics issues with such an

implementation if carbon credits are desired.

In areas where access to affordable feedstock is an issue, such as locations with

competition for biomass, an initial focus on the deployment of biomass to energy

conversion can enable more robust economic scenarios while still capturing some

of the Biochar CCS benefits. Once a predictable carbon credit market with the

Biochar CCS

186

carbon credit accountability for Biochar CCS exists and allows for improved

economic predictions, the re-tooling of the conversion facilities to optimise biochar

production could be considered.

Large-scale deployment for urban waste management using existing municipal

transportation, contractors, and landfills as centralised collection points where

infrastructure can be co-located. Additional revenues can be generated from

householder tipping fees.

The use of mobile units for more rural, forest, and seasonal feed stocks, with a

focus on undervalued biomass and waste. If more permanent facilities are

preferred, economies of scale can be achieved by sharing the CAPEX and OPEX

between a collective of district/regional growers/forestry and will also minimise

transportation costs.

Focus on the food security of developing nations. In 2011, Lal suggests that adoption

of Biochar CCS and other soil conservation and agricultural practices in developing

countries could increase the carbon pool in the root zone by 1 tpa, thereby increasing

crop yield by 24 - 32 million tpa for food grains and 6-10 million tpa for roots and

tubers. Further research is needed to verify the different plant responses to different

biochars under varying soil conditions (Waters et al., 2011; Winsley, 2007). The

application of biochar to nutrient-poor soil, produced by individual household

cooking stoves using the latest pyrolysis techniques, will help with food security

through soil enrichment/amendment, as well as offsetting atmospheric GHG.

Regulation and logistics of power grids must be supportive of electricity generation from

dispersed sites.

Economic values need to be assigned to the environmental benefits of Biochar CCS to

encourage the technology’s uptake, i.e. soil carbon sequestration, reduced GHG

emissions, and reduced nitrate leaching into waterways, etc. (Winsley, 2007).

More governmental and international body funding to support:

R&D efforts to validate and optimise the environmental and agricultural benefits

of Biochar CCS.

Demonstration projects of the Biochar CCS life cycle at different scales,

especially commercial-scale.

Biochar CCS

187

5.5 CONCLUSIONS

As a renewable carbon neutral/negative mitigation solution, Biochar CCS represents

significant potential to sequester CO2 at the volumes needed to achieve atmospheric GHG

targets through deployment at both household and commercial-scales. Using sustainable

feedstocks, Biochar CCS has the potential to offset 12% current anthropogenic CO2

emissions, or 1.8Gt CO2-e pa of the 15.4 Gtpa emitted annually in comparison to 10% for

bioenergy (Woolf et al., 2010). Simultaneously, it can provide multiple economic and

environmental benefits: food security, soil enrichment, waste and environmental

management, etc. The technology is proven, affordable and can be applied immediately

worldwide in a variety of locations at scales ranging from the individual household to

industrial scale. Commercial-scale demonstration projects (>500 Mtpa) would assist in

technology uptake. There are revenue streams from the biochar, energy products and

specialised high value commodities. R&D efforts focus on validating the environmental and

agricultural benefits, as well as optimising pyrolysis systems. More field and industrial trials

are needed to establish long-term effects. The main issue is the sustainability of feedstocks,

but lessons have been learned from first-generation biofuels to ensure the technology

elements of the Biochar CCS life cycle are applied properly with sustainability protocols and

certification. Biochar CCS is one of the few technologies to address climate change that

creates net economic as well as environmental benefits with low risk.

Biochar CCS

188

Biochar CCS

189

REFERENCES

Amonette, J.E. and Joseph, S. (2009), Characteristics of Biochar: Microchemical Properties: In

Biochar for Environmental Management: Science and Technology, [Lehmann, J., and Joseph, S.

(Eds.)], Earthscan, London, United Kingdom, 3, 33-52.

Anderson, P., Taylor, P., and Wever, P. (2010), CHAB Micro‐Gasification for 1GtCO2/yr

Mitigation‐Sequestration: Presented at IBI 2010, September 12 –15, 2010, Rio de Janeiro, Brazil.

Badger, P.C.and Fransham, P. (2006), Use of mobile fast pyrolysis plants to densify biomass and

reduce biomass handling costs – A preliminary assessment: Biomass Bioenergy, 30, 4, 321-325.

Bailis, R. (2009), Modelling climate change mitigation from alternative methods of charcoal production

in Kenya: Biomass Bioenergy , 33, 1491-1502.

Ballina Information (2012), Ballina gets $8.5 biochar venture: Ballina Information, 14 June 2012.

http://www.ballina.info/blog/2012/06/13/ballina-gets-8-5m-biochar-venture/ (accessed 9 February

2013).

Baum, E., and Weitner, S. (2006), Biochar application on soils and cellolosic ethanol production:

Clean Air Task Force, Washington.

Business Environment Network [BEN] (2011), Pac Pyro bring WtE to Melbourne: Business

Environment Network, 6 September 2011. http://ben-

global.com/storyview.asp?storyID=9586427&section=Advanced+Technologies&sectionsource=s14

50114#.UWTt4Mqre3M (accessed 9 February 2013).

Business Environment Network [BEN] (2012), $4.3m brings second Pac Pyro plant closer: Business

Environment Network, 12 June 2012. http://www.ben-

global.com/storyview.asp?storyID=9588438&section=Advanced+Technologies&sectionsource=s14

50114#.UWTsxcqre3M (accessed 9 February 2013).

Brown, R. (2009), Biochar Production Technology: In Biochar for Environmental Management:

Science and Technology, [Lehmann, J., and Joseph, S. (Eds.)], Earthscan, London, United Kingdom,

8, 127-146.

Brown, T.R., Wright, M.M., and Brown, R.C. (2011), Estimating profitability of two biochar

production scenarios: slow pyrolysis vs fast pyrolysis: Biofuels, Bioproducts and Biorefining, 5, 1,

54-68.

Biochar CCS

190

Coomes, D.A., Allen, R.B., Scott, N.A., Goulding, C., and Beets, P. (2002), Designing systems to

monitor carbon stocks in forests and shrublands: Forest Ecology and Management, 164, 1, 89-108.

Cowie, A.L., Downie, A.E., George, B.H., Singh, B.P., Van Zwieten, L., and O'Connell, D. (2012), Is

sustainability certification for biochar the answer to environmental risks?: Pesquisa Agropecuária

Brasileira, 47, 5, 637-648.

De Gryze, S., Cullen, M., Durschinger, L., Lehmann, J., Bluhm, D., Six, J. and Suddick, E. (2010),

Evaluation of the Opportunities for Generating Carbon Offsets from Soil Sequestration of Biochar:

An issues paper commissioned by the Climate Action Reserve.

http://www.climateactionreserve.org/wp-

content/uploads/2009/03/Soil_Sequestration_Biochar_Issue_Paper1.pdf (accessed 23 February

2013).

Demirbas, A., and Arin, G. (2002), An overview of biomass pyrolysis: Energy Sources, 24, 5, 471-

482.

Downie, A., Munroe, P., Cowie, A., Van Zwieten, L., and Lau, D. (2012), Biochar as a geoengineering

climate solution: Hazard identification and risk management: Critical reviews in environmental

science and technology 42, 3, 225-250.

Driver, K., and Gaunt, J. (2010), Bringing Biochar Projects into the Carbon Marketplace: An

introduction to carbon policy and markets, project design, and implications for biochar protocols.

http://www.biocharprotocol.com/sg_userfiles/policy_primer.pdf (accessed 24 January 2013).

Ewing, M., and Msangi, S. (2009), Biofuels production in developing countries: assessing tradeoffs in

welfare and food security: Environmental Science and Policy, 12, 520-528.

European Commission. Directorate General for Research [EU DGR] (2003), Energy: issues,

options and technologies, science and society, 20624 [European Opinion Research Group (Ed.)]:

European Commission.

Food and Agriculture Organisation [FAO] (1983), Simple technologies for charcoal Making: FAO

Forestry Paper 41: Food and Agriculture Organisation of the United Nations, Rome, Italy.

Forster, P., Ramaswamy, V., Artaxo, P., Berstsen, T., Betts, R., Fahey, D.W., Haywood, J., Lean, J.,

Lowe, D.C., Myhre, G., Nganga, J., Prinn, R., Raga, G.M.S., and Van Dorland, R. (2007), Changes

in atmospheric constitutents and in radiative forcing: In Climate change 2007: The physical science

basis, [Solomon S., Qin, D., Manning M., et al (eds.)], Contribution of working group I to the fourth

assessment report of the Intergovernmental Panel on Climate Change: Cambridge University Press,

Cambridge, United Kingdom, 129-234.

Biochar CCS

191

Fowles, M. (2007), Black carbon sequestration as an alternative to bioenergy: Biomass and Bioenergy,

31, 426–432.

Gaunt, J. and Cowie, A. (2009), Biochar, greenhouse gas accounting and emissions trading: In

Biochar for Environmental Management: Science and Technology, [Lehmann, J., and Joseph, S.

(Eds.)], Earthscan, London, United Kingdom, 11, 317-340.

Glaser, B., Haumaier, L., Guggenberger, G. and Zech, W. (2001), The “Terra Preta” phenomenon: A

model for sustainable agriculture in the humid tropics: Naturwissenschaften, 88, 37-41.

Harris, P. (1999), On charcoal: Interdisciplinary Science Reviews, 24, 4, 301-316.

International Biochar Initiative [IBI] (2012), Standardized Product Definition and Product Testing

Guidelines for Biochar That Is Used in Soil. http://www.biochar-

international.org/sites/default/files/Guidelines_for_Biochar_That_Is_Used_in_Soil_Final.pdf

(accessed 23 January 2013).

International Biochar Initiative [IBI] (2012), Phoenix Energy’s business model: Building small

profitable plants: [in September 2012 newsletter]. http://www.biochar-

international.org/profile/Phoenix_Energy (accessed 24 April 2013).

International Energy Agency Bioenergy [IEA Bioenergy] (2007), Biomass Pyrolysis: IEA

Bioenergy [Bridgwater, T]: Aston University, United Kingdom, T34:2007:01.

Intergovernmental Panel on Climate Change [IPCC] (2000), Land Use, Land Use Change, and

Forestry,: A Special Report: Intergovernmental Panel on Climate Change, Cambridge University

Press, Cambridge and New York.

Intergovernmental Panel of Climate Change [IPCC] (2007), Climate Change 2007: The Physical

Science Basis: Contribution of Working Group I to the Fourth Assessment Report of the

Intergovernmental Panel on Climate Change, [Solomon, S., Qin, D., Manning, M., Marquis, M.,

Averyt, K., Tignor, M. M. B., Miller Jr., H.L., and Chen, Z.], Cambridge University Press,

Cambridge, UK and New York, NY, USA.

Ipsos-MORI (2010), Experiment Earth: Report on a Public Dialogue on Geoengineering: Natural

Environment Research Council, Swindon, United Kingdom.

www.nerc.ac.uk/about/consult/geoengineering-dialogue-final-report.pdf (accessed 21 February

2013).

Biochar CCS

192

Itaoka, K., Saito, A., Paukovic, M., de Best-Waldhober, M., Dowd, A-M., Jeanneret, T., Ashworth, P.

and James, M. (2012), Understanding how Individuals Perceive Carbon Dioxide: Implications for

Acceptance of Carbon Dioxide Capture and Storage: CSIRO, Report EP 118160, Australia.

Jahirul, M. I., Rasul, M. G., Chowdhury, A. A., and Ashwath, N. (2012), Biofuels production through

biomass pyrolysis—A technological review: Energies, 5, 12, 4952-5001.

Laird, D.A. (2008), The charcoal vision: A win-win-win scenario for simultaneously producing

bioenergy, permanently sequestering carbon, while improving soil and water quality: Agronomy

Journal, 100, (1) 178 - 181.

Laird, D.A., Brown, R.C., Amonette, J.E., Lehmann, J. (2009), Review of the pyrolysis platform for

coproducing bio-oil and biochar: Biofuels, Bioproducts and Biorefining, 3, 547-562.

Lal, R. (2004), Soil carbon sequestration impacts on global climate change and food security: Science,

304, 5677, 1623-1627.

Lal, R. (2011), Sequestering carbon in soils of agro-ecosystems: Food Policy, 36, S33-S39.

Lehmann, J., Gaunt, J. and Rondon, M. (2006), Biochar sequestration in terrestrial ecosystems – a

review: Mitigation and Adaptation Strategies for Global Change, 11, 403-427.

Lehmann, J., Skjemstad, J., Sohi, S., Carter, J., Barson, M., Falloon, P.,Coleman, K.,Woodbury, P. and

Krull, E. (2008), Australian climate–carbon cycle feedback reduced by soil black carbon: Nature

Geoscience, 1, 832 – 835.

Lehmann, J., and Joseph, S., (2009a), Biochar for Environmental Management: An Introduction: In

Biochar for Environmental Management: Science and Technology, [Lehmann, J., and Joseph, S.

(Eds.)], Earthscan, London, United Kingdom, 1, 1-12.

Lehmann, J., and Joseph, S., (2009b), Biochar Systems: In Biochar for Environmental Management:

Science and Technology, [Lehmann, J., and Joseph, S. (Eds.)], Earthscan, London, United Kingdom,

9, 147-168.

Lehmann, J., Czimczik, C., Laird, D., and Sohi, S. (2009c), Stability of Biochar in the Soil: In Biochar

for Environmental Management: Science and Technology, [Lehmann, J., and Joseph, S. (Eds.)],

Earthscan, London, United Kingdom, 11, 183-205.

Levitan, D. (2010), Refilling the carbon sink: Biochar’s potential and pitfalls: Yale Environment 360.

http://e360.yale.edu/feature/refilling_the_carbon_sink_biochars_potential_and_pitfalls/2349/

(accessed 25 January 2013).

Biochar CCS

193

Liang, B., Lehmann, J., Soloman, D., Kinyangi, J., Grossman, J., O’Neill, B., Skjemstad, J.O., Thies,

J., Luizão, F.J., Petersen, J., and Neves, E.G. (2006), Black carbon increases cation exchange

capacity in soils: Soil Science Society of America Journal, 17, 69-74.

MacDicken, K. G. (1997), A Guide to Monitoring Carbon Storage in Forestry and Agroforestry

Projects: Winrock International Institute for Agricultural Development, Arlington, VA, USA.

Malkow, T. (2004), Novel and innovative pyrolysis and gasification technologies for energy efficient

and environmentally sound MSW disposal: Waste Management, 24, 53–79.

Manning, D.A.C. and Lopez-Capel, E. (2009), Test Procedures for Determining the Quantity of

Biochar within Soils: In Biochar for Environmental Management: Science and Technology,

[Lehmann, J., and Joseph, S. (Eds.)], Earthscan, London, United Kingdom, 17, 301-315.

Marland, G., Fruit, K., and Sedjo, R. (2001), Accounting for sequestered carbon: the question of

permanence: Environmental Science & Policy, 4, 6, 259-268.

McCarl, B.A., Peacocke, C., Chrisman, R., Kung. C.C. and Sands, R.D. (2009), Economics of Biochar

Production, Utilization and Greenhouse Gas Offsets: In Biochar for Environmental Management:

Science and Technology, [Lehmann, J., and Joseph, S. (Eds.)], Earthscan, London, UK, 19, 341-

357.

Measurable Offsets (2010), WorldStove certified as carbon offset: WorldStove’s five-step program

and the LuciaStove are now a fully certified carbon sequestration method: Measureable Offsets.

http://measurableoffsets.com/certified.html (accessed 24 April 2013).

Miles, T. (2009), The Economics of Biochar Production: Presentation to the Pacific Northwest Biochar

Association, Richland, Washington, USA.

http://trmiles.com/Presentations/Biochar/TRMiles%2005%2020%2009%20PNWBiochar.pdf

(accessed 25 January 2013).

Moscowitz, C.M. (1978), Source Assessment: Charcoal Manufacturing - State of the Art: EPA-600/2–

78-004z: US Environmental Protection Agency, Cincinnati, OH, USA.

Müller, A., Schmidhuber, J., Hoogeveen, J., and Steduto, P. (2008), Some insights in the effect of

growing bio-energy demand on global food security and natural resources: Water Policy, 10, 83-84.

Natural Resources Defense Council [NRDC] (2010), Biochar: Assessing the Promise and Risks to

Guide U.S. Policy, NRDC Issue Paper November 2010, New York, NY, USA.

Biochar CCS

194

National Energy Technology Laboratory, Department of Energy [NETL DOE] (2010), Terrestrial

Sequestration of Carbon Dioxide: National Energy Technology Laboratory (NETL), Department of

Energy (DOE). http://www.netl.doe.gov/technologies/carbon_seq/refshelf/BPM_Terrestrial.pdf

(accessed 25 April 2013).

Neves, E.G., Petersen, J.B., Bartone, R.N. and Silva, C.A.D. (2003), Historical and socio-cultural

origins of Amazonian Dark Earths: In Amazonian Dark Earths: Origin, Properties, Management,

[Lehmann, J., Kern, D.C., Glaser, B, and Woods, W.I. (Eds.)]: Kluwer Academic Publishers,

Dordrecht, The Netherlands, 29-50.

O’Connell, D., Braid, A., Raison, J., Handberg, K., Cowie, A., Rodriguez, L. and George, B. (2009),

Sustainable Production of Bioenergy: A review of global bioenergy sustainability frameworks and

assessment systems: Rural Industries Research and Development Corporation, Barton, Australia.

Owens, S., and Driffill, L. (2008), How to change attitudes and behaviours in the context of energy:

Energy Policy, 36, 4412–4418.

Pacific Pyrolysis (2013), Technology. http://pacificpyrolysis.com/technology.html (accessed 1 March

2013).

Phoenix Energy (2013), Onsite biomass gasification plants. http://www.phoenixenergy.net/powerplan

(accessed 24 April 2013).

Post, W. M., Izaurralde, R. C., Mann, L. K., and Bliss, N. (2000), Monitoring and verifying changes of

organic carbon in soil: Climatic Change, 51, 1, 73-99.

Read, P. and Lermit, J. (2005), Bio-energy with carbon storage (BECS): A sequential decision

approach to the threat of abrupt climate change: Energy, 30, 2654–2671.

Read, P. (2009), Policy to Address the Threat of Dangerous climate Change: A Leading Role for

Biochar: In Biochar for Environmental Management: Science and Technology, [Lehmann, J., and

Joseph, S. (Eds.)], Earthscan, London, United Kingdom, 8, 127-146.

Roberts, K.G., Gloy, B.A., Joseph, S., Scott, N.R., and Lehmann, J. (2009), Life cycle assessment of

biochar systems: Estimating the energetic, economic, and climate change potential: Environmental

Science and Technology, 44, 2, 827-833.

Rondon, M., Ramirez, J. A., and Lehmann, J. (2005), Charcoal additions reduce net emissions of

greenhouse gases to the atmosphere: in Proceedings of the 3rd USDA Symposium on Greenhouse

Gases and Carbon Sequestration in Agriculture and Forestry, 21-24.

Biochar CCS

195

Scottish Environmental Protection Agency [SEPA] (2012), Position Statement – Manufacture and

Use of Biochar from Waste.

http://search.sepa.org.uk/sepa/C=eJwdy9kNgDAMBNEthgKgImSSFURKHGMDonyOz3nS7FhKT

5s4KqhQBE1gW1ceJWFEZWuCCWLm*X4hUoPyomO4LeYqusa3*kWVpTLj8JOfzC65nPFuD!9

MIeY_?action=previewandq=biocharandpreviewindex=0andurl=http%3a%2f%2fwww.sepa.org.uk

%2fwaste%2fwaste_regulation%2fguidance__position_statements.aspx (accessed 21 February

2013).

Sohi, S.P. (2012), Carbon storage with benefits: Science, 338, 6110, 1034-1035.

Sombroek, W., Ruivo, M.L., Fearnside, P.M., Glaser, B., and Lehman, J. (2003), Amazonian Dark

Earths as carbon stores and sinks: In Amazonian Dark Earths: Origin, Properties,Management,

[Lehmann, J., Kern, D.C., Glaser, B. and Woods W.I. (Eds)], Kluwer Academic Publishers,

Dordrecht, Netherlands, 7, 125-139.

Standards Australia (2012), Australian Standard AS-4454 (2012) Composts, Soil Conditioners and

Mulches: Standards Australia Limited, Sydney, NSW, Australia.

Surridge, T.M. (2011), The WorldStove’s Five-Step Program’s Social Value Creation: Copenhagen

Business School. http://worldstove.com/wp-content/uploads/download/research_paper.pdf

(accessed 25 April 2013).

The Climate Trust (2010), Carbon Market Investment Criteria for Biochar Projects: Final Report

prepared for the California Energy Commission, [Weisberg, P., Delaney. M. , and Hawkes, J. (main

authors)]: CEC-500-02-004, CA, USA.

The Royal Society (2009), Geoengineering the Climate: Science, Governance and Uncertainty: The

Royal Society, [Chairman: Shepherd, J. G.], RS Policy document 10/09, London, UK.

United Nations Framework Convention on Climate Change [UNFCCC] (1992), United Nations

Framework Convention on Climate Change.

http://unfccc.int/files/essential_background/background_publications_htmlpdf/application/pdf/conv

eng.pdf (accessed 21 February 2013).

United Nations Framework Convention on Climate Change [UNFCCC] (1998a), Kyoto Protocol to

the United Nations Framework Convention on Climate Change.

http://unfccc.int/resource/docs/convkp/kpeng.pdf (accessed 31 August 2011).

Biochar CCS

196

United Nations Framework Convention on Climate Change [UNFCCC] (1998b), Report of the

Conference of the Parties on its Third Session, held at Kyoto from 1 to 11 December 1997:

Addendum: Part Two: Action Taken by the Conference of the Parties at its Third Session.

http://unfccc.int/resource/docs/cop3/07a01.pdf#page=31 (accessed 26 April 2013).

United Nations Framework Convention on Climate Change [UNFCCC] (2000), 5.3.4 Project

Duration: In IPCC Special Report on Land Use, Land Use Change and Forestry.

http://www.grida.no/publications/other/ipcc_sr/?src=/climate/ipcc/land_use/267.htm (accessed 26

April 2013).

United Nations [UN] (2009), Transport of Dangerous Goods: Model Regulations Volume 1.

http://www.unece.org/fileadmin/DAM/trans/danger/publi/unrec/rev16/English/Volume1.pdf

(accessed 26 February 2013).

URS Australia (2010), Pre-Feasibility Assessment of a Thermal Conversion Facility for the Australian

Capital Territory: Final Report prepared for the Department of the Environment, Climate Change,

Energy and Water.

http://www.environment.act.gov.au/__data/assets/pdf_file/0012/210504/URS_thermal-conversion-

technologies_Final.pdf (accessed 10 February 2013).

Verheijen, F.G.A., Jeffery, S., Bastos, A.C., van der Velde, M., and Diafas, I. (2009), Biochar

Application to Soils - A Critical Scientific Review of Effects on Soil Properties, Processes and

Functions: EUR 24099 EN, Office for the Official Publications of the European Communities,

Luxembourg.

Waters, D., Zwieten, L., Singh, B. P., Downie, A., Cowie, A. L., and Lehmann, J. (2011), Biochar in

soil for climate change mitigation and adaptation: In Soil Health and Climate Change: Springer

Berlin Heidelberg, 15, 345-368.

West, T.O. and Marland, G. (2002), A synthesis of carbon sequestration, carbon emissions, and net

carbon flux in agriculture: comparing tillage practices in the United States, Agriculture Ecosystems

and Environment, 91, 217-232.

Whitman, T. Scholz, S.M. and Lehmann, J. (2010), Biochar projects for mitigating climate change: An

investigation of critical methodology issues of carbon accounting: Carbon Management , 1, 1, 89 -

107.

Wilson, K., and Reed, D. (2012), Implications and Risks of Potential Dioxin Presence in Biochar:

White Paper for the International Biochar Institute. http://www.biochar-

international.org/sites/default/files/IBI_White_Paper-

Implications_of_Potential_%20Dioxin_in_Biochar.pdf (accessed 13 March 2013).

Biochar CCS

197

Winsley, P. (2007), Biochar and bioenergy production for climate change mitigation: New Zealand

Science Review, 64, 1, 5-10.

Woolf, D., Amonette, J.E., Street-Perrott, F.A., Lehmann, J., and Joseph, S. (2010), Sustainable

biochar to mitigate global climate change: Nature Communications: 1, 56, 1-9.

WorldStove (2009), WorldStove Five-Step Program: A Five-Step Program to Reduce GHG Emissions,

Conserve Fuelwood, and bring Industry to Developing Nations: WorldStove.

http://worldstove.com/wp-content/uploads/download/five_step.pdf (accessed 24 April 2013).

Discussion and Conclusions

198

Discussion and Conclusions

199

CHAPTER 6

DISCUSSION AND CONCLUSIONS

OVERVIEW

This chapter reviews the applicability and challenges in developing the methodology,

potential modifications for specific scenarios and general improvements which could be made.

I discuss the outcome of the analysis of Geosequestration, Algae CCS and Biochar CCS, the

three carbon capture and sequestration (CCS) technologies to which the methodology is

applied, and the impact for a portfolio management approach. I finish the chapter with my

conclusions.

6.1 DISCUSSION

Methodology

The methodology provides a standardised format for decision makers to assess CCS

technologies from different business, legal, social, scientific and engineering disciplines. It is

intended as a first pass risk assessment and big picture comparison of the current status of

alternative carbon sequestration options for the purpose of a carbon management

strategy/portfolio. CCS technologies to suit corporate strategy, situations or regions can be

identified that represent the lowest risk, and an understanding of the main challenges which

the various technologies may face can be gained. While the methodology was applied in this

thesis to a global comparison of various available technologies, it can also be applied to a

portfolio of projects within a given carbon sequestration technology class.

The first pass evaluation of a portfolio by utilising this methodology will highlight the

strengths of a CCS technology, and also single out areas with the need for further

improvements or a more in-depth analysis. At the current stage, with the methodology applied

to three CCS technologies, the outcome of the assessment is in line with the evaluation by

published sources (ZEP, 2012; US DOE, 2010; SBC Energy Institute, 2012, 2013; NNFCC,

Discussion and Conclusions

200

2012; IEA, 2011; IPCC, 2011; Waters et al., 2011; NRDC, 2010; Downie et al., 2012; Global

CCS Institute, 2012). The overall methodology framework has been designed for a high level of

flexibility to allow for easy modification to account for technology progress, increased amount

of public domain reference data, and the impact of recent events.

Currently, the methodology ranking system is based on a qualitative assessment of sub-

criteria to establish a quantitative score for each of the evaluation criterion. This approach was

used to allow for the assessment of both quantitative and qualitative data under one rating

system. An additional challenge was the diversity of influence factors which ranged from the

purely technical to socio-economical. This complicated the definition of cut-off points in the

form of numerical values. Even though purely technical sub-criteria like Storage Capacity

suggest that a hard cut-off point could be established, the influence of economic parameters and

social constraints to establish the actual available capacity, as well as the immature state of

some technologies with wide ranging estimates, complicate this significantly. To address this

issue, I broke down each of the evaluation criteria into sub-criteria, defined by assessing a very

narrow topic. Besides reducing the user bias to some extent by providing a guideline and

framework for the assessment, it also allows for introduction of hard cut-off limits at a low-

level, thereby providing flexibility in the adaptation.

This approach was calibrated and tested with Geosequestration and proved to be a workable

solution. A challenge for the end-user which may arise from this approach is the risk of double-

dipping, as factors like Regulatory Framework or Economics are assessed at two different

levels. They represent at one level the global view for the given Geosequestration technology in

the form of an evaluation criterion, but are also assessed at a lower level, specifically for a given

element in the Technology Sub-Matrix. To prevent double-dipping, assigning the same score on

both levels, will require some discipline from the end user. For this reason, it is suggested to

justify assigned scores with case studies where available.

While I aim to provide a global view of the analysed technologies, a bias towards

experiences in certain geographical areas exists due to the majority of projects being located

there. Therefore offshore Geosequestration reviews as an option with very limited public

opposition, but the technology is at the current time only deployed in one country, Norway, and

at an advanced stage of implementation in just another one, Australia. Similar region-specific

bias can be found for other regions due to their character of being a research hub. I attempt to

reflect this in my analysis by analysing overall worldwide trends. A separate approach if

sufficient case studies are available may be the use of an average of the technology score for

each geographic region, thereby to some extent suppressing the extreme project density bias.

Discussion and Conclusions

201

In assessing Algae CCS and Biochar CCS, very limited public domain information is

available for certain evaluation criteria and no case studies at commercial scale could be

identified. For these technologies, I was required to develop expectation scenarios for the

calibration stage of the methodology to assess the applicability of the same and iterate for model

updates when necessary. This can have introduced a positive or negative bias towards the

technology due to the limited number of references on which the expectation scenario was

based. As more information and demonstration projects are published, the relevance of the

evaluation criteria in their current form should be reviewed. For the technology analysis, I use

analogues in cases where no public domain information is available to evaluate the sub-criteria.

If in-house experience or confidential data is available to the end user, a review of the score may

be appropriate. As the framework and rating system are designed for easy update and revision,

this can be achieved without a significant resource demand.

Future improvements to the methodology could be introduced through a larger dynamic

range of the rating scale. While this may not be relevant for the current global overview, the

application to a portfolio of projects within the same technology class may require a more

detailed rating to highlight the benefits of one project over the other. A further suggestion would

be to introduce weights to account for region specific or corporate strategy specific emphasis.

At the current stage, all the evaluation criteria are equally weighted. For example, a heavier

weighting can be applied to the evaluation criterion of Public Acceptance when assessing

Geosequestration projects in tightly populated areas where policy makers are very receptive to

public opinion. This would be the case in countries, such as the Netherlands, where a strong

negative perception of onshore storage of carbon dioxide (CO2) in geological formations exists

and has led to delays or cancellation of projects (Kuijper, 2011; Reuters, 2011). This risk would

be more accurately reflected by a higher Public Acceptance weight for a project in the

Netherlands in comparison to more accepting jurisdictions where industrial development may

trump public concerns. Similarly, a low Environmental Impact score from infrastructure

footprint should have a much higher weight in environmentally or culturally sensitive areas

where planning permissions are more difficult to obtain than in a location where such issues will

not be encountered.

If it is the case that some of the criteria become obsolete or of minor global importance, the

methodology framework should be reviewed. On a global scale, this may happen if a unified

global Regulatory Framework is introduced under which project operators can plan and operate.

This may already be the case when assessing a project portfolio within the same jurisdiction. In

this situation, a differentiation of projects by the Regulatory Framework evaluation criteria may

not be relevant. On a global scale, a fully-fledged review of the methodology would be

Discussion and Conclusions

202

recommended, whereas for the more local scenario, the weighting scheme may provide a more

suitable alternative to modifying the methodology with the respective evaluation criteria being

zero-weighted.

Further validation of the methodology should be performed by applying it to a larger number

of CCS options to ensure that no bias towards positive or neutral outcomes was introduced, and

that true negative outcomes exist as the three analysed technologies yielded only neutral and

positive outcomes. Another improvement could be to analyse available projects with statistical

measures to provide harder cut-off criteria for some or all of the evaluation criteria, thereby

addressing some of the issues of potential user bias in the outcome which currently exists.

While statistical metrics were originally intended as part of the rating system, the dispersed

information available for the best documented technology, Geosequestration, and the limited

documentation and lack of case studies for all other technologies in the public domain, made

this approach unfeasible at the current time.

CCS Technology Comparisons

In the following, I discuss the outcome of the comparative analysis of the three CCS

technologies to which the methodology was applied. The technologies were selected, based on a

review of a diverse range of proposed CCS technologies, and it was decided to limit the analysis

to technologies at a relatively advanced stage to ensure at least a minimum amount of public

documentation exists. The overall technology ratings, based on the current status quo as

documented, range from a medium probability of success (2) for Algae CCS, medium-high

(2.3) for Geosequestration, to Biochar CCS with a high probability of success (2.7) as detailed

in Table 6.1. The reason for the difference in rating between Biochar CCS and

Geosequestration, a technology with significant worldwide effort and existing case studies,

becomes more apparent when I contrast the technologies across the evaluation criteria. This will

also highlight why Algae CCS is scoring significantly lower than the other two technologies.

All three analysed technologies share a NIMBY attitude towards the associated infrastructure

developments which are needed, but the association of Geosequestration with the extraction

industry, the additional concerns regarding health and environmental risks, and the potential risk

of lowered property values and damage due to active physical activities like drilling and

injection, negatively impact the attitude about this technology. Biochar CCS and Algae CCS

benefit from their perception as more natural processes with the same biological basis as the

positively-perceived forestation option, while the scientific concepts behind Geosequestration

Discussion and Conclusions

203

are generally not widely known by the public. In addition, the public and political attitude

towards carbon neutral/negative bio energy from renewable biomass is generally more positive.

This all results in a markedly lower Public Acceptance score for Geosequestration when

compared to the other two technologies and makes this one of the major barriers of adoption for

Geosequestration.

For all three technologies, a consistent regulatory framework is lacking, resulting in the

Regulatory Framework score being determined by how well existing regulations can be applied

or amended for these technologies. To a large extent, Algae CCS and Biochar CCS benefit from

coverage of their activities under various other regulations. This is not the case for

Geosequestration, especially when it comes to issues like cross-border transportation and

liability transfer which strongly impact the risk profile of a project. Additional negative

regulations, restricting access to onshore storage sites, further exacerbate this issue. While

government incentives are available for all of the CCS technologies, Biochar and Algae CCS

already benefit from mandates with regards to biofuel and bioenergy usage, while

Geosequestration has currently no consistent mandatory requirement. This has significant

impact on the long term economics of such a project. While a more unified regulatory

framework would benefit all the CCS technologies, Geosequestration is most likely to gain the

most benefit due to the limited versatility of existing regulations. In addition, a carbon price,

mandatory requirement for CO2 emissions reduction, and contract for difference for power

generation would improve economic planning certainty.

The Economics of Algae CCS as well as Geosequestration are strongly impacted by the costs

of the capture process which is not present for the Biochar CCS option as it relies on a natural

process. Algae CCS and Biochar CCS benefit from the ability to offset the costs by multiple

revenue streams, while the only current additional revenue stream for Geosequestration stems

from enhanced hydrocarbon recovery. This makes Geosequestration, especially for industries

other than the petroleum industry, very dependent on government support or a carbon price,

either in the form of direct contributions, or in the form of mandated CO2 capture. The main

challenge for Algae CCS at the current stage is the unpredictability of the revenue stream due to

insufficient research and development (R&D) into the range of factors which impact the

biomass yield and conversion, therefore there are many uncertainties in making any economic

forecasts for a project. Economics is one of the highlights of Biochar CCS, with relatively

modest capital expenditure and operational running costs as well as the availability of a range of

revenue streams, while this represents a significant barrier for the two other technologies.

Discussion and Conclusions

204

Storage Quality is one of the large enablers of Biochar CCS and Geosequestration, even

though the means of sequestration are different. For Geosequestration, a large volume of

geological storage capacity is available worldwide, and large volumes can be sequestered within

a relatively short time span once a project is operational. Biochar CCS benefits from the

applicability to a large range of feedstock together with the possibility to deploy the technology

nearly everywhere and at different scales. Although Biochar CCS projects are not yet

demonstrated at commercial-scales, the potential also exists to achieve significant carbon

mitigation by adoption at the household level around the world. For Algae CCS, storage quality

is a major disadvantage as the CO2 is only temporarily stored and released after combustion of

the biofuel. A solution to this temporary storage could be to combine Algae CCS with Biochar

CCS, using the biomass as a feedstock and thereby achieving a more permanent form of storage.

The Environmental Impact score for Algae CCS is strongly influenced by the significant

areal space and water requirements of the cultivation facilities. While this can be mitigated to

some extent if Algae CCS is combined with waste water treatment or similar activities which

reduce water consumption, it still has a much larger infrastructure footprint than

Geosequestration. The environmental footprint for Geosequestration is generally limited to the

injection site as well as the transportation infrastructure, and is balanced by the positive effect of

reducing anthropogenic CO2 emissions. Biochar CCS has the benefits of removal of

atmospheric CO2 with limited infrastructure footprint, and additionally yields the potential for

soil improvement and water retention.

An analysis of the Science and Technology Sub-Matrix emphasises the observations made

above about the evaluation criteria (Tables 6.2, 6.3 and 6.4). For Geosequestration, the lower

score in comparison to Biochar CCS is not determined by the purely technical sub-criteria, but

by the technology enablers like Regulation, Public Acceptance and Economics. In these

categories, Biochar CCS scores significantly better than Geosequestration, while the other

scores are rather comparable. Algae CCS suffers the most from the additional need for

significant R&D and the lack of any demonstration projects.

As seen from the contrasting of the different technologies, the reason for a lower score of

Geosequestration is not associated with technical challenges, but with factors such as social

licence to operate. While the Economics, at least on the cost of capture can be addressed

through further R&D, factors such as the lack of a Regulatory Framework and relatively low

Public Acceptance will require a wider multi-disciplinary approach. As these two evaluation

criteria are linked, with political will driven by public opinion, it is recommended to address

Public Acceptance by educational programmes and community engagement, placing

Discussion and Conclusions

205

Geosequestration in the context of climate mitigation as an important part of a low-carbon

energy mix strategy but not at the expense of renewables. The use of neutral parties like

scientists or even NGOs for this purpose is recommended as governments and industry are not

perceived as trustworthy sources of information. As Biochar CCS is impacted to a very limited

degree by the above factors, either due to flexibility in current regulations, or because of the

familiarity which the wider public has with the science and technology concepts, it may prove

to have a wider applicability spectrum. The scalability of this technology will allow deployment

virtually anywhere where sustainable feedstock can be found. Algae CCS is the technology

which is the least mature of the analysed technologies, indicated by the absence of any kind of

demonstration projects. Here, further R&D about algae types to optimise biomass production, as

well as cultivation and conversion systems are the highest priorities.

The three CCS technologies evaluated share some similar technology elements and can be

aligned in a synergistic process. My ideal scenario proposal for the most effective and economic

solution to mitigate atmospheric CO2 would be to combine all three together. Firstly, algae

captures the CO2 through photosynthesis using and treating wastewater. During the biorefinery

conversion process, bio-energy and liquid biofuels are produced from the algae biomass, and the

CO2 is captured from the flue gas for reutilisation in algae cultivation, with the excess injected

into suitable geological formations. Finally, the remaining biomass is pyrolised to biochar and

applied to soil thereby gaining the environmental and agricultural benefits. This solution is

unlikely to be available for the next five-ten years.

Table 6.1 Master Matrix – Overall Technology Ratings for Geosequestration, Algae CCS and Biochar CCS.

Table 6.2. Science and Technology Sub-Matrix Scorecard: Results of Application of Methodology to

Geosequestration.

Discussion and Conclusions

206

Table 6.3. Science and Technology Sub-Matrix Scorecard: Results of Application of Methodology to Algae CCS.

Table 6.4. Science and Technology Sub-Matrix Scorecard: Results of Application of Methodology to Biochar CCS.

Example of Carbon Management Portfolio

Based on the application of my methodology to the three CCS options as described in this

and previous chapters, I give some recommendations for an example portfolio for a specific

industry group, LNG operators on the North West Shelf of Australia. Geosequestration would

be the first technology of choice to dispose of the CO2 produced from offshore gas reservoirs

and generated from processing. This industry group has the knowledge and experience in-house

to deploy the technology with all the expertise required for storage characterisation, separation

technologies, injection, and reservoir monitoring. Source and sink matching are not an issue as

the storage space is in the immediate vicinity of the processing facilities, also reducing the costs

for additional transportation infrastructure. With a number of aging oil fields nearby, the

potential to utilise the CO2 for enhanced hydrocarbon recovery exists as well. Recent R&D

highlights new opportunities in using CO2 for unconventional hydrocarbon production. While

this may require additional infrastructure due to the potentially large distances, the possibility to

mitigate the tremendous water requirements for the production of unconventional resources,

with water being a rare resource in North-Western and Northern Australia, could improve the

economics as well as the public acceptance of such a business model.

Discussion and Conclusions

207

Biochar CCS could provide a potential supplemental technology for a carbon management

portfolio. The benefits of this technology are its scalability, affordability and the multiple

revenue streams, including the aligned industry of biofuels. This option may represent a low

risk alternative or backup plan in case challenges occur with the primary storage method of

Geosequestration. Having an alternative to rely upon would allow the LNG plant to maintain

full production if a similar scenario arises to that experienced by Statoil with the Snøhvit

project, where the injectivity for the storage formation was overestimated. I would suggest

operators could implement this technology initially at more modest scales to gain experience,

such as projects in non Organisation for Economic Cooperation and Development (OECD)

nations providing modern pyrolysis stoves and kilns for households to help with food security

through soil amendment. This can generate certified emission reductions (CER) carbon credits

to offset carbon debt, and mitigate GHG based on accumulated volumes. If a more commercial-

scale project (>.500 Mtpa CO2) is desired, a waste to energy plant sited and/or joint venture with

municipal councils could be a good option. The suggested examples could align with an

operator’s “corporate responsibility” or “green” corporate strategy. If the economic benefits are

in line with those predicted in published literature, Biochar CCS may even provide a standalone

business model while at the same time significantly reducing the risk for the LNG operators.

At the current stage, it is not recommended to include Algae CCS in a carbon management

portfolio. Significant R&D is still required and there is no benefit for carbon management in the

short term. For integrated oil companies with refinery and fuel distribution networks, it is

nevertheless recommended to follow up with biofuels research derived from this technology

application, as there may be the potential to supplement current fossil fuels or to address

mandatory biofuel additions.

Discussion and Conclusions

208

Discussion and Conclusions

209

6.2 CONCLUSIONS

I develop a new semi-quantitative methodology for the feasibility assessment of CCS

technologies and their likelihood for wide-scale commercial adoption. The methodology

analyses a wide range of factors from socio-economic to technical, allowing the ranking of

competing or complementary technologies. It is possible to apply this methodology to either a

portfolio of technologies or a portfolio of projects, either on a regional or global scale. The

methodology was applied to three different CCS technologies, Geosequestration, Biochar CCS

and Algae CCS to analyse their current status.

My analysis suggests that Geosequestration has a medium to high probability of success as a

commercial-scale CO2 sequestration option. This technology allows for the fast storage of large

volumes of CO2 in a large range of locations worldwide and this is one of its the main benefits.

For universal commercial scale adoption, regulatory hurdles, as well as the rather negative

public perception of Geosequestration, have still to be overcome. These need to be combined

with an improvement of the cost profile of the capture process, sufficient government funding,

and a mandatory compliance regime to address the current limited commercial revenue streams

for a Geosequestration project. Nevertheless, Geosequestration is one of the technically most

advanced carbon sequestration options as reflected by its technical scores as well as its existing

industrial deployment.

As a third generation biofuel, Algae CCS benefits from supportive regulatory frameworks.

While it offers the potential for a range of possible revenue streams, its capture potential is

determined by the rate of photosynthesis and climate of the region it is implemented in. The

environmental benefits of this technology are mainly linked to the substitution of fossil fuels as

a carbon neutral alternative as the duration of storage is only temporary. A more permanent

solution of storage will require the combination with Biochar CCS which converts the biomass

into a stable form for storage. The significant R&D needed to maximise production efficiencies

and reliability for sustainable volumes, and to lower the capital and operating expenditure of

capture, cultivation and conversion, together with the large land/offshore acreage requirements

for cultivation ponds, result in a longer term but medium probability of success as a

commercial-scale CO2 sequestration option.

Biochar CCS offers the highest probability of success of the three analysed CO2

sequestration options. As with Algae CCS, it benefits from the ability to refer to an existing

regulatory framework, even though it was not specifically designed for this CCS option. The

technical readiness of the technology is comparable to Geosequestration, but the public

Discussion and Conclusions

210

acceptance level is significantly higher. This, and the scalable nature of pyrolysis technology

itself, makes it ready for immediate application at different scales worldwide, as long as

sustainability of the biomass feedstock is kept a priority. Significant global R&D has been

undertaken in this area as governments, scientists and industry recognise its multiple

environmental, agricultural and economic benefits. Further R&D is only needed to improve

conversion efficiencies and optimise agronomical productivity. The demonstration of large scale

projects (>500 Mtpa) would assist in technology uptake. This CO2 mitigation option offers low

risk and is one of the few technologies capable of creating net economic as well as

environmental benefits.

The application of my methodology to these three technologies and the consistency of results

with published literature suggest that the methodology can provide a valuable contribution to

the ranking and prioritisation of CCS projects. The methodology framework is designed to be

flexible to allow for review and modification to allow for technology progress and increased

volumes of, or access to data. The assessment of further technologies can contribute to the

evaluation of the robustness of the methodology and may yield further insight into the various

approaches which can be taken to increase the probability of success for specific CCS

technologies, but also CCS technologies as a whole.

Discussion and Conclusions

211

REFERENCES

Downie, A., Munroe, P., Cowie, A., Van Zwieten, L., and Lau, D. (2012), Biochar as a geoengineering

climate solution: Hazard identification and risk management: Critical reviews in environmental

science and technology 42, 3, 225-250.

European Biofuels Technology Platform/Zero Emission Fossil Fuel Power Plants [EBTP/ZEP]

(2012), Biomass with CO2 capture and storage (Bio-CCS): The way forward for Europe: The

Advisory Council of the European Technology Platform for Zero Emission Fossil Fuel Power Plants

(ZEP) and the Steering Committee of the European Biofuels Technology Platform (EBTP).

http://www.biofuelstp.eu/bio-ccs.html (accessed 2 July 2012).

Global CCS Institute (2012), The Global status of CCS: 2012: Global CCS Institute, Canberra, ACT,

Australia.

International Energy Agency [IEA] (2011), Carbon Capture and Storage Legal and Regulatory

Review: Edition 2: International Energy Agency, OECD [Garrett, J., Beck, B., Lipponen, J.,

McCoy, S., Houssin, D., Volk, D., Yoshimura, T., Finkenrath, M. and Gaghen, R.]: Paris, France.

Intergovernmental Panel on Climate Change [IPCC] (2012), Intergovernmental Panel on Climate

Change Special Report on Renewable Energy Sources and Climate Change Mitigation (SRREN) [O.

Edenhofer, R. Pichs‐Madruga, Y. Sokona, K. Seyboth, P. Matschoss, S. Kadner, T. Zwickel, P.

Eickemeier, G. Hansen, S. Schlömer, C. von Stechow (eds)], Cambridge University Press, New

York, NY, USA.

Kuijper, M. (2011), Public acceptance challenges for onshore CO2 storage in Barendrecht: Energy

Procedia, 4, 6226-6233.

Natural Resources Defense Council [NRDC] (2010), Biochar: Assessing the Promise and Risks to

Guide U.S. Policy, NRDC Issue Paper November 2010, New York, NY, USA.

NNFCC (2012), Research Needs in Ecosystem Services to Support Algal Biofuels,Bioenergy and

Commodity Chemicals Production in the UK: A project for the Algal Bioenergy Special Interest

Group [Smith, N. and Higson, A.]: NNFCC Project Number: 12-008.

Reuters (2011), Update 2 – Vattenfall drops carbon capture project in Germany: Reuters, 5 December

2011. http://www.reuters.com/article/2011/12/05/vattenfall-carbon-idUSL5E7N53PG20111205

(accessed 7 June 2012).

Discussion and Conclusions

212

Schlumberger Business Consulting Energy Institute [SBC Energy Institute] (2012), Leading the

Energy Transition: Bringing Carbon Capture and Storage to Market, [Soupa, O., Lajoie, B., Long,

A., and Alvarez, H.]: Schlumberger Business Consulting Energy Institute.

Schlumberger Business Consulting Energy Institute [SBC Energy Institute] (2013), Leading the

Energy Transition: Carbon Capture and Storage: Bringing Carbon Capture and Storage to Market:

Factbook Edition, January 2013 Update: Schlumberger Business Consulting Energy Institute.

United States Department of Energy [US DOE] (2010), National Algal Biofuels Roadmap: U.S.

Department of Energy, Office of Energy Efficiency and Renewable Energy, Biomass Program.

Waters, D., Zwieten, L., Singh, B. P., Downie, A., Cowie, A. L., and Lehmann, J. (2011), Biochar in

soil for climate change mitigation and adaptation. In Soil Health and Climate Change: Springer

Berlin Heidelberg, 15, 345-368.