0% found this document useful (0 votes)
49 views489 pages

OIW Report 1

Uploaded by

nuvan
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
49 views489 pages

OIW Report 1

Uploaded by

nuvan
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 489

The Royal Society of Canada Expert Panel:

Report

The Behaviour and Environmental Impacts of


Crude Oil Released into Aqueous Environments
November 2015

Dr. Kenneth Lee (Chair)


Dr. Michel Boufadel
Dr. Bing Chen
Dr. Julia Foght
Dr. Peter Hodson
Dr. Stella Swanson
Dr. Albert Venosa
Behaviour and Environmental Impacts of Crude Oil
Released into Aqueous Environments

An Expert Panel Report prepared at the request of


the Royal Society of Canada
for the Canadian Energy Pipeline Association
and the Canadian Association of Petroleum Producers

The Royal Society of Canada


La Société royale du Canada

November 2015

ISBN: 978-1-928140-02-3

Walter House
282 Somerset West,
Ottawa ON K2P 0J6, Canada

Telephone: +1 (613) 991-6990


Facsimile: +1 (613) 991-6996
Website: http://www.rsc-src.ca

RSC: EPR 15-1

The report should be cited as follows:

Lee, Kenneth (chair), Michel Boufadel, Bing Chen, Julia Foght, Peter Hodson, Stella Swanson,
Albert Venosa. (2015). Expert Panel Report on the Behaviour and Environmental Impacts of Crude
Oil Released into Aqueous Environments. Royal Society of Canada, Ottawa, ON. ISBN: 978-1-
928140-02-3

The opinions expressed in this report are those of the authors and do not necessarily represent
those of the Royal Society of Canada or the opinion or policy of the Canadian Energy Pipeline
Association or the Canadian Association of Petroleum Producers.
The Royal Society of Canada Expert Panels

The Expert Panels of the Royal Society of Canada (RSC) are central to the society since its mission
includes a charge to “…advance knowledge, encourage integrated interdisciplinary understandings and
address issues that are critical to Canadians”.

RSC Expert Panels provide independent, timely and authoritative insights and advice to governments,
industry, non-governmental organizations, and citizens. Topics for expert panels are selected, in part, on
the benefit they are likely to receive from a critical assessment of existing knowledge from a wide range
of disciplinary and sectorial perspectives. While the Society does not shy away from controversial topics,
it does insist that there must be room for critical, balanced, scientific analysis that makes it possible to
achieve a consensus document from a diverse group of experts. Topics where differences in opinion are
based on belief systems that are not founded in informed science are not be appropriate for an RSC Expert
Panel.

Not surprisingly, RSC Expert Panels are often commissioned (and funded) by organizations that may
have a strong interest in the conclusions of the report. However, for a variety of reasons, they not only
want to receive a report that is independent, balanced and objective, but they want that report to be seen to
be so by other stakeholders.

To ensure that this is the case, the RSC Committee on Expert Panels has developed a rigorous set of
procedures1 that guide their work. These procedures include:

1. Defining the Terms of Reference (TOR): Other than writing a cheque, this is the only part of
the Expert Panel process where the commissioning organization has significant input. The RSC
‘Committee on Expert Panels’ works with the funder to develop draft TOR describing the charge
to the panel, ensuring that it is neither too narrow, nor too broad, and, in particular, that it does
not direct the panel towards a pre-determined conclusion. Before the TOR are approved, they are
discussed with experts from the relevant stakeholder communities and, if appropriate,
modifications are made before final approval by the funder and the RSC.
2. Selection of Panel members. Panel selection is the responsibility of the Committee on Expert
Panels and its Oversight Committee, and does not consult with the commissioning organization.
Explicit effort is made to ensure diversity of both expertise and perspectives (bias) in the subject
area. Prospective panelists are interviewed and complete ‘conflict of interest declarations’ before
being selected. Panelists are not reimbursed for the considerable time and effort that they
contribute to the work of RSC Expert Panels.
3. Deliberations of the Panel and Report Preparation. Panel deliberations are ‘in camera’, with
the exception of planned public hearings that are open to all. The panelists decide on the content
of their submitted report and write it themselves. They are identified as the authors of the report
(not the RSC), since RSC Expert Panels are reports to the RSC not from the RSC to the sponsor.
4. Selection of Reviewers. Selection of a Peer Review Monitor and the Peer Reviewers is also the
responsibility of the Committee on Expert Panels and its Oversight Committee. The monitor
manages the Peer Review process and is selected to ensure they have sufficient expertise to be
comfortable in the content area, but have no conflict of interest or perception of bias. The peer
review monitor works with the Panel Chair to ensure that all feedback from reviewers is carefully
assessed and that appropriate revisions are made.
5. Public Release of the Report. The report is released to the media and public (on the RSC
website2). No opportunity is given to the commissioning organization to request modifications of
the report or its recommendations. However, the commissioning organization does receive a copy

1
See http://www.rsc.ca/en/expert-panels/information-about-expert-panels
2
http://www.rsc.ca/en/expert-panels/rsc-reports

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca |2
of the report a few days before it is released so they have an opportunity to prepare responses to
its content by the time the report is publically released.
6. Ownership of the report. The report belongs to the RSC. However, the RSC typically gives
written permission to the panelists / authors to rework the document and submit portions of it to
academic journals or other publications if they wish to do so.

Organizations or individuals interested in discussing opportunities for the RSC to carry out an Expert
Panel on a topic of interest to them should contact the Chair of the Committee on Expert Panels at
[email protected].

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca |3
Preface

Canadians are among the highest per capita consumers of oil in the world, equivalent to about 25 barrels –
or about 4000 litres – of oil per person per year.

Canada is also one of the world’s only developed economies that produces more oil than it demands for
domestic consumption, currently about 1.5 times more (about 6000 L/person/yr). In recent years, the rate
of oil production has been increasing by about 5% per year, primarily through the growth of Alberta’s oil
sands.

To meet both domestic and international demand, the oil must be moved by pipeline, rail, tanker trucks or
ship from where it is recovered to where it will be refined and ultimately used. In recent years, the energy
industries have been developing plans and seeking permission to expand or modify their transportation
networks, especially those involving pipelines. Keystone XL, Northern Gateway, Trans-Mountain, Line 9
Reversal and Canada East are names of pipeline projects that have regularly been a regular feature in the
news.

Hot-button topics at the forefront of those concerned with these pipeline projects include questions about
the consequences of accidental spills of crude oil into aquatic ecosystems. Do we have sufficient
knowledge about how crude oils behave when released into fresh waters, estuaries or oceans to develop
effective strategies for spill preparedness, spill response and remediation? What are the gaps in
knowledge and how should research insights inform policies, regulations and practices in this area?

Among the conditions for the approval of the Gateway project, the National Energy Board required the
applicants to establish ‘a research program regarding the behavior and cleanup (including recovery) of
heavy oils spilled in freshwater and marine aquatic environments’ (Condition 169).3 In mid 2014, the
Royal Society of Canada’s (RSCs) Committee on Expert Panels was approached about carrying out an
independent, arm’s length assessment of the state of the science in this area.

Since the prospective funders of this panel (Canadian Energy Pipeline Association (CEPA) and the
Canadian Association of Petroleum Producers (CAPP)) might be perceived to have a strong interest in the
outcome of the work, the Committee on Expert Panels was especially rigorous in following its procedures
(see previous pages) to ensure an independent, balanced and objective report.

CEPA and CAPP suggested the initial Terms of Reference, and were receptive to changes in those terms
following RSC preliminary consultation with diverse experts. Neither CEPA nor CAPP (or their
membership) had any input into the selection of panelists or reviewers, or into the recommendations of
this panel.

The RSC thanks CEPA and CAPP for their support of this project and appreciates their professional
approach that let us go about our work without interference of any kind.

We wish to thank the Panel Chair, Dr. Ken Lee and his fellow panelists (listed on the next page) for
volunteering their time and expertise to prepare this report. This comprehensive and high quality report
did not emerge without exceptional effort. Thank you!

We also want to thank the Peer Review Monitor, Dr. Robie Macdonald, FRSC and eight expert reviewers
(listed on the next page), who provided extensive comments, criticisms and suggestions on the first draft
of this report. The additions and clarifications that resulted from their feedback significantly added to the
quality of the document.
3
Report of the Joint Review Panel for the Enbridge Northern Gateway Project (Dec 19, 2013) Conditions 169, 170, and 193
http://gatewaypanel.review-examen.gc.ca/clf-nsi/hm-eng.html

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca |4
Finally, a special thank you to Russel MacDonald from the RSC office in Ottawa and Lisa Isaacman from
Bonnyville who provided administrative and editorial support for the work of the panel and to the
members of the Committee on Expert Panels and its Oversight Committee that were involved in selecting
the panelists and reviewers.

David B Layzell, PhD, FRSC


Chair, Committee on Expert Panels

Graham Bell, PhD, FRSC


President, Royal Society of Canada

RSC Committee on Expert Panels (2014-15)

David B. Layzell, PhD, FRSC. (Chair) Professor and Director, Canadian Energy Systems Analysis
Research (CESAR) Initiative, University of Calgary, Calgary, Alberta

Christl Verduyn, PhD, FRSC (Academy I Representative) Professor of English Literature and
Canadian Studies, Mount Allison University, Sackville, New Brunswick

John Myles, PhD, FRSC (Academy II Representative) Emeritus Professor, Sociology and Senior
Fellow in the School of Public Policy, University of Toronto, Toronto, Ontario

Sarah P. Otto, PhD, FRSC (Academy III Representative) Professor, Department of Biology,
University of British Columbia, Vancouver, British Columbia

John P. Smol, OC, PhD, FRSC (Member-at-Large) Canada Research Chair in Environmental
Change, Department of Biology, Queen’s University, Kingston, Ontario

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca |5
Panel Members

Dr. Kenneth Lee (Chair) is the Director of Oceans and Atmosphere, Commonwealth Scientific and
Industrial Research Organisation (CSIRO) in Perth, Australia. This research group provides large-
scale multidisciplinary science to support the management and sustainable development of
Australia’s marine and atmospheric resources. The founding Executive Director of the Centre for
Offshore Oil, Gas and Energy Research (COOGER), Department of Fisheries and Oceans, Canada,
he received the “Prix d’Excellence” for exemplary contributions in science (Fisheries and Oceans
Canada) and the Government of Canada’s “Federal Partners in Technology Transfer - Leadership
Award” for technology transfer from a federal laboratory to the private sector.

Dr. Michel Boufadel is Professor of Environmental Engineering and Director of the Center for Natural
Resources Development and Protection at the New Jersey Institute of Technology. He is a
Professional Engineer in Pennsylvania and New Jersey, a Board Certified Professional Engineer, and
a Fellow in the American Society of Civil Engineers.

Dr. Bing Chen is the leader of the Northern Region Persistent Organic Pollution Control (NRPOP)
Laboratory, Associate Professor of Civil Engineering with the Faculty of Engineering and Applied
Science, and Chair of the Environmental Systems Engineering & Management Program at the
Memorial University of Newfoundland, St. John’s, NL, Canada.

Dr. Julia Foght is Professor Emerita at the University of Alberta, where she was a Professor of
Petroleum Microbiology in the Department of Biological Sciences from 1994 to 2014. Julia received
the Petro-Canada Young Innovators Award in 2001, a McCalla Professorship in 2011 from the
University of Alberta, and in 2014 the Alberta Science & Technology Foundation (ASTech) award
in Innovation in Oil Sands Research.

Dr. Peter Hodson.is an ecotoxicologist and Professor Emeritus with Biology and the School of
Environmental Studies at Queen’s University, and a Past President of the Society of Environmental
Toxicology and Chemistry. He began his career with Fisheries and Oceans and his research
contributed to the development of Great Lakes Water Quality Objectives under the 1972 Canada-US
Great Lakes Water Quality Agreement and to the development of Federal Environmental Effects
Monitoring Programs for the pulp and paper sector. Since 1995, he has focused on the toxicity of
petroleum hydrocarbons to fish embryonic development and the role of contaminants in the decline
of American eel abundance in Lake Ontario.

Dr. Stella Swanson is an aquatic ecologist and risk assessment specialist with Swanson Environmental
Strategies that she operates out of Calgary, AB and Fernie, BC. Her 30-year career has included
management of the Aquatic Biology Group at the Saskatchewan Research Council, and consulting
positions with SENTAR Consultants (now Stantec) and Golder Associates Ltd. (where she attained
the position of Principal).

Dr. Albert Venosa is the former Director of the Land Remediation and Pollution Control Division
(LRPCD), one of the five divisions within the National Risk Management Research Laboratory
(NRMRL), Cincinnati, OH, in the United States Environmental Protection Agency's (U.S. EPA's)
Office of Research and Development. As an environmental scientist and microbiologist, he managed
EPA's oil spill research program from 1990 until 2014, during which time he became an expert in
bioremediation and dispersant research and other response technologies. Prior to that, he led U.S.
EPA's national program on wastewater disinfection research for 13 years. He is now retired after 45
years of government service.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca |6
Support of the Work of the Panel

We are grateful to the following individuals:

Christopher Stone, (Calgary, Alberta) for his assistance in compiling references and identifying
prospective panellists and reviewers.

Lisa Isaacman, (Bonnyville, Alberta) for her administrative work with the panel and editorial work on
the report

Russel MacDonald, (Officer, Expert Panels & the College, the Royal Society of Canada, Ottawa,
Ontario) for his assistance in managing the funds, helping with logistics, taking notes and producing
the final report.

Peter Christie (Science Writer, Kingston, Ontario) for his assistance with some critical sections of the
report.

Peer Reviewers

The Royal Society of Canada would like to acknowledge and sincerely thank the Peer Review Monitor
and eight peer reviewers who provided valuable, detailed input into the first draft of this report. The
following reviewers have permitted the RSC to release their names:

Dr. Robie Macdonald, FRSC, (Peer Review Monitor) Fisheries and Oceans Canada, Victoria, British
Columbia

Dr. Ronald Atlas, Department of Biology, University of Louisville, Louisville, Kentucky

Dr. Tracy Collier, Science Director, Puget Sound Partnership, Bainbridge Island, Washington

Dr. Merv Fingas, Independent Consultant, Edmonton, Alberta

Dr. Jeffrey Hutchings, FRSC, Department of Biology, Dalhousie University, Nova Scotia

Dr. Don Mackay, OC, Department of Chemistry, Trent University, Peterborough, Ontario

Dr. Jacqueline Michel, Research Planning, Inc., Columbia, South Carolina

Dr. Jeffrey Short, JWS Consulting, Juneau, Alaska

The peer reviewers provided many very constructive comments, but they were not asked to endorse the
conclusions or recommendations, nor did they see the final draft of the report before it was released. The
identification of reviewers is done to recognize them for their valuable contributions and does not imply
that they endorse the report or agree with its content. Responsibility for the final content of this report
rests entirely with the Panel.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca |7
TABLE OF CONTENTS

LIST OF ACRONYMS ................................................................................................................................... 17


GLOSSARY OF TERMS ................................................................................................................................ 20

EXECUTIVE SUMMARY .............................................................................................................................. 26

CHAPTER 1: INTRODUCTION ..................................................................................................................... 38


1.1 Sponsors and Mandate of the Panel ............................................................................................ 38
1.2 Scope of the Panel’s Review ......................................................................................................... 39
1.3 Consultation with Stakeholders ................................................................................................... 39
1.4 Context for Oil Spills in Canada .................................................................................................. 40
1.4.1 Spills from Oil Tankers in Marine Waters .................................................................................. 40
1.4.2 Accidental Releases from Offshore Oil and Gas Exploration and Production ........................... 43
1.4.3 Inland Freshwater Spills ............................................................................................................. 45
1.4.4 Spills from Pipelines ................................................................................................................... 46
1.4.5 Other Releases from Production Sources into Fresh Water ....................................................... 47
1.4.6 Spills Due to Train Derailments ................................................................................................. 48
1.4.7 Spills Due to Truck Transport Accidents .................................................................................... 49
1.4.8 Summary of Crude Oil Spills in Canada ..................................................................................... 50
1.5 Organization of This Report ........................................................................................................ 51

CHAPTER 2: CHEMICAL COMPOSITION, PROPERTIES AND BEHAVIOUR OF SPILLED OILS ................. 53


2.1 Chemical Composition of Oils ..................................................................................................... 54
2.1.1 Four Major Chemical Fractions of Crude Oil ............................................................................ 54
2.1.1.1 Saturates ................................................................................................................................. 54
2.1.1.2 Aromatics ................................................................................................................................ 60
2.1.1.3 Resins ...................................................................................................................................... 61
2.1.1.4 Asphaltenes ............................................................................................................................. 61
2.1.2 Minor Components of Crude Oils ............................................................................................... 62
2.2 Bulk Properties of Oil ................................................................................................................... 64
2.2.1 Density, Specific Gravity and API Gravity ................................................................................. 64
2.2.2 Viscosity and Pour Point............................................................................................................. 65
2.2.3 Volatility and Flash Point ........................................................................................................... 66
2.2.4 Water Solubility .......................................................................................................................... 67
2.2.5 Surface Tension and Interfacial Tension .................................................................................... 67
2.2.6 Adhesion ...................................................................................................................................... 68
2.3 Examples of Oil Types .................................................................................................................. 68
2.3.1 Condensates and Refined Products Used as Diluents for Bitumen ............................................ 71
2.3.2 Light Crude Oils ......................................................................................................................... 71
2.3.3 Medium Crude Oils ..................................................................................................................... 72
2.3.4 Shale Oils (Tight Oil or Light Tight Oil) .................................................................................... 72
2.3.5 Waxy Crude Oils ......................................................................................................................... 72
2.3.6 Sour Crude Oils .......................................................................................................................... 72
2.3.7 Heavy Oils, Bitumen and Diluted Bitumen Products .................................................................. 73
2.4 Weathering of Oil Spilled in Aquatic Environments ................................................................. 75
2.4.1 Weathering Processes at the Water Surface ............................................................................... 78
2.4.1.1 Spreading ................................................................................................................................ 78
2.4.1.2 Evaporation, Aerosolization and Atmospheric Re-deposition ................................................ 79
2.4.1.3 Photooxidation ........................................................................................................................ 82
2.4.1.4 Emulsification ......................................................................................................................... 83

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca |8
2.4.2 Weathering Processes in the Water Column ............................................................................... 85
2.4.2.1 Dissolution .............................................................................................................................. 85
2.4.2.2 Natural dispersion .................................................................................................................. 86
2.4.2.3 Submergence, sinking, sedimentation and re-suspension of oil .............................................. 87
2.4.2.4 Formation of tar balls, tar patties and tar mats...................................................................... 93
2.4.2.5 Biodegradation ....................................................................................................................... 93
2.4.3 Oil Interactions with Shorelines and Sediments ......................................................................... 94
2.4.3.1 Sorption, penetration and sequestration ................................................................................. 95
2.4.4 Perspective .................................................................................................................................. 96
2.4.5 Oil Interactions with Ice ............................................................................................................. 96
2.5 Summary of Identified Research Needs and Recommendations .............................................. 99
2.5.1 Research Needs ........................................................................................................................... 99
2.5.1.1 Short Term (High Priority) ..................................................................................................... 99
2.5.1.2 Medium Term .......................................................................................................................... 99
2.5.1.3 Long Term ............................................................................................................................. 100
2.5.2 Operational Preparedness Needs ............................................................................................. 101

CHAPTER 3: EFFECT OF ENVIRONMENT ON THE FATE AND BEHAVIOUR OF OIL .............................. 102
3.1 Key Environmental Factors Affecting Spilled Oil Fate and Behaviour ................................. 103
3.1.1 Presence of microbial species in oil-impacted water, sediments, shorelines and wetlands ..... 103
3.1.2 Temperature .............................................................................................................................. 107
3.1.2.1 Physical-Chemical Properties of the Oil Influenced by Temperature .................................. 107
3.1.2.2 Effect of Temperature on Biodegradation............................................................................. 107
3.1.3 Dissolved Oxygen...................................................................................................................... 109
3.1.4 Nutrient supply and stoichiometric ratios of C:N:P ................................................................. 110
3.1.5 Salinity ...................................................................................................................................... 112
3.1.6 pH.............................................................................................................................................. 112
3.2 Oil Impacts on Different Aquatic Environments ..................................................................... 113
3.2.1 Sand and gravel shorelines ....................................................................................................... 113
3.2.1.1 Porosity of the substrate ....................................................................................................... 114
3.2.1.2 Shoreline morphology ........................................................................................................... 114
3.2.1.3 Wave exposure ...................................................................................................................... 114
3.2.2 Estuarine shorelines .................................................................................................................. 115
3.2.3 Riverine shorelines .................................................................................................................... 115
3.2.4 Rocky cliffs and beaches ........................................................................................................... 116
3.2.5 Lakes ......................................................................................................................................... 116
3.2.6 Wetlands .................................................................................................................................... 117
3.2.7 Permafrost................................................................................................................................. 118
3.2.8 Ice-covered Arctic marine environments .................................................................................. 119
3.2.9 Offshore Surface Spills at Sea and in Large Freshwater Lakes (e.g., Great Lakes). ............... 121
3.2.9.1 Spreading .............................................................................................................................. 121
3.2.9.2 Evaporation........................................................................................................................... 121
3.2.9.3 Dissolution ............................................................................................................................ 122
3.2.9.4 Emulsification ....................................................................................................................... 122
3.2.9.5 Photooxidation ...................................................................................................................... 123
3.2.9.6 Sedimentation/sinking ........................................................................................................... 123
3.3 Importance of Oil Type in Relation to the Environment ........................................................ 124
3.3.1 Light oils (diesel, heating oil, light crudes) .............................................................................. 124
3.3.2 Medium crude oils (most crude oils) ......................................................................................... 124
3.3.3 Heavy oils.................................................................................................................................. 124
3.4 Summary of Research Needs and Recommendations.............................................................. 125

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca |9
3.4.1 High Priority Research Needs .................................................................................................. 125
3.4.1.1 Arctic Ecosystems ................................................................................................................. 125
3.4.2 Medium Priority Research Needs ............................................................................................. 125
3.4.2.1 Persistence of Oil in Non-Marine Environments .................................................................. 125
3.4.2.2 High Risk Characteristics and Treatability of Dilbit ............................................................ 126
3.4.3 Long Term Research Needs ...................................................................................................... 126
3.4.3.1 Oil-Suspended Particulate Matter Aggregates and Dilbit Emulsification ............................ 126

CHAPTER 4: OIL TOXICITY AND ECOLOGICAL EFFECTS ..................................................................... 127


4.1 Effects of Oil on Aquatic Organisms ......................................................................................... 131
4.1.1 Immediate effects of an oil spill ................................................................................................ 131
4.1.2 Plants ........................................................................................................................................ 132
4.1.3 Invertebrates ............................................................................................................................. 134
4.1.4 Fish ........................................................................................................................................... 135
4.1.4.1 Acute toxicity ......................................................................................................................... 135
4.1.4.2 Chronic and sublethal toxicity .............................................................................................. 138
4.1.4.3 Hydrocarbons causing chronic toxicity ................................................................................ 143
4.1.4.4 Physiological and pathological responses ............................................................................ 145
4.1.4.5 Gene mutations and cancer................................................................................................... 146
4.1.4.6 Population and ecosystem consequences of oil toxicity ........................................................ 147
4.1.5 Reptiles and amphibians ........................................................................................................... 150
4.1.6 Birds .......................................................................................................................................... 151
4.1.7 Mammals ................................................................................................................................... 153
4.1.8 Ecosystem services .................................................................................................................... 155
4.2 Factors Affecting Oil Toxicity to Fish ....................................................................................... 156
4.2.1 Species tested ............................................................................................................................ 156
4.2.2 Type of oil and oil characteristics ............................................................................................. 159
4.2.3 Chemically- vs. mechanically-dispersed oil .............................................................................. 161
4.2.4 Sediment contamination ............................................................................................................ 164
4.2.5 Photo-enhanced toxicity ............................................................................................................ 164
4.2.6 Arctic vs. temperate spills ......................................................................................................... 165
4.2.7 Marine vs. freshwater spills ...................................................................................................... 167
4.2.8 Interactive and cumulative effects............................................................................................. 170
4.3 Limitations of Current Toxicity Test Methods with Oil .......................................................... 171
4.3.1 Constant vs. time-varying exposure concentrations ................................................................. 171
4.3.2 Exposure time............................................................................................................................ 172
4.3.3 Measurements of oil concentrations in toxicity tests ................................................................ 173
4.3.4 Confounding of risk assessments by oil droplets ...................................................................... 174
4.3.5 Realism in toxicity tests ............................................................................................................. 176
4.4 Research Needs and Recommendations .................................................................................... 177
4.4.1 Research Needs ......................................................................................................................... 177
4.4.2 Operational/preparedness Needs .............................................................................................. 178

CHAPTER 5: MODELING OIL SPILLS IN WATER .................................................................................... 179


5.1 Modeling of Oil Weathering....................................................................................................... 179
5.1.1 Oil evaporation ......................................................................................................................... 180
5.1.2 Oil dissolution ........................................................................................................................... 181
5.1.3 Photooxidation .......................................................................................................................... 182
5.1.4 Oil emulsification ...................................................................................................................... 182
5.1.5 Oil dispersion ............................................................................................................................ 183
5.1.6 Oil particle aggregation............................................................................................................ 184

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 10
5.1.7 Oil biodegradation in open water ............................................................................................. 186
5.1.8 Oil biodegradation within beaches ........................................................................................... 186
5.2 Oil Transport ............................................................................................................................... 187
5.2.1 Spreading .................................................................................................................................. 187
5.2.2 Transport in lakes and sea ........................................................................................................ 188
5.2.3 Transport in streams and rivers ................................................................................................ 188
5.2.4 Transport in estuaries ............................................................................................................... 189
5.2.5 Transport of oil within beaches ................................................................................................ 190
5.2.6 Floodplains ............................................................................................................................... 191
5.2.7 Wetlands .................................................................................................................................... 191
5.2.8 Underwater release ................................................................................................................... 191
5.2.9 Ice .............................................................................................................................................. 192
5.3 Research Recommendations in Priority Order ........................................................................ 193

CHAPTER 6: REVIEW OF SPILL RESPONSE OPTIONS ............................................................................ 194


6.1 Natural Processes ........................................................................................................................ 196
6.1.1 Monitored Natural Attenuation (MNA)..................................................................................... 196
6.1.2 Evaporation............................................................................................................................... 196
6.1.3 Photooxidation .......................................................................................................................... 196
6.2 Physical Response Methods........................................................................................................ 197
6.2.1 Containment and recovery ........................................................................................................ 197
6.2.2 Sorbents..................................................................................................................................... 200
6.2.2.1 Selection of sorbents for floating oil spills on water............................................................. 201
6.2.3 Oil removal from shorelines ..................................................................................................... 202
6.2.3.1 Sand beaches ......................................................................................................................... 202
6.2.3.2 Pebble or shingle beaches..................................................................................................... 202
6.2.3.3 Cobble beaches ..................................................................................................................... 202
6.2.3.4 Boulder beaches .................................................................................................................... 203
6.2.3.5 Bedrock shoreline with no beach .......................................................................................... 203
6.2.3.6 Low pressure washing or sediment flushing of a variety of beach types .............................. 204
6.2.4 Vegetation cutting ..................................................................................................................... 205
6.2.5 Removal of oiled sediment ........................................................................................................ 206
6.2.6 Oiled sediment reworking ......................................................................................................... 207
6.2.7 Physical dispersion (Oil mineral aggregates) .......................................................................... 208
6.2.8 In Situ burning .......................................................................................................................... 208
6.3 Biological and Chemical Methods ............................................................................................. 210
6.3.1 Bioremediation .......................................................................................................................... 210
6.3.1.1 Wetlands ................................................................................................................................ 212
6.3.2 Phytoremediation ...................................................................................................................... 213
6.3.3 Chemical dispersion .................................................................................................................. 218
6.3.4 Surface washing by chemical washing agents .......................................................................... 224
6.3.5 Solidifiers .................................................................................................................................. 224
6.3.6 Herding agents .......................................................................................................................... 225
6.3.7 Debris and detritus removal ..................................................................................................... 226
6.4 Factors Affecting Spill Response and Cleanup Effectiveness ................................................. 227
6.4.1 Oil types and properties ............................................................................................................ 227
6.4.2 Environmental and ecological factors ...................................................................................... 228
6.4.2.1 Weather ................................................................................................................................. 228
6.4.2.2 Wave height ........................................................................................................................... 229
6.4.2.3 Ice conditions ........................................................................................................................ 230
6.4.2.4 Available daylight and visibility ........................................................................................... 231

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 11
6.4.2.5 Ecological considerations ..................................................................................................... 232
6.4.3 Technical and economic factors ............................................................................................... 232
6.4.4 Other factors ............................................................................................................................. 233
6.5 Research Needs and Recommendations .................................................................................... 234
6.5.1 High Priority Research Needs .................................................................................................. 234
6.5.1.1 Cold, harsh environments and controlled field studies ......................................................... 234
6.5.1.2 Sediments .............................................................................................................................. 234
6.5.1.3 Dispersants ........................................................................................................................... 234
6.5.2 Medium Priority Research Needs ............................................................................................. 234
6.5.2.1 Bioremediation, including phytoremediation and wetlands ................................................. 234
6.5.2.2 Herding Agents and In Situ Burning (ISB) ........................................................................... 235
6.5.3 Low Priority Research Needs ................................................................................................... 235
6.5.3.1 Berm Relocation, surf washing, sediment reworking ............................................................ 235
6.5.3.2 Sorbents, surface washing agents (SWA), solidifiers ............................................................ 235

CHAPTER 7: PREVENTION AND RESPONSE DECISION MAKING ........................................................... 236


7.1 Prevention and Preparedness .................................................................................................... 236
7.1.1 Prevention ................................................................................................................................. 237
7.1.1.1 Tanker safety and spill prevention ........................................................................................ 237
7.1.1.2 Pipeline safety and spill prevention ...................................................................................... 239
7.1.1.3 Oil transport by rail: safety and spill prevention.................................................................. 242
7.1.1.4 Oil transport by truck: safety and spill prevention ............................................................... 244
7.1.2 Oil spill monitoring ................................................................................................................... 244
7.1.2.1 Onsite monitoring ................................................................................................................. 244
7.1.2.2 Remote sensing...................................................................................................................... 245
7.1.2.3 Shoreline monitoring............................................................................................................. 246
7.1.3 Oil and sample analysis ............................................................................................................ 247
7.1.3.1 Physical analysis ................................................................................................................... 247
7.1.3.2 Chemical analysis ................................................................................................................. 247
7.1.3.3 Biological monitoring ........................................................................................................... 247
7.1.4 Baseline data ............................................................................................................................. 248
7.1.5 Sensitivity and vulnerability ...................................................................................................... 249
7.1.5.1 Sensitivity .............................................................................................................................. 249
7.1.5.2 Vulnerability ......................................................................................................................... 250
7.1.5.3 Vulnerability analysis ........................................................................................................... 250
7.1.6 Preparedness and contingency planning .................................................................................. 251
7.1.6.1 Preparedness and response regime ...................................................................................... 251
7.1.6.2 Technical preparation ........................................................................................................... 252
7.1.6.3 Contingency planning ........................................................................................................... 253
7.1.6.4 Legal requirements and assessment ...................................................................................... 253
7.1.6.5 Training, drills and education .............................................................................................. 254
7.2 Response Decision Support ........................................................................................................ 254
7.2.1 Early warning ........................................................................................................................... 255
7.2.1.1 Spill imaging and analysis .................................................................................................... 255
7.2.1.2 Early warning indicators ...................................................................................................... 255
7.2.2 Response technology screening and evaluation ........................................................................ 257
7.2.2.1 Technology standards ........................................................................................................... 257
7.2.2.2 Screening and performance evaluation ................................................................................ 258
7.2.3 Spill physical and numerical simulation ................................................................................... 258
7.2.4 Net Environmental Benefit Analysis (NEBA) ............................................................................ 258
7.2.5 Performance evaluation and post-event management .............................................................. 260

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 12
7.2.5.1 Performance evaluation and impact assessment .................................................................. 260
7.2.5.2 Adaptive management ........................................................................................................... 261
7.2.6 Risk assessment ......................................................................................................................... 265
7.2.6.1 Spill risk assessment.............................................................................................................. 265
7.2.6.2 Environmental/ecological risk assessment ........................................................................... 268
7.2.6.3 Health risk assessment .......................................................................................................... 269
7.2.7 Cleanup process simulation and control................................................................................... 270
7.2.7.1 Gaps in cleanup process modeling ....................................................................................... 270
7.2.7.2 Special considerations for harsh environments .................................................................... 271
7.2.8 Response operation optimization .............................................................................................. 271
7.2.9 Information technologies for response decision making .......................................................... 272
7.2.9.1 Geomatics, remote sensing, and geographical information systems .................................... 272
7.2.9.2 Data mining and artificial intelligence ................................................................................. 272
7.2.9.3 Results visualization.............................................................................................................. 273
7.3 Summary ...................................................................................................................................... 273
7.4 Research Needs and Recommendations .................................................................................... 274
7.4.1 High Priority Research Needs .................................................................................................. 274
7.4.2 Medium and low priority research needs.................................................................................. 274
7.4.3 Additional recommendations for response operations and preparedness ................................ 275

CHAPTER 8: RISKS FROM OIL SPILLS .................................................................................................... 276


8.1 The Arrow Spill ........................................................................................................................... 278
8.1.1 Important factors affecting the consequences of the Arrow spill .............................................. 281
8.1.1.1 Oil properties and behaviour ................................................................................................ 281
8.1.1.2 Effect of the environment on fate and behaviour .................................................................. 281
8.1.1.3 Oil toxicity ............................................................................................................................. 281
8.1.1.4 Spill response ........................................................................................................................ 281
8.1.2 Lessons learned from the Arrow spill ....................................................................................... 282
8.1.3 Research needs related to the Arrow spill ................................................................................ 282
8.2 The Baffin Island Oil Spill Project (BIOS) ............................................................................... 282
8.2.1 The near-shore study................................................................................................................. 283
8.2.2 The shoreline study ................................................................................................................... 284
8.2.3 Important factors affecting the consequences of the experimental BIOS spills ........................ 286
8.2.3.1 Oil properties and behaviour ................................................................................................ 286
8.2.3.2 Effect of the environment on fate and behaviour .................................................................. 286
8.2.3.3 Oil toxicity ............................................................................................................................. 286
8.2.3.4 Spill response ........................................................................................................................ 286
8.2.4 Lessons learned from the BIOS project .................................................................................... 287
8.2.5 Research needs based on the BIOS project............................................................................... 287
8.3 Other Arctic Oil Spill Field Trials ............................................................................................. 288
8.3.1 Review of conclusions arising from other experimental spills in ice covered waters ............... 288
8.3.2 Research needs regarding spills in Arctic conditions ............................................................... 289
8.4 The Exxon Valdez Oil Spill (EVOS) .......................................................................................... 290
8.4.1 Important factors affecting the consequences of the Exxon Valdez spill .................................. 295
8.4.1.1 Oil properties and behaviour ................................................................................................ 295
8.4.1.2 Effect of the environment on fate and behaviour .................................................................. 296
8.4.1.3 Oil toxicity ............................................................................................................................. 296
8.4.1.4 Spill response ........................................................................................................................ 296
8.4.2 Lessons learned from the Exxon Valdez spill ............................................................................ 296
8.4.3 Research needs related to the Exxon Valdez spill, in the Canadian context............................. 297
8.5 Deepwater Horizon (DWH) Blowout ........................................................................................ 297

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 13
8.5.1 Dispersant use ........................................................................................................................... 300
8.5.2 Deposition and effects on shorelines ........................................................................................ 301
8.5.3 Effects on fisheries .................................................................................................................... 302
8.5.4 Effects on invertebrates ............................................................................................................. 303
8.5.5 Effects on trophic level dynamics.............................................................................................. 303
8.5.6 Effects on ecosystem services .................................................................................................... 304
8.5.7 Important factors affecting the consequences of the Deepwater Horizon Blowout .................. 306
8.5.7.1 Oil properties and behaviour ................................................................................................ 306
8.5.7.2 Effect of the environment on fate and behaviour .................................................................. 306
8.5.7.3 Oil toxicity ............................................................................................................................. 307
8.5.7.4 Spill response ........................................................................................................................ 307
8.5.8 Lessons learned from the Deepwater Horizon Blowout ........................................................... 307
8.5.8.1 Lessons learned according to the USGS (2015) ................................................................... 307
8.5.8.2 Other lessons learned (from the RSC Panel’s review of the literature) ................................ 308
8.5.9 Research Needs Related to the Deepwater Horizon Blowout, in the Canadian Context .......... 308
8.6 Pine River, British Columbia pipeline break ........................................................................... 309
8.6.1 Immediate effects, response and cleanup .................................................................................. 309
8.6.2 Water quality ............................................................................................................................. 310
8.6.3 Sediments .................................................................................................................................. 310
8.6.4 Periphyton and benthic invertebrates ....................................................................................... 310
8.6.5 Fish ........................................................................................................................................... 311
8.6.6 Important factors affecting the consequences of the Pine River spill ....................................... 312
8.6.6.1 Oil properties and behaviour ................................................................................................ 312
8.6.6.2 Effect of the environment on fate and behaviour .................................................................. 312
8.6.6.3 Oil toxicity ............................................................................................................................. 312
8.6.6.4 Spill response ........................................................................................................................ 312
8.6.7 Lessons learned from the Pine River spill................................................................................. 313
8.6.8 Research needs related to the Pine River spill.......................................................................... 313
8.7 Wabamun Lake, Alberta Train Derailment ............................................................................. 314
8.7.1 Early behaviour and effects of the spill..................................................................................... 314
8.7.2 Shorelines .................................................................................................................................. 315
8.7.3 Weathering ................................................................................................................................ 316
8.7.4 Observed effects on aquatic biota ............................................................................................. 316
8.7.4.1 Algae and macrophytes ......................................................................................................... 316
8.7.4.2 Zooplankton .......................................................................................................................... 317
8.7.4.3 Benthic invertebrates ............................................................................................................ 317
8.7.4.4 Fish ....................................................................................................................................... 317
8.7.4.5 Effects of the cleanup ............................................................................................................ 318
8.7.5 Analysis of the response to the spill .......................................................................................... 318
8.7.6 Public consultation and involvement ........................................................................................ 319
8.7.7 Aboriginal community concerns ............................................................................................... 319
8.7.8 Important factors affecting the consequences of the Wabamun Lake spill ............................... 319
8.7.8.1 Oil properties and behaviour ................................................................................................ 319
8.7.8.2 Effect of the environment on fate and behaviour .................................................................. 319
8.7.8.3 Oil toxicity ............................................................................................................................. 320
8.7.8.4 Spill response ........................................................................................................................ 320
8.7.9 Lessons learned from the Wabamun Lake spill......................................................................... 320
8.7.10 Research needs related to the Wabamun Lake spill .................................................................. 321
8.8 Spill from a Ruptured Pipeline into Talmadge Creek and the Kalamazoo River, Michigan .... 321
8.8.1 Oil-particle interactions and submergence ............................................................................... 322
8.8.2 Environmental effects ................................................................................................................ 322

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 14
8.8.3 Effects of cleanup activities....................................................................................................... 323
8.8.4 Evaluation of the spill response ................................................................................................ 325
8.8.5 Important factors affecting the consequences of the Kalamazoo River spill ............................ 325
8.8.5.1 Oil properties and behaviour ................................................................................................ 325
8.8.5.2 Effect of the environment on fate and behaviour .................................................................. 325
8.8.5.3 Oil toxicity ............................................................................................................................. 326
8.8.5.4 Spill response ........................................................................................................................ 326
8.8.6 Lessons learned from the Kalamazoo Spill ............................................................................... 326
8.8.7 Research needs related to the Kalamazoo spill ........................................................................ 326
8.9 Other Recent Spills ..................................................................................................................... 326
8.9.1 Westridge 2007 Spill, Burnaby, British Columbia .................................................................... 326
8.9.1.1 Lessons learned ..................................................................................................................... 327
8.10 Release of Fuel Oil from the Marathassa Grain Carrier into English Bay, British Columbia .. 328
8.10.1 Lessons learned ......................................................................................................................... 329
8.11 Release of Bitumen Emulsion from a Pipeline in Alberta ....................................................... 330
8.11.1 Lessons learned ......................................................................................................................... 330
8.12 Overall Conclusions Arising from the Case Studies ................................................................ 330
8.13 Prediction of Risk of Crude Oil Spills in Canada .................................................................... 333
8.13.1 Risk assessment for marine spills in Canadian waters south of 60th parallel .......................... 333
8.13.2 Risk assessment methods ........................................................................................................... 333
8.13.2.1 Spill frequency ...................................................................................................................... 333
8.13.2.2 Environmental sensitivity ...................................................................................................... 334
8.13.3 Results of the risk assessment ................................................................................................... 336
8.13.4 Major Conclusions and Recommendations from the Transport Canada Risk Assessment Report . 338
8.14 Risk Assessment for Marine Spills in Canadian Waters North of 60th Parallel ................... 339
8.14.1 Recommendations from the Arctic Marine Risk Assessment Report......................................... 341
8.15 Assessment of Marine Oil Spill Risk and Environmental Vulnerability for the State of Alaska .. 341
8.16 Assessment of Risks of Hypothetical Spills from the Proposed Northern Gateway Pipeline ..... 343
8.16.1 Review and Recommendations Based on the Key Findings of the Northern Gateway Pipeline
Ecological Risk Assessments ................................................................................................................. 345
8.17 Assessment of Risks of Hypothetical Spills from the Proposed Trans Mountain Pipeline
Expansion Project ................................................................................................................................... 349
8.17.1 Semi-quantitative assessment of spills at selected locations ..................................................... 349
8.17.1.1 Stochastic modeling of spills at locations near and at the mouth of the Fraser River ......... 351
8.17.1.2 Comparative assessments of Unmitigated and Mitigated Spills at the Westridge Marine
Terminal .............................................................................................................................................. 353
8.18 Assessment of Accidents and Malfunctions for the Energy East Pipeline Project ................ 354
8.19 Summary Comments and Recommendations Regarding Risk Assessments of Oil Spills
within Environmental Impact Assessments .......................................................................................... 355
8.20 Assessment of Risks Associated with Transport of Crude by Rail ......................................... 357
8.21 Overall Conclusions Regarding Assessment of Risks of Oil Spills ......................................... 358
8.22 Research Needs and Recommendations .................................................................................... 360
8.22.1 High priority research needs (short-, medium-, and long-term)............................................... 360
8.22.1.1 Gap: Baseline information .................................................................................................... 360
8.22.1.2 Gap: National, accessible databases for use in assessment of oil spills ............................... 361
8.22.1.3 Gap: Shoreline sensitivity in the Arctic ................................................................................ 361
8.22.1.4 Gap: Shoreline sensitivity in inland Canada ........................................................................ 361
8.22.1.5 Gap: Efficacy of oil spill responses for various types of oil ................................................. 361
8.22.1.6 Gap: Sufficient information to conduct NEBA for Arctic spills ............................................ 362
8.22.1.7 Gap: Finer-scale relative risk analysis that allows focused preparedness and response
planning .............................................................................................................................................. 362

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 15
8.22.1.8 Gap: Limited data from past spills in Canada ...................................................................... 362
8.22.2 Medium-priority research needs (medium-term and long-term) .............................................. 363
8.22.2.1 Gap: Risks of residual oil in Arctic ecosystems .................................................................... 363
8.22.2.2 Gap: Risks of deposition to deep sea sediments.................................................................... 363
8.22.2.3 Gap: The potential contribution of hyporheic flow to oil spill risk in rivers ........................ 363
8.22.2.4 Gap: Data supporting assessment of effects on ecosystem services ..................................... 363
8.22.3 Long-term research needs ......................................................................................................... 364
8.22.3.1 Gap: Effects of oil spills on trophic dynamics ...................................................................... 364
8.22.3.2 Gap: Potential for population-level effects ........................................................................... 364
8.22.4 Operational preparedness needs .............................................................................................. 364

CHAPTER 9: SUMMARY CONCLUSIONS AND RECOMMENDATIONS ...................................................... 367

APPENDIX A ............................................................................................................................................. 380


APPENDIX B.............................................................................................................................................. 387
APPENDIX C ............................................................................................................................................. 388
APPENDIX D ............................................................................................................................................. 391

REFERENCES ............................................................................................................................................ 414

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 16
LIST OF ACRONYMS

A-Nat: “natural” site on the Athabasca River, AB, unaffected by oil sands development
ACR: acute-chronic ratio
ADEC: Alaska’s Department of Environmental Conservation
AER: Alberta Energy Regulator; Alberta Government agency
AhR: aryl hydrocarbon receptor.
AMOP: Arctic and Marine Oilspill Program; an international technical forum held in Canada since 1978
ANS: Alaska North Slope crude oil
API: American Petroleum Institute
AMRA: Arctic Marine Risk Assessment
ASMB: Alberta Sweet Mixed Blend
ASTM: formerly the American Society for Testing and Materials; an international organization that
publishes consensus technical standards and analytical methods
AUV: autonomous underwater vehicle
BaP: Benzo(a)pyrene, a 5-ringed PAH
BAT: best available technology
bbl: barrel; a unit of volume that varies between countries and for different products. A barrel of crude oil
is equivalent to 159 litres.
BSEE: Bureau of Safety and Environmental Enforcement, U.S. Department of the Interior
BOEM: Bureau of Ocean Energy Management, U.S. Department of the Interior
BSD: blue sac disease
BTEX: benzene, toluene, ethylbenzene and xylene isomers: monoaromatic hydrocarbons commonly
present in petroleum, particularly in light crude oils and diluents
CAPP: Canadian Association of Petroleum Producers
CAT: Catalase
CBBO: Cosco Busan bunker oil
CEPA: Canadian Energy Pipeline Association
CEWAF: chemically-enhanced water-accommodated fraction of oil
CNRL: Canadian Natural Resources Ltd.; an oil sands mining company
CROSERF: Chemical Response to Oil Spills Ecological Effects Research Forum
CYP1A: Cytochrome P450 1A
DFO: Department of Fisheries and Oceans (Canada)
DNA: deoxyribonucleic acids
DW: Dry weight
DWH: Deepwater Horizon; subsurface blowout of very light crude oil in the Gulf of Mexico in 2010
E-Nat: “natural” site on the Ells River, AB, unaffected by oil sands development
EBSA: Ecologically and Biologically Significant Areas
EC50: median effective concentration
ECOD: Ethoxycoumarin-o-deethylase
ECRC-SIMEC: Eastern Canada Response Corporation, certified by Transport Canada to respond to oil
spills from tankers and oil-handling facilities in the Great Lakes, Quebec and Atlantic regions.
EDCF: effects-driven-chemical-fractionation
EEM: environmental effects monitoring
EEPP: Energy East Pipeline Project

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 17
EI2V: a microbial isolate from Elrington Island, PWS
EIA: environmental impact assessment
ERA: ecological risk assessment
ERCB: Energy Resources Conservation Board (Alberta), now replaced by AER
EROD: Ethoxyresorufin-o-deethylase
ESI: environmental sensitivity index
EVOS: Exxon Valdez Oil Spill; spill of Alaska North Slope crude oil into Prince William Sound in 1989
FAC: Biliary fluorescent aromatic compounds
FID: flame ionization detector
GC: gas chromatography
GSI: gonad somatic index
GST: glutathione transferase
GTH: glutathione
HC: hydrocarbon
HEWAF: high energy water-accommodated fraction of oil
Hexaco: hexacosaner
HFO: heavy fuel oil, e.g., Bunker C
HLB: Hydrophilic-lipophilic balance
HMW: high molecular weight
HPLC: high performance liquid chromatography
IFO: intermediate fuel oil, e.g., IFO 180
IP: intraperitoneal
ISB: in situ burning (of oil on water)
ITOPF: International Tanker Owners Pollution Federation; a not-for-profit marine ship pollution
response advisory body
Kow: octanol-water partition coefficient
LC50: median lethal concentration
LMW: low molecular weight
LOEC: lowest observable effect concentration
LOET: lowest observable effects time
LWO: less weathered oil
MC-252: Macondo crude oil
MDA: malondialdehyde
MESA: Medium South American crude oil
MoA: Mode of action
MOA: Mechanism of action
MMW: medium molecular weight
MNA: monitored natural attenuation
MS: mass spectrometry
MWO: more weathered oil
NAS: National Academy of Sciences (United States)
NAs: naphthenic acids
NEB: National Energy Board (Canada)
NEBA: net environmental benefits analysis

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 18
NGP: Northern Gateway Pipeline
NOAA: National Oceanographic and Atmospheric Administration, U.S. Department of the Interior
NOECC: no observable effect concentration
NRC: National Research Council (in both Canada and the United States)
NTSB: National Transportation Safety Board (United States)
OHCB: obligate hydrocarbonoclastic bacteria
OMA: oil-mineral aggregate
OPA: oil-particle aggregate
OWD: oil-water dispersion
PAC: polycyclic aromatic compounds
PAH: polycyclic aromatic hydrocarbon
PBCO: Prudhoe Bay Crude Oil
PDMS: polydimethylsiloxane
PHE: phenanthrene
PHMSA: U.S. Pipeline and Hazardous Materials Safety Administration
ppm: parts-per-million; a unit of concentration; e.g., mg/L or mg/kg.
ppt: parts-per-thousand; a unit of concentration; e.g., g/L or g/kg.
PTO: pole treating oil
PWS: Prince William Sound
RBCC: red blood cell count
RO: response organization
ROS: reactive oxygen species
ROV: remotely operated vehicle
RTG-2: rainbow trout gill (cell line)
S-nat: “natural” site on the Steepbank River, AB, unaffected by oil sands development
SAGD: steam-assisted gravity drainage
SAR: synthetic aperture radar
SARA: an acronym for the four major chemical classes of petroleum constituents: saturates, aromatics,
resins and asphaltenes
SCAT: shoreline cleanup and assessment technique
SCO: synthetic crude oil
SOD: superoxide dismutase
SSD: species sensitivity distributions
TAN: total acid number
TCRA: Transport Canada Risk Assessment
THC: total hydrocarbon (sum of concentrations of all hydrocarbon measured)
TLM: target lipid model
TMPEP: Trans Mountain Pipeline Expansion Project
TPAH: total polycyclic aromatic hydrocarbon
TPH: total petroleum hydrocarbon
TSB: Transportation Safety Board of Canada
TSPS: Tanker Safety Panel Secretariat, Transport Canada
TU: toxic unit
UCM: unresolved complex mixture

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 19
UME: unusual mortality event
U.S. EPA: United States Environmental Protection Agency
USGS: U.S. Geological Survey
v/v: volume to volume (dilution)
VEC: valued ecosystem component
VHS: viral hemorrhagic septicemia
VOCs: volatile organic compounds
WAF: water-accommodated fraction of oil
WCMRC: Western Canada Marine Response Corporation
WSF: water-soluble fraction of oil
WWP: waste water pond

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 20
GLOSSARY OF TERMS

Acute-chronic ratio: the ratio of the acute and chronic toxicity values for a given compound, usually the
average of the ratios for a variety of species; used to estimate the chronic toxicity of a compound
or a mixture of compounds, from the measured or modeled acute toxicity when no chronic
toxicity data are available.
Alberta Sweet Mixed Blend: a major blend of light crude oil exported from Alberta; also used by
Environment Canada as a reference for laboratory tests.
Annual Exceedance Probability: the chance or probability of a natural hazard event (usually a rainfall or
flooding event) occurring annually and is usually expressed as a percentage. Bigger rainfall
events occur (are exceeded) less often and will therefore have a lesser annual probability.
Aromatics: class of hydrocarbons comprising one or more benzene ring structures having alternating
double bonds; may have one or more alkyl side chains in various positions, but no heteroatoms.
Aryl hydrocarbon receptor: a cellular protein that binds planar or plate-like polycyclic aromatic
compounds with a shape and dimensions that resemble 2,3,7,8 tetrachorodibenzo(p)dioxin.
Asphaltenes: class of petroleum compounds of high molecular weight and complexity; defined as the oil
fraction that precipitates in low molecular weight n-alkanes (e.g., C5-C7) but is soluble in
toluene.
Bioaccumulation (or bioconcentration): the tendency of substances to accumulate in the body of
organisms; the net uptake from their diet, respiration or transfer across skin and loss due to
excretion or metabolism. The bioaccumulation factor (BAF) or bioconcentration factor (BCF) is
the ratio of concentrations in tissue to concentrations in a source, i.e., water or diet.
Bioavailability (or biological availability): compound that is in a physical or chemical form that can be
assimilated by a living organism; also, the proportion of a chemical in an environmental
compartment (e.g., water) that can be taken up by an organism.
Biodegradation: a natural process of microbial transformation of chemicals, such as oil under aerobic or
anaerobic conditions; oil biodegradation usually requires nutrients, such as nitrogen and
phosphorus; transformation may be complete, producing water, carbon dioxide and/or methane,
or incomplete, producing partially-oxidized chemicals.
Biomagnification: a food chain or food web phenomenon whereby a substance or element increases in
concentration at successive trophic levels; occurs when a substance is persistent and is
accumulated from the diet faster than it is lost due to excretion or metabolism.
Biomarkers: a term used in two different ways, depending upon discipline. In petroleum chemistry, a
biomarker is a relic chemical relating its presence to the original biological source (microbial,
plant or animal); biomarkers are usually poorly or non-biodegradable and so persist in the oil,
enabling their use as internal standards in petroleum analysis. In environmental toxicology, a
biomarker is a biochemical process, product or cellular response that indicates the organism’s
exposure to a pollutant and/or the toxic effects of the pollutant.
Bioremediation: an intervention strategy to enhance biodegradation of spilled oil (or other
contaminants), ranging from no remedial action other than monitoring (natural attenuation) to
nutrient addition (biostimulation) to inoculation with competent microbial communities
(bioaugmentation).
Bitumen: heaviest class of petroleum, having high viscosity and density; widely considered to represent
the residue from lighter oils that have undergone biodegradation over geological time.
Blue sac disease: a disease syndrome of fish embryos caused by a variety of environmental stresses,

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 21
including exposure to PAH and to alkyl PAH.
Cn: a molecule, such as a hydrocarbon, having n carbon atoms.
Chemically-enhanced water-accommodated fraction of oil (CEWAF): a solution of hydrocarbons and
a suspension of oil droplets created when a chemical dispersant is added to oil and water with
stirring. See also WAF and HEWAF.
Conventional crude oil: commonly defined as liquid petroleum that flows in the reservoir and in
pipelines and is recovered from traditional oil wells using established methods, including primary
recovery and water flooding (e.g., condensates, light and medium crude oils), versus
unconventional crude oils.
Crude oil: synonymous with petroleum; a naturally-occurring and typically liquid complex mixture of
thousands of different hydrocarbon and non-hydrocarbon molecules.
CYP1A: a member of the cytochrome P450 family of proteins; an enzyme of vertebrates, including fish,
birds and mammals, that catalyzes the addition of oxygen to double bonds as a first step in the
metabolism and excretion of PAHs.
cyp1a: the gene that codes for CYP1A proteins.
Dilbit: bitumen diluted with a lighter petroleum class, such as condensate or naphtha, typically 70%
bitumen and 30% diluent; also see Synbit.
Dispersion: suspension of oil droplets in water accomplished by natural wind and wave action,
production of biological materials (biosurfactants) and/or chemical dispersant formulations.
Dispersant: a chemical or mixture of chemicals applied, for example, to an oil spill to disperse the oil
phase into small droplets in the water phase.
EC50: the concentration of a substance that causes sublethal effects on the median or 50 th percentile
organism tested within a specified exposure time (e.g., 14 d EC50)
Effects-driven-chemical-fractionation (EDCF): the step-wise fractionation, toxicity testing and
chemical analysis of complex mixtures of compounds to isolate and identify the constituents
responsible for the toxic effects of the whole mixture.
Ecological risk assessment: process for analyzing and evaluating the possibility of adverse ecological
effects caused by environmental pollutants.
Emulsification: formation of water droplets in an oil matrix (water-in-oil) or conversely oil droplets in a
water matrix (oil-in-water) achieved by the action of agitation, such as wind and wave activity;
can be unstable, separating into oil and water phases again soon after formation, or stable for
months or years (e.g., ‘chocolate mousse’, a water-in-oil emulsion).
Environmental impact assessment: the process of measuring or estimating the environmental effects of
pollutants, such as oil spills, relative to conditions at a reference site or to a time prior to a spill.
Evaporation: the physical loss of low molecular weight components of an oil to the atmosphere by
volatilization.
Flame ionization detector (FID): used with analytical instruments like gas chromatographs to detect
components of petroleum by combustion ionization, hence GC-FID
Fracking: see hydraulic fracturing.
Gas chromatography (GC): an analytical method used to characterize petroleum components; GC is
combined with different detection methods, hence GC-FID, GC-MS, etc.
HEWAF: A solution of hydrocarbons and a suspension of oil droplets created when oil and water are

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 22
mixed by high energy agitation. See also WAF and CEWAF.
Heteroatom: in petroleum, an atom such as nitrogen, sulfur and/or oxygen that is part of a hydrocarbon
skeleton, such as found in the Resins fraction of crude oils
High molecular weight (HMW): relative term referring to the molecular mass of chemicals; in oil,
asphaltenes would be typical of HMW compounds.
High performance liquid chromatography (HPLC): an analytical method for separating chemicals in
solution.
Hydrocarbon: a chemical that is composed of only carbon and hydrogen; chemicals containing
heteroatoms, such as nitrogen, sulfur and/or oxygen, are not hydrocarbons, even though they may
petroleum constituents.
Hydraulic fracturing: also known as fracking’; an unconventional method for recovering liquid
petroleum from shale oil deposits that otherwise would not release the gas and/or fluid in the
reservoir.
Hydrogen sulfide (H2S): a toxic, flammable and corrosive gas sometimes associated with petroleum.
Hydrophilic-lipophilic balance (HLB): a measure of the properties of dispersants and surface-active
agents at the interface of polar and non-polar liquids, such as oil and water. The higher the
surfactant HLB value, the more hydrophilic (water-loving) it is.
Hyporheic flow: the perculating flow of water through the sand, gravel, sediment and other permeable
soils under and beside the streambed.
Intermediate fuel oil: the heaviest commercial class of refined petroleum diluted with a lighter refined
product, commonly burned in furnaces, boilers or ship engines; e.g., IFO 180.
Isomers: chemicals that have the same molecular formula (i.e., elemental composition) but different
structures; may also have different properties, including water solubility, biodegradability and
toxicity.
Kow: the partition coefficient describing the equilibrium concentration ratio of a dissolved chemical in
octanol versus in water, in a two-phase system at a specific temperature; used in prediction of
toxicity.
LC50: the concentration of a substance toxic to the median or 50th percentile organism tested within a
specified exposure time (e.g., 96 h LC50).
Lacustrine: relating to lakes.
Low molecular weight: relative terms referring to the molecular mass of chemicals; in oil,
monoaromatics and aliphatics up to C10 would be typical of these compounds.
Mass spectrometry: an analytical method used for detailed characterization of petroleum components,
often in combination with GC, hence GC-MS.
Mechanism of action (MOA): describes a functional or anatomical change at the molecular level.
Medium molecular weight: relative terms referring to the molecular mass of chemicals; in oil, 3 to 6-
ringed PAH and aliphatics up to C20 would be typical of MMW compounds.
Mineralization: complete oxidation of a compound (e.g., hydrocarbon) to carbon dioxide and water; may
be accomplished by a single species of organism or by a community of microbes.
Mode of action (MoA): describes a functional or anatomical change at the cellular level, resulting from
the exposure of a living organism to a substance.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 23
Monitored natural attenuation: a remediation strategy in which there is no intervention but the site is
monitored using various parameters.
Monoaromatics: aromatic hydrocarbons having only a single benzene ring; may also have one or more
alkyl side chains.
Naphthenic acids: a class of polar petroleum compounds that contributes to aquatic toxicity and to
petroleum infrastructure corrosion by contributing to TAN.
Natural attenuation: remediation of a contaminated site by natural processes alone, without human
intervention; also see monitored natural attenuation.
Obligate hydrocarbonoclastic bacteria: bacteria that have adapted to use hydrocarbons as their sole
source of carbon and energy for growth and metabolism.
Oil-mineral aggregate (OMA): floc containing oil adhering to mineral particle, which may float, sink or
re-suspend in a water column; a process initially referred to as clay-oil flocculation.
Oil-particle aggregate (OPA): more general term than OMA, describing oil adhering to particles that
may be inorganic (minerals) and/or organic, including microbial cells.
Partitioning: the diffusion of compounds between two immiscible liquid phases, including water and oil
droplets and water and lipid membranes.
Petroleum: synonymous with crude oil; a naturally-occurring complex mixture of thousands of different
hydrocarbon and non-hydrocarbon molecules.
Photooxidation: oxidation due to the influence of photic energy, usually from UV light.
Photo-enhanced toxicity: increased toxicity due to photooxidation in vivo.
Polycyclic aromatic hydrocarbons: (PAHs) a subclass of aromatic hydrocarbons having two or more
fused benzene rings; may also have one or more alkyl side chains, generating large suites of
isomers; some are considered ‘priority pollutants’ because of their toxicity and/or potential
carcinogenicity.
Reactive oxygen species (ROS): a group of chemicals that can cause cellular damage due to their high
reactivity; are released during reactions that add oxygen to double bonds. In turn, ROS can react
with double bonds in lipids, proteins and nucleic acids to change their structure and function.
Resins: a solubility class of poorly characterized, polar petroleum compounds in which each molecule
contains one or more atoms of nitrogen, sulphur and/or oxygen in a hydrocarbon skeleton.
Riparian: related to being situated on or dwelling on the bank of a river or other body of water, such as a
lake or tidewater.
Riverine: relating to lakes.
SAGD (steam-assisted gravity drainage): an unconventional in situ method to produce bitumen from
deep oil sands deposits without surface mining operations.
Saturates: class of hydrocarbons that may be straight-chain, branched-chain or cyclic, in which all carbon
atoms have single bonds to either carbon or hydrogen.
Shale oil: also known as ‘tight oil’ (but not to be confused with ‘oil shale’); liquid petroleum that is
produced from shale oil reservoirs, typically by hydraulic fracturing methods.
Sour crude: petroleum that has a >1% total sulfur content that may be present as hydrogen sulfide and/or
as organic forms of sulfur in a hydrocarbon backbone.
Species sensitivity distributions (SSD): compare the cumulative proportion of species (percentile)

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 24
affected by a chemical to a toxicity endpoint measured for each species (e.g., 96 h LC50); SSD
model assumes that species sensitivity is randomly distributed.
Sweet crude: petroleum with low total sulphur content, variously defined as <0.5% or <1% sulphur.
Synbit: bitumen diluted with synthetic crude oil, typically at 1:1 ratio.
Synthetic crude oil: a partially-refined fraction of bitumen; may be used as a diluent to make dilbit for
transport.
Target lipid model: estimates the aqueous concentration of organic compounds, or mixtures of organic
compounds, that cause toxicity by narcosis.
Total acid number (TAN): a measure of the acidity determined by the amount of potassium hydroxide
in milligrams that is needed to neutralize the acids (typically naphthenic acids) in one gram of oil;
used by refineries as an indicator of potential corrosion and scale production.
Total petroleum hydrocarbons (TPHs): the total mass of all hydrocarbons in an oil or environmental
sample, including the volatile and extractable (non-volatile) hydrocarbons; may be further defined
by stating the analytical method used, e.g., GC-detectable TPH or TPH-F (TPH measured by
fluorescence), which vary in their rigour.
Total polycyclic aromatic hydrocarbons (TPAHs): including alkyl-PAHs and parent (unsubstituted)
PAHs; the sum of all concentrations of PAHs measured by GC-MS.
Toxic unit (TU): ratio between the concentration of a compound in water and a toxic endpoint (e.g., 96 h
LC50).
Unconventional crude oils: petroleum that does not flow readily in the reservoir and/or must be
produced by using unconventional methods, such as surface mining of shallow bitumen deposits,
steam-assisted gravity drainage (SAGD) for in situ extraction of deep bitumen deposits, cyclic
steam injection for heavy oils, or horizontal drilling with hydraulic fracturing for recovery of light
shale oils.
Unresolved complex mixture (UCM): petroleum constituents that are not resolved by conventional GC
and appear as a ‘hump’ in the gas chromatogram; comprises many hundreds or thousands of
unresolved isomers.
Unusual mortality event (UME): term coined by NOAA to describe a greater-than-usual rate of
mortality of marine mammals.
Volatile organic compounds (VOCs): chemicals having high vapour pressure at room temperature (and
corresponding low boiling point) that therefore tend to evaporate or sublimate into the air; for
example, BTEX.
Water-accommodated fraction of oil (WAF): hydrocarbons that will partition from oil to water during
gentle stirring or mixing; may contain droplets, in contrast to water-soluble fractions (WSF).
Water-soluble fraction of oil (WSF): aqueous solution of hydrocarbons that partition from oil; does not
include droplet or particulate oil. See also CEWAF and HEWAF.
Weathering: a suite of changes in spilled oil composition and properties brought about by a variety of
environmental processes including spreading, evaporation, photooxidation, dissolution,
emulsification and biodegradation, among others.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 25
EXECUTIVE SUMMARY

A panel of leading experts on oil chemistry, behaviour and toxicity reviewed the current science relevant
to potential oil spills into Canadian marine waters, lakes, waterways and wetlands. The review, which
examined spill impacts and oil spill responses for the full spectrum of crude oil types (including bitumen,
diluted bitumen and other unconventional oils), is among the most comprehensive of its kind. It surveyed
scientific literature, key reports and selected oil spill case studies, including tanker spills, an ocean rig
blowout, pipeline spills and train derailments. The Panel also consulted industry, government and
environmental stakeholders across the country.

The Panel found that the dozens of crude oil types transported in Canada exist along a chemical
continuum, from light oils to bitumen and heavy fuels, and the unique properties of each of these oil types
(their chemical ‘fingerprints’) determine how readily spilled oil spreads, sinks, disperses, and impacts
aquatic organisms, including wildlife, and what proportion ultimately degrades in the environment.
Despite the importance of oil type, the Panel concluded that the overall impact of an oil spill, including
the effectiveness of an oil spill response, depends mainly on the environment and conditions (weather,
waves, etc.) where the spill takes place and the time lost before remedial operations.

The Panel recommends that this critical research should concentrate on seven general high-priority
research needs:

High-Priority Research Needs Identified by the Expert Panel

1. Research is needed to better understand the environmental impact of spilled crude oil in high-risk
and poorly understood areas, such as Arctic waters, the deep ocean and shores or inland rivers
and wetlands.
2. Research is needed to increase the understanding of effects of oil spills on aquatic life and wildlife
at the population, community and ecosystem levels.
3. A national, priority-directed program of baseline research and monitoring is needed to develop an
understanding of the environmental and ecological characteristics of areas that may be affected
by oil spills in the future and to identify any unique sensitivity to oil effects.
4. A program of controlled field research is needed to better understand spill behaviour and effects
across a spectrum of crude oil types in different ecosystems and conditions.
5. Research is needed to investigate the efficacy of spill responses and to take full advantage of
‘spills of opportunity’.
6. Research is needed to improve spill prevention and develop/apply response decision support
systems to ensure sound response decisions and effectiveness.
7. Research is needed to update and refine risk assessment protocols for oil spills in Canada.

Timeframes for conduct of the recommended studies are provided by the Panel. This executive summary
highlights these central conclusions and other priority research questions identified in the report
chapters.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 26
What happens when crude oil spills into oceans, into lakes or into the waterways that wind through our
forests, fields and towns? Canada produces some three million barrels of oil every day, importing
hundreds of thousands more, and all of it travels somewhere. In this lake-and-river-rich country with the
world’s longest coastline, crossing water is inevitable when oil is transported. Whether in vast tankers
traversing the sea or in pipelines, trucks and trains passing countless rivers, lakes and wetlands, oil is on
the move. Some is drilled directly from the seabed where offshore rigs perch above ocean waves.

Meanwhile, accidents happen. Headline-grabbing calamities, such as the Deepwater Horizon blowout in
the Gulf of Mexico in 2010, the Exxon Valdez spill off Alaska in 1989 and the Arrow spill off the coast of
Nova Scotia in 1970, are periodic reminders that oil spills can shock the environment, the economy and
the communities affected by them—at least in the short-term. Water is fouled. Wildlife is tarred. Fisheries
and other industries struggle to recover.

The good news is that transporting oil at sea is safer than it has ever been. According to the International
Tanker Owners Pollution Federation, large tanker spills occurred almost 14 times more often during the
1970s on average than they do today. Undersea blowouts during oil production and exploration are also
rare (although Canadian offshore exploration and drilling is expected to increase). Less known is how
much oil spilled from pipelines, trains and trucks reaches our lakes, rivers and wetlands (where oil can
become trapped and remain concentrated causing more harm or creating more concern because towns and
cities are nearby). However, while big oil spills from grounded tankers, oil rigs, pipeline ruptures or train
wrecks are guaranteed newsmakers, in truth most of the oil-related chemicals that make it into our oceans
arrive from natural seepage, routine tanker maintenance and runoff from land.

Even so, the potential impact of spills into Canadian waters during the transport of oil can be profound.

The Royal Society of Canada Expert Panel report addresses this impact. Its purpose is to better
understand what we know and, perhaps more importantly, what we need to find out. Chief among the
report’s aims is to provide a roadmap to research questions concerning how crude oils, including diluted
bitumen and other unconventional oils, behave and how they affect ecosystems and communities after
spilling into the changeable and weather-affected environments of Canada’s vast marine and inland
waters.

The Expert Panel has highlighted hundreds of conclusions and identified a long list of research needs in
its extensive report. This executive summary highlights only the most pressing of these research
priorities. Answers to these research questions are considered by the Panel to be essential for equipping
policy makers, oil industry decision-makers, oil spill responders and other Canadians with critical tools to
better anticipate spills and their consequences and to better protect Canada’s marine and inland waters
from the adverse effects of spilled oil. The Panel has consolidated seven general, high-priority research
needs (see the box above) in this executive summary from hundreds of research priorities identified in the
report chapters. The rationale for these research needs, and more detailed descriptions, can be found
within the report itself.

CHAPTER 1: INTRODUCTION

The Royal Society of Canada (RSC) Expert Panel was established in response to a request from the
Canadian Energy Pipeline Association (CEPA) and the Canadian Association of Petroleum Producers
(CAPP). The request was the result of a widespread recognition within industry, governments and
elsewhere that Canadians should know what to expect in the event of an accidental spill, and that those
who move oil and respond to spills have the information they need to protect our environment, economy
and communities across the country.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 27
The Panel, composed of international experts on oil chemistry, behaviour and toxicity, reviewed the
current science relevant to crude oils spilled into Canadian marine waters, lakes, waterways and wetlands
(spills of gasoline, diesel and other refined fuels were not considered). The Panel relied on scientific
literature, key reports and selected oil spill case studies, including well-known tanker spills (e.g., the
Arrow spill in 1970 and the Exxon Valdez spill in 1989), the Deepwater Horizon spill of 2010, pipeline
spill ruptures and train derailments.

The Panel’s work also involved extensive consultations with key industry, government and environmental
stakeholders across the country. These included representatives from CEPA and CAPP, government
agencies in Canada (Environment Canada, Natural Resources Canada, Fisheries and Oceans Canada,
Alberta Innovates) and the United States (the National Oceanographic and Atmospheric Administration),
private sector consultants, oil spill response agencies (e.g., the Eastern Canada Response Corporation),
non-government organizations (e.g., Greenpeace), as well as other academics and interested individuals.
Formal consultations included public forums involving open, online access held in Calgary in February
2015 and in Halifax in April 2015. A third Panel meeting in June 2015 included informal discussions with
attendees at the 38th international Arctic and Marine Oilspill Program (AMOP) technical conference in
Vancouver.

These consultations and the scientific review examined spill impacts and oil spill responses for the full
spectrum of oil types, from ultra-light condensates and light oils to bitumen, diluted bitumen and heavy
fuels. Many of the largest knowledge gaps were found to be associated with the chemical composition
and environmental behaviour of emerging petroleum types, including diluted bitumens and other
unconventional oils.

CHAPTER 2: CHEMICAL COMPOSITION, PROPERTIES AND BEHAVIOUR OF SPILLED OILS

While many Canadians think of the oil travelling by pipeline, train, truck or tanker as much the same, the
crude oil crossing the country or plying its offshore waters each day represents dozens of different types.
This may seem trivial to some, but if the oil spills into water, the type of oil involved can make a world of
difference: How much damage it does, how easy it is to cleanup and how readily the oil degrades in the
environment.

Each of the oil types transported in Canada is a complex mixture of thousands of chemicals. While these
different types can be thought as existing along a kind of chemical continuum (from ultra-light oil
condensates and light oils to heavy crude oils and the thick bitumen commonly associated with Canada’s
oil sands), each has its unique ‘chemical’ fingerprint.

The Panel found that this chemical fingerprint is a key predictor of not only the physical properties of the
oil (e.g., how heavy or thick it is), but also its behaviour in the environment (e.g., how it spreads, sinks or
disperses in water), its toxic effects on aquatic organisms and humans, and its susceptibility to
degradation by ‘weathering’ (i.e., changes to the oil caused by exposure to sunlight, waves, weather
conditions and microorganisms in the environment). How the fingerprint of each spilled oil type changes
in the environment is an important tool for spill responders for monitoring cleanup efforts and setting
cleanup goals.

Although the Panel found the chemical composition and behaviour of many oil types have been well-
studied, more research is needed to better understand the chemistry, properties and spill behaviour of
newer, less-familiar oils, such bitumen, diluted bitumen blends and other unconventional oils.

CHAPTER 3: EFFECT OF ENVIRONMENT ON THE FATE AND BEHAVIOUR OF OIL

The unique features of the environment where an oil spill takes place are at least as important as the type
of oil in determining effects on aquatic ecosystems.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 28
Saltwater straits, freshwater lakes, running rivers and dense wetlands are home to distinctive
combinations of physical characteristics, water and sediment chemistry and natural communities of
microorganisms that can transform oil as it spills and spreads. Microorganisms, for example, degrade
various hydrocarbons found in different oil types to varying degrees, and their impact is often an
important part of oil spill cleanup strategies. Sunlight, wind, waves and weather conditions can physically
and chemically transform a spill. Temperature, dissolved oxygen, nutrient supply, salinity and pH also
alter the composition and behaviour of contaminating oil. These changes to the chemistry of oil are
crucial factors affecting how spilled oil spreads, affects aquatic organisms and people or lingers in the
environment.

Indeed, the Panel found that, despite the importance of oil type, the overall impact of an oil spill,
including the effectiveness of an oil spill response, depends mainly on the environmental characteristics,
the conditions where the spill takes place and the speed of response.

The impact on spilled oil of the characteristics and conditions of a spill site has been carefully studied in
some environments, but knowledge gaps remain. In particular, research is needed to better understand
what happens to oil spilled into the cold, icy, yet ecologically sensitive waters of the Arctic, where
interest in oil exploration, production and shipping is on the rise. Similarly, little is known about the fate
of oil and its impact in permafrost areas or in marine environments covered in ice. Microorganisms that
break down oil are considered less active when temperatures are near freezing, but this relationship may
not be as clear as we think and further study is needed.

High-Priority Research Need #1

Research is needed to better understand the environmental impact of spilled crude oil in high-risk
and poorly understood areas, such as Arctic waters, the deep ocean and shores or inland rivers and
wetlands.

i. Research is needed to assess the complex interactions among physical, chemical and biological
factors unique to Arctic conditions (e.g., extreme cold temperatures, permafrost ecosystems, snow
and ice) and different types of spilled crude oil. (Timeframe: Within 5 years)
ii. Research is needed to assess the fate and behaviour of oil spilled into freshwater ecosystems,
especially in northern bogs, fens and areas of permafrost. (Timeframe: Within 5 years)
iii. Research is needed to evaluate risks associated with the shipment of fuel oil to communities in the
Arctic. (Timeframe: Within 5 years)
iv. Research is needed to assess the risks of deep sea blowouts in the Beaufort Sea and in areas of the
Atlantic coast that support commercial and subsistence fisheries, including research into the
behaviour of oil on the surface with and without ice and the effects of subsurface oil plumes,
residual oil deposited on deep sea sediments, oil stranded along shorelines and in backwater,
marshy areas, and the impact of dispersant additions. (Timeframe: 5-10 years)
v. Research is needed to assess the risks of pipelines in Arctic freshwater environments, with an
emphasis on the Mackenzie River. (Timeframe: 5-10 years)
vi. Research is needed to investigate the fate of unrecovered oil in rivers where it can interact with ice,
substrates, woody debris, bed sediments, groundwater and engineered structures. (Timeframe: 5-10
years)

CHAPTER 4: OIL TOXICITY AND ECOLOGICAL EFFECTS

Oil spills can have significant consequences for aquatic ecosystems. These effects can be both short-lived
and long-lasting. In the days following a spill, floating oil smothers mollusks, plants and other species at
the shoreline. Oil on birds and mammals destroys their thermal insulation and buoyancy. Some chemical

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 29
components of spilled oil dissolve in water and kill fish and other aquatic creatures (before they typically
break down quickly and disappear). Other chemicals, such as polycylic aromatic hydrocarbons (PAHs),
can persist in the water and cause chronic health effects for aquatic species that show up months or years
later.

Light oils contain more compounds that are acutely toxic to aquatic organisms than medium or heavy oils.
On the other hand, heavy oils contain more of the chronically toxic alkyl PAHs. The Panel could not
conclude that diluted bitumens present a greater or lesser health risk to most species than other oils
because there are too few data available on toxicity. However, there may be a greater risk to bottom- and
sediment-dwelling organisms due to the tendency for diluted bitumens to sink in fresh water under certain
conditions.

The characteristics of the oil spill location and its environment determine how spilled oil affects aquatic
biota. Oil spills into fresh water, for example, are generally smaller than marine spills, but they may have
a greater relative impact because the oil can’t be diluted and degraded by the large volumes of water
available at sea. Inland shorelines and sediments are more likely to become fouled, and less time is
available to contain a freshwater oil spill before it contaminates sensitive habitats.

It is not only the spills themselves that threaten ecosystems, but oil spill cleanup can be damaging as well.
Physical cleanup (e.g., removing oiled vegetation or tarred shoreline) destroys habitat and can cause
erosion or the buildup of silt. Habitat damage reduces the abundance and productivity of native species
and fosters invasive species. Using chemicals to disperse spilled oil often means surface oil is transferred
to subsurface water at concentrations that can be toxic to aquatic life (especially to fish embryos). More
research is needed on spill cleanup methods that limit habitat damage and the threats to wildlife.

Oil spill impacts on aquatic ecosystems are difficult to measure. In many cases, information about the
ecology of a site before a spill occurs (i.e., baseline data) is scarce or missing altogether. Baseline
monitoring is typically the responsibility of provincial and federal government departments as part of
environmental and natural resource management. Coordination and collaboration is needed between the
oil industry and these government agencies to ensure that monitoring addresses the needs for data to
assess the distribution and effects of spilled oil in ecosystems most at risks of spills.

Assessment of oil spill impacts on ‘ecosystem services’ should be considered. Ecosystem services are the
benefits provided by ecosystems to humans that contribute to making human life both possible and
fulfilling. Further research is also required to better understand how the toxicity of spilled oil is affected
by its interaction with the environment in which it was spilled.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 30
High-Priority Research Need #2

Research is needed to increase the understanding of effects of oil spills on aquatic organisms,
populations, communities and ecosystems.

i. Research is needed to investigate the cumulative and interactive effects of co-exposure to oil and
other human-induced and natural environmental stressors, such as industrial and municipal
pollution, extreme temperatures, salinity, low oxygen concentrations and elevated concentrations
of suspended sediments. (Timeframe: 5-10 years)
ii. Research is needed on the effects of spilled oil on populations and community structure of aquatic
biota. (Timeframe: 5-10 years)
iii. Research is needed to understand the indirect effects of oil spills on ecological processes, such as
interactions within and among trophic levels in aquatic food chains. (Timeframe: 10+ years)
iv. A program of research is needed on the resilience of aquatic ecosystems affected by oil spills,
particularly at sites of past spills and in ecosystems unique to northern Canada (e.g., bogs, fens,
etc.) at a high risk of oil exposure. (Timeframe: 10+ years)
v. Research is needed to investigate the socioeconomic impacts of oil spills as a first step in
implementing an ecosystem services approach to oil spill impact assessments. (Timeframe: 10+
years)

CHAPTER 5: MODELING OIL SPILLS IN WATER

Knowing what to expect and how to respond when oil spills into an ocean, lake or river is no mean feat.
The complex chemistry of each oil type makes it difficult to predict how the oil will act and change when
it meets the equally complex water chemistry, ecology and conditions at the particular site where a spill
takes place. For this predictive work, scientific models are invaluable tools.

Scientific modeling creates conceptual or mathematical representations of complex real-world phenomena


that can’t be readily observed. Scientific models of oil spills use what we know from experiments,
previous spills and other information to approximate what happens when oil of a particular type spills in
particular circumstances. Scientists are constantly adding information and refining these models to
improve their accuracy for predicting spill consequences and for understanding the best spill responses.

Early models of oil behaviour and transport relied heavily on experimental observations. Since the early
1980s, advances in oil spill modeling focused mainly on oil dispersion (the formation of oil droplets), the
formation of oil particle aggregates, emulsification, evaporation and the general transport of oil in open
water as well as in other types of ecosystems. More recently, researchers have developed more advanced
numerical models of these various processes to better predict oil’s behaviour and changes in situations
where no direct measurements can be made, such as in deep water or in the Arctic.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 31
The Panel found that while scientific modeling has made many advances in predicting how the
environment can influence spilled oil and its behaviour (through dispersion, biodegradation, dissolution,
etc.), more research is needed to improve models of oil-in-ice effects, oil dispersion by waves, oil droplet
formation from blowouts, the formation of oil particle aggregates and the biodegradation of oil droplets
under various environmental conditions (such as temperature, salinity, nutrient availability, light and
chemical dispersants).

High-Priority Research Need #3

A national, priority-directed program of baseline research and monitoring is needed to develop an


understanding of the ecological characteristics of areas that may be affected by oil spills in the
future.

i. Research is needed to collect and evaluate baseline information from high-risk, poorly
understood areas, such as the Arctic and other less-studied Canadian environments. (Timeframe:
Within 5 years)
ii. Research is needed to understand the current status of sensitive and highly-valued species and
vulnerable habitats for specific, pre-defined locations in Canada representing a range of human
disturbance, from relatively undisturbed to highly disturbed. (Timeframe: Within 5 years)
iii. Research is needed to create ecosystem sensitivity maps, prioritized according to recent relative
risk assessments, the intensity of current and potential future human use, the relative sensitivity
of ecosystems and geographic gaps (e.g., in large areas of inland Canada). (Timeframe: Within 5
years)
iv. Research is needed to understand the natural variability of population and community metrics
(e.g., abundance, diversity, productivity) across physical and chemical gradients, as well as
across time (seasonal and annual). (Timeframe: Within 5 –10 years)
v. Research is needed to identify other anthropogenic stressors that could influence the effects of oil
spills. (Timeframe: Within 5–10 years)

CHAPTER 6: A REVIEW OF SPILL RESPONSE OPTIONS

Just as types of crude oil are far from uniform and the environments and conditions where spills occur are
many, effectively responding to oil spills is complicated. Decisions about what response is best and the
likelihood of success depend not only on the oil type, environment and weather conditions, but also on
technical and logistical factors (such as the responders’ knowledge and skills, the availability of personnel
and equipment, time constraints, regulatory approvals, health and safety criteria, etc.), as well as financial
concerns (such as the cost and economic impacts of the spill). Other considerations include the level of
community engagement.

There are three main categories of oil spill responses. The first simply relies on natural processes to
disperse and degrade spilled oil. For instance, naturally occurring microorganisms can remove or break
down some of the hydrocarbons and other chemicals in the oil (called ‘natural attenuation’). Evaporation
can also help remove volatile and lighter weight components of spilled oil, while exposure to sunlight and
oxygen causes the natural photooxidation of some of the oil’s aromatic compounds. The second type of
response involves physically containing and removing spilled oil, often using booms and skimmers on the
water or washing and scraping at shore. Thick slicks of oil can also be burned at a spill site. The third
response type uses biological and chemical methods. This can involve methodologies to enhance the
growth of oil-degrading microbes and/or plants on contaminated sites (phytoremediation) or the
application of chemical dispersants that break up oil slicks into small droplets that become diluted into the
water column where they are eventually also biodegraded.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 32
Choosing the best response or combination of responses depends on the unique circumstances of each
spill. Among these are weather, wave height, ice conditions, daylight and ecological factors, including the
risk to fish, invertebrates and other wildlife. Technical and economic factors also play a role, as well as
the inherent effectiveness of the response strategy being considered.

The Panel found that most of what is known about oil spill response technologies has been developed
through laboratory work and case studies. A better understanding of appropriate spill responses in the
Arctic and in snow and ice conditions is vital. The Panel recommends carefully controlled field studies to
help close this knowledge gap (without significant negative impact on the environment). Research is also
needed to better understand less familiar response methods, such as anaerobic biodegradation in
sediments, and emerging technologies (e.g., bioventing, air sparging, etc.) for aiding in the cleanup of
anaerobic or anoxic sea floor and lake bottom environments contaminated by sunken oil. What happens to
chemically-dispersed oil in both the deep sea and on its surface also needs to be studied using controlled
empirical experiments.

High-Priority Research Need #4

A program of controlled field research is needed to better understand spill behaviour and effects
across a spectrum of crude oil types in different ecosystems and conditions.

i. Controlled field experiments on oil spills (sanctioned by the federal government through a new
permitting system) with rigorous statistical designs are needed at a variety of sites representing
different coastal marine and freshwater ecosystems and conditions. (Timeframe: Within 5 years
and beyond)
ii. Research is needed at the site of previous oil spills in Canada to increase our understanding of the
effects of spilled oil over the long-term and of the extent of natural cleanup. (Timeframe: Within 5
years)

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 33
High-Priority Research Need #5

Research is needed to investigate the efficacy of spill responses and to take full advantage of ‘spills of
opportunity’.

i. Research is needed to help develop effective oil spill response measures tailored to the Arctic,
including studies that explore the interactions of oil with permafrost and ice or that examine the
microbial degradation of oil at low temperatures. (Timeframe: Within 5 years and beyond)
ii. Advanced planning and contingency funds are needed to support research on the fate, behaviour
and effects of real-world oil spills as they occur (‘spills of opportunity’) in the short, medium and
long-term, including studies of the relative effectiveness of response measures. (Timeframe: Within
5 years)
iii. Indigenous peoples and their traditional knowledge should be involved in the development of
research protocols, in oil spill preparedness, cleanup and remediation/restoration, including
involvement in the investigations of ‘spills of opportunity’. (Timeframe: Within 5 years)
iv. Research is needed to address the long-standing remediation question “how clean is clean?”
(Timeframe: 5-10 years)
v. Research is needed to develop and improve methods for remediation, reclamation or restoration of
damaged marine and freshwater habitats following oil spills. (Timeframe: 5-10 years)
vi. Research is needed on the efficacy and environmental impacts of conventional and new oil spill
remediation options, particularly in Arctic and freshwater ecosystems. (Timeframe: 5-10 years)

CHAPTER 7: PREVENTION AND RESPONSE DECISION MAKING

The best way to protect aquatic environments from the sometimes devastating impacts of spilled oil is to
prevent spills from happening in the first place. That is, reducing the likelihood of accidental spills is
always more effective than managing the risks (to the environment or to the economy) after a spill has
occurred. This principle is particularly true in the sensitive ecosystems where spills can cause catastrophic
or irreversible consequences, such as in the Arctic where industrial activities (e.g., offshore oil and gas,
mining), urban growth and climate-related changes to navigation routes are expected to increase tanker
traffic in the years ahead.

All oil spill strategies emphasize prevention as the prior emergency management activity. Effective
prevention combines an understanding of the science and technologies associated with oil operations and
potential oil spills with a clear understanding of the environment and conditions in which these activities
are taking place. For example, properly designed pipelines or tankers can be built to withstand anticipated
conditions (waves, wind, ice, etc.) that increase spill risk. Similarly, established procedures, proper
inspection and maintenance of equipment and training for extreme and adverse circumstances help reduce
the chances of a spill.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 34
High-Priority Research Need #6

Research is needed to improve spill prevention and develop/apply response decision support
systems to ensure sound response decisions and effectiveness.

i. A national guidance program for post-spill monitoring is needed to collect reliable, adequate,
credible and consistent information on the fate and effects of oil in the environment. This
program should be developed based upon consultations among industry, government,
Indigenous organizations and community stakeholders. (Timeframe: Within 5 years)
ii. Research is needed to develop methods to support the monitoring of oil spill impacts and the
fate of released oil. (Timeframe: 5-10 years)
iii. Research is needed to develop methods for the derivation of comprehensive mass balances for
spilled and recovered oil. (Timeframe: 5-10 years)
iv. Research is needed to develop modeling methods to simulate and optimize individual and
collective cleanup processes (e.g., booming, in situ burning, skimming, dispersion and
bioremediation) for supporting response decision-making. (Timeframe: Within 5 years)
v. Research is urgently required on development and demonstration of oil spill response decision
support systems, which can dynamically and interactively integrate monitoring and early
warning, spill modeling, vulnerability/risk analysis, response process simulation/control, system
optimization and visualization. (Timeframe: Within 5 years)
vi. Research investment is needed on trial tests and field validation of new prevention and decision-
making methods to demonstrate feasibility, increase confidence for implementation and improve
response capabilities. (Timeframe: Within 5 years)
vii. Research is needed to better quantify modeling uncertainties, evaluate their propagation and
mitigate their impacts on spill response decision-making. (Timeframe: 5+ years)
viii. Further research and development are desired on environmental forensics, remote sensing and in
situ measurement, early warning and diagnosis, and biological monitoring to improve spill
prevention and decision-making. (Timeframe: 5+ years)
ix. Special attention of the above research should be given to some emerging issues (e.g., diluted
bitumen, aging/subsea pipelines, railcars and the Arctic) to enhance effectiveness and
confidence of prevention and response strategies and decisions. (Timeframe: Within 5 years)

Prevention policies and measures are often best informed by the timely monitoring and analysis of the
causes and outcomes of spills when they do occur. Importantly, knowing the characteristics and
ecological features of an environment before a spill occurs is central to understanding how it has been
affected. Developing baseline data in areas where oil is transported should be an important priority for
research.

Indeed, despite the best prevention efforts, spills happen, and then making sound and timely decisions
about how to respond is a critical second line of defense. The Panel found that management strategies
should be in place to identify the lead decision-making agencies in the case of a spill and to present
clearly the steps to contain potential damage to human health, businesses and the environment. Decisions
concerning what to do following an oil spill can occasionally mean weighing the potential benefits of a
response against its possible harm or against the pros and cons of another approach altogether. Net
environmental benefit analysis (NEBA) provides a helpful framework and has been widely used for
supporting these decisions. The review also disclosed the limited research efforts in simulating, predicting
and optimizing cleanup processes (e.g., in situ burning, skimming and dispersion) individually and
collectively and evaluating their effects on response decisions. Inadequate decision support is one of the
major challenges that limit the efficiency of current response practices. Due to limited attention and
investment, existing decision support systems are rare and lack dynamic and interactive support from
other modeling tools (risk analysis, spill modeling, NEBA, process simulation, etc.) and field validation.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 35
In addition, uncertainty is a major hindrance to improving efficiency and confidence of decision-making.
These are especially true for the Arctic waters where the window of opportunity for the application of
some response measures is significantly short. The Panel also noticed that advances in monitoring and
information technologies such as remote sensing, geographic information systems, artificial intelligence
and visualization have provided a set of cost-effective and powerful tools that can play a more important
role in better addressing complexity and dynamics of spills and supporting sound response making and
operations.

High-Priority Research Need #7

Research and work are needed to update, refine and focus risk assessments of oil spills in Canada.

i. Follow-up relative risk assessments are needed to build upon the Transport Canada assessments of
marine spills, focusing on high-sensitivity areas. (Timeframe: Within 5 years)
ii. Research is needed to update and refine risk assessment methods to include such things as
credible spill scenarios, analyses of seasonal differences in fate, transport and effects of oil
(particularly for spills in winter) and the prediction of chronic toxicity. (Timeframe: Within 5
years)
iii. A comprehensive national database is needed to track the fate, behaviour and effects of various
types of oil spilled and the efficacy of current and emerging oil spill countermeasures over a range
of environmental conditions. (Timeframe: 5-10 years)
iv. Research is needed to expand species sensitivity distributions (SSDs) for acute and chronic
toxicity of oil to aquatic biota. SSDs should be expressed as measured concentrations of total
petroleum hydrocarbons and total polycyclic aromatic hydrocarbons. (Timeframe: 5-10 years)
v. Research is needed to extend models of chronic toxicity to a wider array of species and
environmental (temperature, salinity, etc.) and life history variables. (Timeframe: 10+ years)

CHAPTER 8: RISKS FROM OIL SPILLS

Learning from history is important to understanding what’s known about the risks posed by potential oil
spills in Canada and, most significantly, what needs further study. The Panel reviewed the circumstances
and outcomes for selected oil spill cases involving tanker accidents (e.g., the Arrow spill in 1970 and the
Exxon Valdez spill in 1989), a major ocean-rig blowout (i.e., the Deepwater Horizon Gulf of Mexico spill
in 2010), pipeline spills and train derailments that occurred in marine and fresh water in Canada and the
United States over the past few decades.

Chief among the Panel’s conclusions is that each case was unique in the combination of different
physical, chemical and biological factors at the spill location, as well as in the cleanup and recovery
measures used in the wake of each accident. These varied combinations of factors were critical for either
increasing or decreasing the overall impact for each spill.

Delays in responding to the spilled oil affected the outcome of all case studies examined. Indeed, despite
the obvious importance of weather, remote location and technological challenges facing each accident,
human error (individuals and organizations) played a dominant role in affecting the impact of the spills
across all case studies. Absent or inadequate planning, limited data analysis, inadequate training, poor
communication, insufficient personnel and equipment, poor or no information sharing, and lapses in
regulatory oversight were common to most, if not all, spill case studies.

From its case study review, the Panel found that the ability of aquatic ecosystems to recover from the
shock of an oil spill may be influenced by the presence of other longer-term environmental stresses (e.g.,
habitat degradation caused by urban development, fishing pressure or water pollution from sewage

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 36
discharges or agriculture). In most cases, the lack of pre-spill baseline data (i.e., information about the
natural environment and ecology of each area) hampered efforts to predict or monitor the long-term
effects of the oil spills on populations and communities of aquatic life. Similarly, monitoring following
spills was not conducted according to any standard or consistent national protocols. The Panel’s review of
risk assessments of oil spills in Canada revealed a number of challenges, notably the lack of readily
accessible data for use in the assessments and the need for increased sophistication of both exposure and
effects analyses. In many cases, even if data were accessible, they were extremely limited, particularly for
the Arctic and large portions of inland rivers, lakes and wetlands. The Panel found that the assumptions
used in the risk assessments sometimes were overly optimistic given the experience gained from oil spill
case studies. This was especially true for the spill response times assumed in the assessments.

CHAPTER 9: CONCLUSIONS

Crude oil spills are infrequent in Canada’s coastal or inland waters. But the consequences of these spills
into sensitive waterbodies can be profound. They can significantly affect not only the environment but
also the economy of affected areas as well as human health and safety.

Canada’s offshore oil and gas industry, meanwhile, is expected to grow. The production and transport of
unconventional oils, such as diluted bitumens and Bakken crude oil, are likely to increase. Spills of these
oils from offshore platforms, pipelines, tankers, rail and other sources will continue to pose risks to
Canadian aquatic environments and the communities that rely on them.

The Royal Society of Canada expert Panel prepared this report—based on a review of the science and
consultations with key stakeholders—to better understand what is behind these risks. Among the Panel’s
many conclusions is a long list of research needs, including seven key research areas that should become
top research priorities.

In particular, the Panel recommends that research needs to identify where most oil spills occur and why (e.g.,
pipeline spills into wetlands are more common than these spills into rivers; oil from truck spills are more likely
to enter storm sewers before reaching rivers; etc.). Researchers need to examine past spill response records,
current risk management processes and regulations to identify their weaknesses. Other critical knowledge gaps
include developing a better understanding of environmental sensitivities that affect the impact of spilled oil.
More research is also needed to understand how the type of oil, its source, the environment and the level of
preparedness of spill responders combines to influence spilled oil’s fate and effects.

These research gaps are significant. The data needed to assess oil spill risks in Canada are often either
absent or widely scattered among government agency, industry and academic sources. Information
needed to reliably assess the environmental sensitivity of areas at risk from oil spills is also very limited
for large portions of Canada. Input of traditional knowledge from Indigenous peoples and other interested
parties is needed.

Examples that demonstrate the need for more research are numerous. Many are documented in the Panel
report. While scientific advances have significantly reduced the threat of oil spills in Canadian waters
over the past few decades, much about the fate and effects of oil spills remains poorly understood.

To meet the research and oil spill response needs identified in the expert report, the Panel recommends
the conduct of coordinated multi-disciplinary research programs between industry, government and
academia to further study the effects of oil spills on various marine and freshwater ecosystems, including
wetlands. The program should also include Indigenous people for the provision of traditional knowledge
and expertise. The science from these studies will provide a much needed database on the interaction and
effects of spilled oil with its surrounding environment that will support science-based decision-making
following future spill incidents to protect our aquatic environment and its living resources.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 37
CHAPTER 1: INTRODUCTION

Abstract

This chapter describes the rationale and mandate for a Royal Society of Canada Expert Panel established
to provide a scientific review of our current knowledge and understanding of the behaviour and
environmental impacts of a range of crude oils, including diluted bitumen, which may be accidentally
released into Canadian marine or freshwater ecosystems.

To provide context for its evaluation of the risks associated with spills, an overview of global and
Canadian information is provided on the probable causes and predicted frequency and size of crude oil
spills into marine or fresh waters from exploration, production and transport operations. The global
frequency of large spills (>700 tonnes) from oil tankers has decreased significantly in the past four
decades and, fortunately, large spills due to tanker incidents in Canadian marine waters have been rare.
Return periods (i.e., time between spills of a given volume) in Canada range between about 40 and 240
years (depending on spill volume); however, these return periods may not provide a reliable basis for
predicting future occurrences in circumstances where the frequency and/or volumes of shipments are
increasing. Diffuse sources, such as natural seeps and runoff from land-based sources, account for the
majority of petroleum hydrocarbon inputs to oceans. Thus, while continued efforts to reduce tanker spills
are necessary, attention should also focus on sources such as urban runoff and recreational boating
because the spills are chronic and often occur in sensitive ecosystems. While freshwater spills associated
with land-based extraction of crude oils (including the extraction of bitumen) and the transport of
petroleum hydrocarbons (by pipeline, rail tankers and trucks) are expected to be at much smaller volumes
than their marine counterparts, they may cause significant damage because of their higher probability to
occur within populated areas and the proximity to water bodies lacking the dilution and dispersion
capacity typically found within the marine environment.

Notwithstanding the apparent relative infrequency of crude oil spills into aqueous environments in
Canada, the consequences of spills into sensitive marine and freshwater systems can be substantial, both
with respect to impacts on human health, safety and the environment and with respect to economic
impacts. Therefore, a primary focus of this Report is on the consequences of spills, no matter how low the
probability may be. This focus is commensurate with the key questions posed by the sponsors and
stakeholders.

1.1 Sponsors and Mandate of the Panel

The Expert Panel was established by the Royal Society of Canada (RSC) in response to a request from the
Canadian Energy Pipeline Association (CEPA) and the Canadian Association of Petroleum Producers
(CAPP) who recognized the need for a scientific review of our current knowledge and understanding of
the behaviour and environmental impacts of a range of crude oils, including diluted bitumens,
accidentally released into Canadian marine or freshwater ecosystems. The Panel was asked to address the
following questions:

1. How do the various types of crude oils, including diluted bitumens, compare in the way they
behave when mixed with fresh, brackish or salt waters under a range of environmental
conditions?
2. How do the various crude oils compare in their chemical composition and toxicity to organisms
in aquatic ecosystems?
3. How do microbial processes affect crude oils in aquatic ecosystems, thereby modifying their
physical and chemical properties, persistence and toxicity?

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 38
4. Are the research and oil spill response communities able to relate, with reliable predictions, the
chemical, physical and biological properties of crudes to their behaviour, persistence, toxicity and
ability to be remediated in water and sediments?
5. How should these scientific insights be used to inform optimal strategies for spill preparedness,
spill response and environmental remediation?
6. Given the current state of the science, what are the priorities for research investments?

1.2 Scope of the Panel’s Review


The terms ‘conventional’ and ‘unconventional’
The Panel considered accidental releases of
oils are related to the techniques used to extract
conventional and unconventional light, medium
them from reservoirs, not to their chemical
and heavy crude oils, including diluted bitumens.
composition. Conventional oils include liquid
The Panel evaluated reports on the environmental
crudes that flow in the reservoir and in
behaviours of these oils and the effectiveness of
pipelines. Such oils (e.g., condensates, light and
current and emerging oil spill response models
medium crude oils) are recovered from
and technologies available for use in the event of
traditional oil wells using established methods
accidental releases. Releases of biofuels or refined
like primary recovery and waterflooding.
petroleum products (e.g., gasoline, diesel, aviation
Unconventional oils are produced using
fuel, heating oil, etc.) were determined to be
unconventional methods, such as surface mining
outside of the scope of this Report by the project
of shallow oil sands, steam-assisted gravity
stakeholders, although spills of Bunker C and IFO
drainage of bitumen (SAGD) and horizontal
(intermediate fuel oil) were included for
drilling with hydrofracturing (‘fracking’) for
comparison.
recovery of light shale oils (‘tight oils’).
The scope of this Report was intentionally focused
on the accidental release of crude oils at the
exploration and production source or as they are transported to refineries. Thus, the transport of crude oil
by pipeline, ship, rail and truck was considered both with respect to the probability of accidental releases
and consequences in the aqueous environment. Unplanned releases from offshore platforms, including
subsurface blowouts, were included in light of the recent Deepwater Horizon (DWH) blowout and the
expected growth of Canada’s offshore oil and gas industry. The Panel focused on the Canadian
environment; however, it reviewed and considered the applicability of case studies from the United States
and other countries. Case studies were selected to represent different types of crude oil spilled into either
marine or freshwater systems. Efforts were also made to find case studies applicable to cold climates.

This Expert Panel Report is not expected to provide exhaustive overview of previous oil spill research and
case studies that have been cited in numerous existing reviews. While this information is clearly
considered in the response to the above questions, emphasis has been placed on the expertise and
experience of the individual Panel members to identify the most relevant and current literature in
scientific journals and academic, government and consultant reports on current policies and practices in
oil spill prevention and response, including the assessment of environmental impacts and remediation.
The Panel recognizes the existence of numerous review articles related to recent oil spill events and the
large amount of important research currently being conducted (e.g., research on diluted bitumen, and
studies related to Natural Resources Damage Assessment following the DWH spill in the Gulf of
Mexico). The conclusions and recommendations are based on a consideration of the large body of
literature available up to the date of writing.

1.3 Consultation with Stakeholders

The Panel consulted with representatives of CEPA and CAPP, government agencies in Canada
(Environment Canada, Natural Resources Canada, Fisheries and Oceans Canada, Alberta Innovates) and
the U.S. (National Oceanographic and Atmospheric Administration [NOAA]), private sector consultants,

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 39
first responders (e.g., Eastern Canada Response Corporation [ECRC-SIMEC]), non-government
organizations (e.g., Greenpeace), as well as individuals (e.g., academics). The formal consultations
included public forums with open online access that were hosted by the Panel in February (Calgary, AB)
and April (Halifax, NS) 2015 that involved presentations, as well as question and answer sessions. A third
Panel meeting in June 2015 included informal discussions with attendees at the 38th international Arctic
and Marine Oilspill Program (AMOP) technical conference that included a dedicated session on diluted
bitumens.

Several stakeholders emphasized that the Panel’s work could be used as supporting information for
government and industry policy and practice regarding spill prevention and emergency response. Some
stakeholders were interested in receiving relevant information regarding transportation routing (e.g., to
avoid or minimize risk to highly sensitive environments or species). Considerable interest was expressed
in gaining an understanding of the relative risks associated with the transportation and handling of
different types of crude oil. Some stakeholders wanted to know about the best available and emerging
technologies for spill cleanup, as well as monitoring requirements to track the effectiveness of cleanup.

Stakeholders expected the Panel to take a comparative approach in order to better understand the relative
impacts of conventional crude oils and diluted bitumen. The results of the comparisons among crude oil
types (and impacts in different marine and freshwater ecosystems) would then be useful when making
decisions regarding preparedness and appropriate spill response procedures.

The stakeholders requested that the Panel clearly summarize what is known and not known about the key
issues and make recommendations for actions and further research. Particular interest was expressed in
the performance of different cleanup technologies and the long-term effects of spills.

1.4 Context for Oil Spills in Canada

The Panel delivers the following brief overview of information on the causes, frequency and size of crude
oil spills into marine or fresh waters in Canada to provide context for its evaluation of the risks associated
with spills.

1.4.1 Spills from Oil Tankers in Marine Waters

Although this report considers the full range of Canadian aquatic ecosystems, most of the research,
response and media attention throughout the decades has focused on marine oil spills, particularly large-
scale accidents involving oil tanker vessels. The largest spill due to an oil tanker accident was the Atlantic
Empress in 1979 off the island of Tobago in the West Indies, where 287,000 tonnes of oil were released
after a collision with another vessel, followed by fire and an explosion that sank the tanker in deep waters.
No impact studies were carried out, so it is not known what quantity of oil burned or sank, but only minor
shore pollution was reported on nearby islands (ITOPF 2015a). The Amoco Cadiz grounding off the coast
of Brittany in 1978 resulted in the release of 223,000 tonnes of light Iranian and Arabian crude oil and
4,000 tonnes of bunker fuel into heavy seas—causing the formation of an emulsion that increased the
volume of pollutant by up to five-fold. Numerous shoreline types were affected and much of the oil
became buried in sediments and entrapped in salt marshes and estuaries (ITOPF 2015b). The spill of
37,000 tonnes of Alaska North Slope crude oil into Prince William Sound, AK, from the Exxon Valdez in
1989 was small in comparison, but led to the mortality of thousands of seabirds and marine mammals, a
significant reduction in population of many intertidal and subtidal organisms, and reports of long-term
environmental impacts (Spies et al. 1996). Public and political awareness about the risks involved in the
storage and transportation of oil and oil products as a consequence of this spill resulted in changes to
regulations, including: the enactment of the 1990 Oil Pollution Act by the U.S. Congress, which regulates
spills in the United States; regulations within Canada under the Canada Shipping Act (2001), which

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 40
regulates tankers; and the Canada Arctic Waters Pollution Prevention Act (1985, most recently amended
in 2014), which includes regulations regarding the design of Arctic class ships.

Fortunately, large spills due to tanker incidents in Canadian marine waters have been rare. In terms of
national involvement, one of the largest spills was in 1988 when the tanker Odyssey loaded with over
132,000 tonnes of North Sea Brent crude oil broke into two and sank 700 miles (~1,100 km) off the coast
of Nova Scotia beyond Canada’s Exclusive Economic Zone1 (ITOPF 2015c). Fire started on the stern
section as it sank and the surrounding oil caught fire. Due to rough weather, the Canadian Coast Guard
was only able to come within about 2 km of the vessel while it was on fire. No shoreline effects occurred
because of the distance of the spill from the coastline. Another well-known tanker accident off the
Canadian coast occurred in 1970 when the Arrow, hauling 9,500 tonnes of Bunker C fuel oil from Aruba
to Nova Scotia, encountered severe weather and gale force winds near Port Hawkesbury, NS, at the
entrance to Chedabucto Bay. The tanker ran aground on Cerberus Rock spilling most of the fuel oil cargo
and contaminating 75 miles of shoreline with residues of thick black sludge, which can still be found (Lee
et al. 2003; Owens et al. 2008). Within the scientific community, the Arrow spill was unique because a
section of the affected shoreline (Black Duck Cove) was intentionally left ‘untreated’ to enable scientific
assessment of natural recovery processes.

The frequency of large spills (>700 tonnes) from oil tankers has decreased significantly in the past four
decades. According to data from the International Tanker Owners Pollution Federation (ITOPF 2015d),
an average of 1.8 large spills per year occurred internationally in the period 2010-2014 compared to an
average of 24.5 large spills per year between 1970 and 1979. One large spill was recorded in 2014 - the
sinking of a tanker in the South China Sea loaded with a cargo of about 3,000 tonnes of bitumen. The
U.S. Department of the Interior’s Bureau of Ocean Energy Management (BOEM) and Bureau of Safety
and Environmental Enforcement (BSEE) produced a comprehensive summary report on the occurrence
rates for offshore oil spills from U.S. Outer Continental Shelf (OCS) Platform and Pipeline Spill Data
(1964 - 2010), Worldwide Tanker Spill Data (1974 - 2008), and Barge Spill Data for U.S. waters (1974 -
2008) (Anderson et al. 2012). According to this report, the rate of offshore oil spills in the U.S. has been
declining since 1994, due mostly to major regulatory changes in the early 1990s (arising from the Exxon
Valdez spill) that substantially eliminated the use of single-hull tankers by requiring double hulls or their
equivalent. In 2013, a report by the Canadian Tanker Safety Panel Secretariat (TSPS) published two
tables produced by GENIVAR, a professional services firm, showing the risk calculations for the
probability and potential impacts of ship-source oil spills in Canada (Table 1.1) and internationally
(Table 1.2). The significant figures in these two tables are from those reported by the data sources. The
return periods (i.e., time between spills of a given volume) presented in these tables may not be a reliable
basis for predicting future occurrences in circumstances where the frequency and/or volumes of
shipments are increasing (e.g., along the southern British Columbia coastline if the proposed
TransMountain Pipeline Expansion Project, with associated tanker traffic, is approved).

1
The exclusive economic zone (EEZ) is an area of the sea adjacent to and beyond the territorial sea, extending out to 200
nautical miles from the baseline. Within the EEZ, a coastal state has sovereign and jurisdictional rights over exploration and
management (e.g., scientific research and protection of the marine environment) and economic exploitation of living and non-
living resources in the waters above the seabed, in the seabed and beneath the seabed.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 41
Table 1.1 Spill frequency estimates for ship-source spills occurring in Canada, calculated using only previous
Canadian spill occurrences. Data from TSPS (2013)
Return period, years a
Spill volume, m3 10-100 100-1,000 1,000-10,000 > 10,000
Crude oil –b – – –
Refined Cargo oil 1.7 10.0 – –
Bunker oil 0.5 1.7 – –
Total 0.4 1.4 – –
a
estimated average number of years between spills
b
probability could not be estimated due to the lack of spills in Canada in this category in the previous 10 years,
indicating that the probability of a spill in this size range and category is remote.

Table 1.2 Spill frequency estimates for ship-source oil spills in Canada, based on international and Canadian
data. Data from TSPS (2013)
Return period, years a
Volume, m3 10-100 100-1,000 1,000-10,000 > 10,000
Crude oil 46.4 69.2 51.6 242.3
Refined Cargo oil 1.7 10.0 42.2 –b
Bunker oil 0.5 1.7 154.8 –
Total 0.4 1.4 20.2 242.3 c
a
estimated average number of years between spills
b
probability could not be estimated due to the lack of spills in the previous 10 years in this category, indicating that
the probability of a spill in this size range and category is remote
c
i.e., a crude oil spill of >10,000 tonnes in Canadian waters is predicted to occur once every 242 years, based on
the occurrence of only two such spills worldwide in the past 10 years, both from single-hulled vessels (the Tasman
Spirit spill of ~30,000 tonnes in 2003 in Pakistan, and the Hebei Spirit spill of ~10,500 tonnes in 2007 in South
Korea).

In addition to the estimates above, ITOPF (2014) published actual oil tanker spill statistics, which show a
declining trend in the number of tanker spills greater than 700 tonnes from 1970-2014 (Figure 1.1).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 42
Figure 1.1 Number of large oil tanker spills (>700 tonnes) occurring worldwide from 1970 to 2014 (grey bars),
and average number of spills per decade (red line). Adapted from ITOPF (2014)

The 1989 Exxon Valdez incident increased public and political awareness about the risks involved in the
storage and transportation of oil and oil products. This awareness resulted in changes to regulations,
including: the enactment of the 1990 Oil Pollution Act by the U.S. Congress, which regulates spills in the
United States; regulations within Canada under the Canada Shipping Act (2001) (http://laws-
lois.justice.gc.ca/eng/acts/s-9/), which regulates tankers; and the Canada Arctic Waters Pollution
Prevention Act (1985, most recently amended in 2014) (http://laws-lois.justice.gc.ca/eng/acts/A-12/),
which includes regulations regarding the design of Arctic class ships.

1.4.2 Accidental Releases from Offshore Oil and Gas Exploration and Production

Spills, blowouts and malfunctions may occur during any offshore oil and gas exploration activity. As
defined by the U.S. NOAA, a blowout occurs when operators of a drilling rig are unable to control the
flow of oil, gas or other fluids from the well, causing it to be released into the underground formation,
marine environment and/or atmosphere. Until recently, the risk of blowouts was considered low in the oil
exploration and development industry and thus of minor concern. For example, from 1979 to 1998,
19,821 wells were drilled within the Gulf of Mexico, with only 118 wells resulting in uncontrolled flows
or blowouts - a 0.6% occurrence rate (API 2009). The reported blowouts were not considered a significant
source of hydrocarbon releases to the environment due to effective blowout prevention (BOP) systems
and the fact that they often sealed naturally and ceased flowing within a matter of hours or days. In 2004,
the Canada Nova Scotia Offshore Petroleum Board (CNSOPB) predicted a 1-in-1800 chance per year of
having any sort of deep water blowout off the continental shelf during exploratory drilling, with the
probability of shallow water gas blowouts without a release of oil having a three- to four-fold higher
probability of occurrence (Hurley and Ellis 2004). However, the DWH blowout in the Gulf of Mexico,
which resulted in the continuous discharge of petroleum gas and crude oil into surrounding waters (4.2
million bbl [~600,000 m3] over an 87 day period) and impacts on the Gulf shoreline and salt marshes, has
changed the world’s perception of the environmental risk associated with blowouts, with accompanying
changes in regulatory requirements. For example, Canada’s National Energy Board (NEB) is mandating
same-season relief well capability or equivalent for the Arctic because the tracking and cleanup of oil
spills under ice are major challenges in terms of response.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 43
Only one major oil well blowout has occurred in Canadian offshore waters - the 1984 Uniake G-72 gas
and condensate well off Sable Island approximately 150 nautical miles (~275 km) from Halifax, NS
(Boudreau et al. 1999). In 2004, due to equipment failure, about 170 tonnes of crude oil was accidentally
released from the Terra Nova production, storage and offloading platform on the Grand Banks, about 340
km east-southeast of St. John’s, NL. The primary risk was to seabirds present in relatively high densities
in the area of the spill. It was estimated that about 10,000 seabirds were killed by the spill (Wilhelm et al.
2007) as the oil slick was dispersed by natural physical processes.

Diffuse sources (natural seeps and runoff from land-based sources) account for the majority of petroleum
hydrocarbon inputs to oceans (NRC 2003). The total input of petroleum into the sea worldwide from all
sources over the period 1990-1999 was estimated by the NRC as approximately1.3 million tonnes per
year. Natural petroleum seeps and runoff from land-based sources contributed 46% and 11% of petroleum
hydrocarbon inputs, respectively. Tank vessel spills contributed about 8% of total inputs, while
operational discharges (e.g., bilge discharges and cargo washings) contributed about 3% of inputs.
Furthermore, most spills from tank vessels are relatively small. Data from the ITOPF on oil tanker spills
show that of nearly 10,000 incidents, 81% resulted in releases of less than 7 tonnes (ITOPF 2015d). More
recent comparisons of relative contributions of petroleum hydrocarbons to the world’s oceans confirm
earlier conclusions (Figure 1.2).

Total spillage from tankers Routine annual maintenance


(1970 – 2010) (e.g., bilge cleaning)
6.7 million m3 0.52 million m3

Natural oil seeps (annually)


0.70 million m3

Deepwater Horizon (2010)


0.78 million m3

Exxon Valdez spill (1989)


0.04 million m3
= 100,000 m3

Figure 1.2 Volumes of oil released to marine waters from natural sources and spills. Data from NRC (2003);
Mackenzie (2011)

Examination of the proportion of oil released into North American coastal and offshore marine waters
from natural and human activities reveals that natural seeps and land-based sources accounted for an even
greater proportion of inputs (NRC 2003) (Figure 1.3). The NRC estimated total inputs to North American
waters were 260,000 tonnes per year over the period 1990-1999, with natural seeps contributing 61% of
total inputs. Extraction of petroleum contributed about 1% of total inputs, largely from produced waters.
Transportation of petroleum, including pipeline, tank vessel and coastal facility spills, accounted for 3.5%
of inputs. The category “Consumption of Petroleum” included land-based river and runoff, recreational
marine vessels, operational discharges and jettisoned aircraft fuel; these accounted for 32% of total inputs.
The NRC noted that urban runoff and recreational boating require attention because the spills are chronic
and often occur in sensitive ecosystems.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 44
Figure 1.3 Average annual releases of oil (in kilotonnes) into North American marine waters from various
sources (1990-1999). Data from NRC (2003)

1.4.3 Inland Freshwater Spills

Compared to marine oil spills, inland oil spills have received less attention, but freshwater spills are
common, with more than 2,000 oil spills, on average, taking place each year in the inland waters of the
continental United States (Owens et al. 1993). Although freshwater spills tend to be smaller volumes than
their marine counterparts, they can have a greater potential to pose risks to the environment because of the
greater likelihood that they occur within populated areas close to waterbodies with much less dilution and
dispersion capacity, and involve shorelines that are often immediately adjacent to or directly impacted by
the spill.

There are several recent examples of the substantial impacts of spills into inland waters. The spill of about
1,500 tonnes of Bunker C heavy fuel oil into Wabamun Lake, AB, in 2005 (due to a train derailment)
resulted in residents being advised to avoid all use of lake and well water, and in impacts to habitat for
numerous aquatic organisms and waterfowl (Thormann and Bayley 2008; Martin et al. 2014). The release
of 26,000 bbl (1 bbl = 3,100 tonnes) of dilbit into Talmadge Creek, a tributary of the Kalamazoo River
near Marshall, MI, in July 2010, resulted in the largest inland oil spill and one of the costliest spills in
U.S. history. Another major, high visibility inland spill occurred in March 2013, when an Exxon-Mobil
pipeline carrying Canadian Wabasca heavy crude from the Athabasca oil sands ruptured in Mayflower,
AR, releasing 5,000 to 7,000 bbl (600 to 880 tonnes) of oil mixed with water into a residential
neighbourhood. The oil eventually flowed through storm drains into Lake Conway, a recreational fishing
lake.

Data on spills into fresh water in Canada were available primarily from the Transportation Safety Board
of Canada (TSB) and national provincial regulatory agencies, such as the National Energy Board (NEB),
the Alberta Energy Regulator (AER) and the AER’s predecessor, the Energy Resources Conservation
Board (ERCB). The Panel found apparent differences in the level of data processing and archiving
between government agencies within Canada. For example, spill reports in the Government of
Saskatchewan Petroleum and Natural Gas Spill Report Directory (Government of Saskatchewan 2015)
were not filed in a manner amenable to searching by keyword in order to focus on volume, type of spill or
receiving environment (the only search criterion was spill location).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 45
1.4.4 Spills from Pipelines

The volumes of releases of crude oil from pipeline accidents are usually small and releases are rarely into
waterbodies. The TSB reported that over the 2004-2013 period, 14 pipeline accidents resulting in releases
of crude oil occurred in Canada. Eight of these releases were less than 1 m3 (<1 tonne), two were between
1 and 25 m3, three were between 26 and 1,000 m3, and one release was over 1,000 m3 (TSB 2013b). The
AER reported that the number of crude oil pipeline incidents per 1,000 km of pipe ranged from 1.5 to 3.1
during the period 1990-2000 and from 1.0 to 1.9 during the period 2001-2012 (AER 2013). The AER
stated that the stable spill frequencies in recent years may indicate that current practices may not result in
further improvements; rather, new technologies or management strategies would be required to achieve
significant spill reductions.

Although the incidence of crude oil releases from pipelines into fresh water have been relatively rare,
spills entering rivers or wetlands have caught the public’s attention due to the potential ecological damage
on sensitive sites. Some of these spills are summarized in Table 1.3 to provide context regarding volume,
the type of receiving environment and environmental effects (if noted by the investigators).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 46
Table 1.3 Examples of recent larger-volume pipeline spills in Alberta
Volume Type of
Date and and type of Receiving
Location spilled oil Environment Environmental Effects Reference
January 2001, 3,800 m3 Permanent 3,760 m3 recovered; no TSB (2001)
Enbridge crude oil slough effects noted
pipeline at
Hardisty, AB
April 2011, 4,500 m3 Beaver ponds Beaver, amphibian, bird and ERCB (2013)
Plains Midstream crude oil and muskeg small mammal mortalities;
pipeline near reclamation required
First Nations including re-vegetation
Community of
Little Buffalo,
AB
May 2012, Pace 800 m3 Wetlands No wildlife or aquatic life AER (2014a)
Oil and Gas crude oil mortalities noted; remediation
pipeline near efforts reported as being
Rainbow Lake, effective with vegetative
AB regrowth noted
June 2012, Plains 460 m3 Red Deer River Effects on water supplies of AER (2014b)
Midstream crude oil two communities and on
pipeline near recreational uses; effects on
Sundre, AB wildlife, soils and riparian
vegetation
July 2015, Nexen 5,000 m3 Muskeg Duck and frog mortalities; AER news release
Energy pipeline bitumen adjacent to the concerns focus on effects of https://www.aer.ca
at its Long Lake emulsion pipeline right- high-salinity process water in /compliance-and-
operations, AB of-way the spilled oil: water emulsion enforcement/nexen
-long-lake

1.4.5 Other Releases from Production Sources into Fresh Water

Crude oil releases to fresh water at production sources can be due to blowouts or ‘flow-to-surface’
incidents associated with hydraulic fracturing (‘fracking’) or with in situ oil sands production facilities.
Well blowouts can involve releases of oil, produced water, frac fluids and/or gases. A review of
investigation reports by the AER and previously the ERCB showed that the main consequences were to
workers and public safety due to fire, explosion and toxic gases (particularly hydrogen sulphide gas).
Effects on soils and vegetation in the immediate vicinity were also noted. Only one case was found (for
the period 2006-2015) where release to a waterbody was noted; in this case, wellbore fluids (produced
water and diesel fuel) entered an unnamed watercourse about 100 m from the well (ERCB 2011). The
volumes and/or consequence of the release to the watercourse were not discussed.

A documented release of crude oil to a surface waterbody due to a fracking incident was not found in the
record of reports released by the AER or ERCB. The ERCB investigated one incident that resulted in a
‘misting’ of released fluids on trees and soils; however, no release into water was noted (ERCB 2012).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 47
The release was caused by hydraulic fracturing operations affecting a nearby producing oil well, resulting
in a release of frac and formation fluids, produced water and natural gas.

The use of high-pressure steam to release bitumen from oil sands formations in situ can result in release
of oil to the surface environment. For example, a series of four bitumen releases occurred at the Canadian
Natural Resources Ltd (CNRL) Primrose high-pressure cyclic steam stimulation operations near Cold
Lake, AB, from 2013-2014. The events were related to the use of high-pressure steam in a situation where
vertical hydraulically-induced fractures could propagate, the cementing of wellbores was inadequate, and
natural fractures and faults contributed additional pathways to the surface (Independent Panel Review
2014). Estimates of the total volume released ranged from about 1,200 to 1,900 m 3 (volume reports varied
between CNRL, AER and third-party reviewers). Beaver, migratory bird, amphibian and small mammal
mortalities occurred within the affected wetland complexes. A 20-ha portion of a lake had to be drained to
remove bitumen and re-filled. The AER prohibited future high-pressure steaming at specific CNRL
Primrose locations. CNRL reduced steam volumes and implemented enhanced monitoring and response
to any further bitumen releases pending a final report from the AER.

1.4.6 Spills Due to Train Derailments

The recent increase in transport of crude oil by rail has resulted in an accompanying increase in rail-
related spills. The TSB’s most recent report on railway occurrences stated that over the 2004-2014 period,
an increase in accidents involving release of crude oil was concurrent with an increase in shipments of
crude oil by rail, from 500 car loads in 2009 to 160,000 car loads in 2013 (TSB 2015a). Five rail
accidents resulted in the release of crude oil in 2013, including the Lac Mégantic derailment (see below).
One accident involved a release of crude oil in 2014.

According to the U.S. Pipeline and Hazardous Materials Safety Administration (PHMSA), the United
States has also experienced an increase in train-related oil spills. Between 1975 and 2009, no reported oil
spills occurred during eight of those years, and spills of just one gallon or less were reported in five other
years. However, more recently, data compiled by PHMSA reported a total of 3,700 tonnes of crude oil
spilled from rail cars in 2013, greater than the 2,600 tonnes reported during the previous 37 years
combined. A significant portion of the increased crude-by-rail transportation can be related to shale oil
production in North Dakota, which has greatly expanded its output since the discovery of a new oil field
in 2006 (the Parshall Oil Field producing from the Bakken and Three Forks Formations in the Williston
Basin) (LeFever 2008)

Bakken crude oil produced in North Dakota was involved in three crude-by-rail spills during 2013,
including the derailment in Lac-Mégantic, QC, on July 6, where a subsequent explosion and fire killed 47
people. An estimated 100 m3 of spilled crude oil entered Mégantic Lake and the Chaudière River via
surface flow, underground infiltration and sewer systems (TSB 2014b). In November 2013, derailment of
26 tank cars in Aliceville, AL, resulted in the release of nearly 2,400 tonnes of Bakken crude oil; the U.S.
National Transportation Safety Board (NTSB) noted that the crude oil was released into a wetland (NTSB
2014). The third major spill occurred on December 30 near Casselton, ND, where several tank cars
ruptured when a crude oil unit train derailed after striking another derailed freight train, spilling an
estimated 1,300 tonnes of crude oil (NTSB 2013). A post-accident fire created dense, toxic smoke that
forced a temporary evacuation of the town (NTSB 2014). The NTSB preliminary report did not note any
release to surface waterbodies from the Casselton spill, and the online NTSB docket did not include any
references to reports regarding effects on waterbodies (NTSB 2015).

As reported in the public media, releases of crude oil due to rail accidents continued in 2015. In several
cases, the oil caught fire and towns were evacuated. Most spills occurred on land rather than in aquatic
environments, but one entered a river. Examples during the first six months of 2015 include the
following:

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 48
 16 February – Canada – A Canadian National Railway (CN) freight train transporting crude oil
derailed 80 km south of Timmins, ON. Of the 100 cars in the train, 29 derailed and seven crude
oil tank cars caught fire. No injuries occurred and the burning oil was contained to the area
(Mangione 2015).
 16 February – U.S. – A CSX freight train hauling 109 tank cars of Bakken crude oil derailed near
Mount Carbon, WV. At least seven crude oil tank cars caught fire, forcing the evacuation of
about 200 people, one injury and no deaths (Van Pelt 2015). About 685 m3 of crude oil were
subsequently recovered by CSX and about 9,070 tonnes of soil were removed and shipped for
disposal (Raby 2015).
 5 March – U.S. – A BNSF oil train derailed in a rural area near Galena, IL. Of the 105 cars being
transported, 21 containing Bakken formation crude oil left the track and caught fire. No injuries
were reported. The U.S. EPA reported that sampling of surface water did not detect any oil in the
Mississippi or Galena Rivers (EPA 2015).
 7 March – Canada – A CN freight train derailed near Gogama, in northern Ontario. Forty of the
100 cars derailed, with five entering the Makami River and seven carrying crude oil catching fire
and burning for several days (TSB 2015b). Cleanup crews deployed booms to soak up floating
oil, as well as vacuum trucks to recover oil residues in the river. Thousands of tonnes of oily
snow were removed. A drinking water advisory issued for Minisinakwa Lake was lifted on June
10, 2015, after analyses confirmed that hydrocarbon concentrations did not pose a risk to human
health (Gillis 2015).
 6 May – U.S. – A BNSF train derailed near Heimdal, ND, which ignited a crude oil fire in six
tanker cars and forced the evacuation of approximately 40 nearby residents. No injuries or
fatalities were reported (Riordan Seville et al. 2015).

1.4.7 Spills Due to Truck Transport Accidents

The total volume of crude oil transported by truck in Canada is an order of magnitude smaller than
transport by pipeline, but the number of incidents per million m3 of crude transported is four-fold greater
(Table 1.4). The rate of incidents per volume transported was lowest for rail transportation. This disparity
remained consistent in 2014, when only one crude-by-rail spill occurred while transporting a total volume
of 9,339,576 m3 of oil (significant figures are those reported by the data sources), an incident rate of 0.11
per million m3 (data from National Energy Board [2015] and TSB [2015a]) versus a thirteen-fold greater
rate for oil transport by truck in 2011 (Table 1.4).
a,b
Table 1.4 Number of accidents, volume shipped and accident rate by mode of crude oil transport . Data from
Young (2014)
Year Mode Number of Volume Shipped (m3) Rate
Accidents (per million m3)
2011 Truck 45 31,459,085 1.43
2011 Pipeline 259 736,285,714 0.35
2012 Rail 1 3,908,266 0.26
a, Volume data from Statistics Canada and Transport Canada; accident numbers from Transport Canada for rail
and truck and from provincial governments and NEB for pipeline. Volumes were estimated using crude oil density of
0.9 kg/l. There were no in-transit rail accidents in 2011; therefore, 2012 data were used for comparison. Pipeline
volume includes some diluent.
b. The number of significant figures in the table are those reported by Young (2014)

Data from the U.S. indicate a similar pattern. The volume of oil spilled per distance transported has been
greatest for tanker trucks among all modes of transport since 1996 (Figure 1.4). However, this metric has

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 49
decreased over the same time period for all spills from all modes of transportation, including oil spills
from trucks.

Figure 1.4 Average volumes of crude oil and petroleum product spilled per distance transported by all modes in
the United States. Notes: Pipelines include onshore and offshore pipelines. The time periods were chosen based on
the available annual data for both spill volume and ton-miles. The values for each time period are averages of
annual data for each six-year period. Source: Frittalli et al. (2014)

Although the incident rate is highest for truck transport, the average spill volume per incident is much
smaller and mainly on land. Unlike other modes of transport, trucks are primarily used to transport oil for
relatively short distances. However, a recent increase in truck oil shipments has occurred, which is not
reflected in the historic data shown in Figure 1.4. Shipment of oil by truck from shale formations in North
Dakota and oil sands in Canada to U.S. refineries increased by 38% between 2011 and 2012
(Christopherson and Dave 2014). Risks to public health and safety from a tanker truck spill would be
greater in the event of a fire and explosion if the spill occurred close to inhabited areas, and risks to the
environment would be greater if the spill was released directly into waterbodies. For example, in 2013 a
tanker truck carrying 35 m3 of jet fuel tipped over into Lemon Creek, BC (a tributary of the Slocan River)
resulting in an evacuation order affecting about 1,500 people, a “do not drink” water order that lasted a
week, some fish and bird mortalities, and effects on water and sediment quality (CBC 2013; Executive
Flight Centre 2014). While this incident did not involve crude oil, it illustrates the effects of hydrocarbon
spills directly into waterbodies. A release of crude oil would be expected to have longer-term impacts due
to the persistence of the non-volatile components of the spill in water, sediments and biota.

1.4.8 Summary of Crude Oil Spills in Canada

Spills of crude oil into marine or freshwater systems in Canada from oil production facilities, tankers,
pipelines, rail and truck transport are infrequent, and the probability of spills decreases with increasing
spill size (Table 1.5). However, as pointed out by SL Ross Environmental Research (2014), while
statistics for spills into the marine environment are readily available in Canada, statistics for inland spills
into fresh water are not. The difficulty in locating inland spill data produced the “no data” results for rail
and road incidents in Table 1.5. The Panel’s review confirmed the difficulty of obtaining clear and

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 50
reliable data on crude oil spills into fresh water. For example, the information presented in the sections
above on spills from production facilities, rail and truck transport had to be gleaned from details within
statistical reports on all incidents, whether involving oil transport or not (e.g., the TSB [2015] report on
railway occurrences), by laborious searching through individual investigation reports (e.g., individual
TSB or AER reports) and/or from media reports. Even when reports were available on spill incidence and
volume (e.g., the AER [2013] report on pipeline performance), spills of crude oil were not always
distinguished from refined product spills; instead, statistics were presented on “hydrocarbon liquids”.

Table 1.5 Annual frequency of oil spills in Canada by source and volume. Source: SL Ross Environmental
Research (2014)
Spill Volume (m3)
Source 1,000-
<1 1-10 10-100 100-1,000 > 10,000
10,000

Offshore
exploration and 20.9 1.7 0.4 0 0 0
productiona

Vessel Spillsb
Crude ndd nd 0.02 0.01 0.19 0.004
Refined nd nd 0.60 0.10 0.024 0.00
Fuel nd nd 1.90 0.60 0.01 0.00
Total nd nd 2.52 0.71 0.05 0.004

Pipelinesc
Release into 0.56 0 0.20 0 0 0
water
Total 0.75 3.8 1.89 0.94 0 0

Rail nd nd nd nd nd nd
Road nd nd nd nd nd nd
a
. 2004 to 2013. Canada-Nova Scotia Offshore Petroleum Board (http://www.cnsopb.ns.ca/environment/incident-
reporting) and http://www.cnsopb.ns.ca/environment/incident-reporting
b
. WSP Canada Inc. (2014).
c
. 2008 to April 2013. https://www.neb-one.gc.ca/sftnvrnmnt/sft/dshbrd/index-eng.html
d.
nd, no data; information was not publically available or was not immediately accessible.

Notwithstanding the apparent relative infrequency of crude oil spills into aqueous environments in
Canada, the consequences of spills into sensitive waterbodies can be substantial, both with respect to
impacts on human health, safety and the environment and with respect to economic impacts. Therefore, a
primary focus of this Report is on the consequences of spills, no matter how low the probability may be.
This focus is commensurate with the key questions posed by the sponsors and stakeholders.

1.5 Organization of This Report

The relationships among the chapters of this report and the key questions posed by the sponsors and
stakeholders are illustrated in Table 1.6. There is overlap in the subject matter among the chapters but the
Panel strove to avoid duplication. In some cases, topics are discussed from a different perspective, or in
more or less detail, depending upon the primary focus of the chapter. Chapter 8 considers all of the major

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 51
topics within a risk framework and uses case studies to examine the importance of physical, chemical and
biological factors, as well as cleanup and recovery measures in determining short-term and long-term risk
of spills. The final chapter (Chapter 9) presents the Panel’s recommendations.
Table 1.6 Relationship between chapters and key questions
Question 4:
Question 3:
Question Question 2: Relationship Question 6:
Microbial
1: Chemical Among Crude Question Spill
Processes
Composition Oil Behaviour, 5: Prevention,
Chapter Affecting
Behaviour and Toxicity Toxicity, Research Preparedness
Crude Oil
of Crude of Crude Oil Characteristics Priorities and
Toxicity and
Oil Types Types and Response
Characteristics
Remediation
2:
Composition,
Properties X X X X
and
Behaviour
3: Effects of
Environment
on Oil Fate X X X
and
Behaviour
4: Toxicity
and
X X X X X
Ecological
Relevance
5: Modeling
Spill X X
Behaviour
6: Spill
Response X X
Options
7: Prevention
and Response
X X
Decision
Making
8: Risks from
X X X X X X
Oil Spills

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 52
CHAPTER 2: CHEMICAL COMPOSITION, PROPERTIES AND BEHAVIOUR OF SPILLED OILS

Abstract

The chemical composition of petroleum1 is of utmost importance for understanding oil spills in aquatic
environments because the chemistry of the oil dictates its physical properties (e.g., density and viscosity),
behaviour (e.g., spreading, sinking, dispersion), biological impacts (e.g., toxicity, susceptibility to
biodegradation) and ultimate fate in the environment. Thus, petroleum chemistry is important to oil spill
responders for predicting hazards, such as fire and explosion, and for selecting suitable cleanup and
remediation approaches. It is also important to regulators for forensic purposes, monitoring and
determining acceptable endpoints for spill remediation.

The petroleum hydrocarbons transported within Canada range from ultra-light condensates and light oils
to heavy oils and bitumens, as well as blends like diluted bitumens and heavy fuel oils. Each oil is a
complex mixture of thousands of chemicals, not all of which are hydrocarbons, and every oil has a unique
chemical ‘fingerprint’. Conventionally, petroleum is separated into four major fractions for analysis:
saturates, aromatics, resins and asphaltenes. Whereas the structures of many saturated and aromatic
petroleum compounds are known and have been studied individually or within an oil matrix, the
chemically diverse, high molecular weight resins and asphaltene fractions have resisted characterization.
Additional minor constituents in oil may include sulphur (e.g., as hydrogen sulphide gas), naphthenic
acids, metals and minerals.

Spilling oil onto water progressively changes its chemical fingerprint through physical, chemical and
biological processes collectively called ‘weathering’. As oil spreads on the water surface, small molecules
evaporate, others may be oxidized by sunlight, and the oil may form emulsions with water. In the water
column oils may disperse as droplets and/or sink and/or form ‘tar balls’ after interacting with suspended
particles, light components may dissolve in the water and others may be selectively biodegraded by
microbes. Oil that reaches the shoreline or sediments may become sequestered and/or re-emerge over
time, but oil interactions with ice are currently not well known. Measurable changes in the ‘fingerprint’ of
spilled oil over time reflect its initial chemistry and the magnitude of weathering processes and therefore
can be used to monitor remediation progress and determine endpoints. Importantly, just as each crude oil
is chemically unique, each oil spill reflects the specific environmental conditions at the spill site acting on
a particular oil over a given time period. The diverse petroleum types form a continuum of chemical
compositions, physical properties and spill behaviours.

To provide background for later chapters on oil fate (Chapter 3), toxicity (Chapter 4), modeling (Chapter
5) and remediation of oil spills (Chapter 6), as well as spill response decision-making (Chapter 7), this
chapter summarizes the common chemical components of crude oils, describes a suite of crudes
representing the compositional range of oils commonly transported in Canada, and briefly outlines the
general behaviour of oil spilled in marine and freshwater environments. Further research is needed to
fully characterize the properties and spill behaviours of emerging oil types, including diluted bitumen
blends and unconventional oils such as shale oil.

1
The terms ‘petroleum’ and ‘crude oil’ are considered here to be synonyms and are used interchangeably.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 53
2.1 Chemical Composition of Oils

Crude oils are naturally-occurring complex


mixtures of thousands of individual compounds, Crude oils are highly complex, diverse mixtures
but primarily they comprise hydrocarbons of chemicals. Each oil has a unique composition
(molecules consisting of only carbon and or ‘chemical fingerprint’ that usually consists of
hydrogen) and lesser proportions of compounds four major classes of chemicals: Saturates,
containing heteroatoms (e.g., nitrogen, sulphur Aromatics, Resins and Asphaltenes (SARA) and
and/or oxygen) in addition to carbon and sometimes minor constituents like metals,
hydrogen. Commonly, small amounts of metals, sulphur, solids and water.
minerals and sometimes inorganic sulphur are also
present. The majority of petroleum components
are derived from organic matter from ancient
aquatic plants, animals and microbes, and their proportions in the oil reflect the source(s) and geological
history of that oil during its formation.

2.1.1 Four Major Chemical Fractions of Crude Oil

Most crude oil constituents can be divided into four major chemical classes known by the acronym
SARA: Saturates, Aromatics, Resins and Asphaltenes, as described below and illustrated in Table 2.1.
Historically, these groups are defined as solubility fractions by liquid-solid chromatography, as explained
in Appendix A (Figure A1). Appendix A also presents an overview of analytical methods commonly
used to chemically characterize oils. Together these fractions contribute to the bulk properties of
petroleum, such as viscosity, specific gravity and susceptibility to biodegradation, as discussed in detail
below (Section 2.2).

It is important to note that Table 2.1 presents only a few examples of the thousands of possible structures
of petroleum components. Indeed, the presence and positions of one or more alkyl groups (e.g., –CH3, –
C2H5) on parent structures, whether saturated or aromatic, can produce myriad ‘isomers’ (chemicals
having the same chemical formula but different structures) and ‘homologous series’ of similar chemicals
that differ by the number of alkyl substituents. Isomers and members of homologous series may have very
different adsorption properties, water solubility and susceptibility to biodegradation. A paraffin with four
carbon atoms can exist in only two isomer forms, but one with 14 carbon atoms theoretically can exist as
60,523 isomers (Speight 2014). Multiple isomers cause technical problems for peak resolution by gas
chromatography (GC) and contribute to the unresolved complex mixture (UCM) or ‘hump’ in gas
chromatograms (Appendix A).

2.1.1.1 Saturates

This chemical class is typically the major


component of petroleum. It exclusively comprises
hydrocarbon molecules having single carbon- The saturate fraction is considered the least
carbon bonds, with all remaining bonds being toxic of the four major petroleum fractions and
saturated with hydrogen atoms (i.e., no double- or the most readily biodegradable, with the
triple-bonded carbon). Three subclasses are exception of steroids and hopanoids that tend to
defined within this fraction: paraffins and resist biodegradation.
isoparaffins having straight or branched chain
structures, respectively; naphthenes
(cycloparaffins, also called alicyclic hydrocarbons) having one or more saturated ring structures; and
fused-ring aliphatics, including hopanoids and steroids, having large, complex structures combining
paraffinic and naphthenic structures (Table 2.1). Olefins (partially unsaturated hydrocarbons), such as
alkenes and alkynes, are scarce or absent in petroleum (Speight 2014).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 54
Petroleum paraffins range from gases and vapours of one to five carbon atoms (C 1-C5) under normal
pressure (e.g., methane gas, CH4) to liquids of C5-C16, to crystalline waxes of C>20. Normal (n-)alkanes
have straight chains with no branching (i.e., no alkyl substitutions on the hydrocarbon backbone) and
therefore have the general formula CnH2n+2. iso-Alkanes are branched structures with alkyl substitutions at
single or multiple points on the backbone, creating numerous series of isomers having the same molecular
weight but different structures.

In general, n-alkanes <C5 are highly volatile (and therefore not likely to remain in the water column for
biodegradation, except at high pressures in deep water), and those between C5 and ~C30 are biodegradable
when present in crude oil. Therefore, n-alkanes of ≤ C30 are unlikely to persist in the environment and
pose little toxicological risk to aquatic life, unless they are ingested in oil droplets, because they are
essentially water-insoluble. iso-Alkanes tend to be more persistent in the environment because they are
more slowly degraded than their n-alkane counterparts; in fact, pristane (Table 2.1) and phytane, which
are common and often prominent iso-alkanes in petroleum (Appendix A, Figure A2), have been used as
short-term petroleum biomarkers because they usually resist biodegradation until the n-alkanes have been
depleted (Head et al. 2006). Therefore, petroleum biomarkers can be used as internal standards to monitor
the biodegradation rates of susceptible oil components, or for forensic evidence to identify sources of
contamination in the environment. For example, triaromatic steroids containing aromatic and aliphatic
moieties (Table 2.1) were used to distinguish naturally-occurring background hydrocarbons from diluted
bitumen (dilbit) contamination in the Kalamazoo River spill (US-EPA 2013a; Chapter 8). See Chapter 7
for a more complete discussion of petroleum biomarkers.

The term ‘biomarker’, as used in petroleum chemistry, designates a chemical compound originally
derived from living organisms that is poorly biodegradable and therefore persists over geological time
in the oil. This differs from the definition of ‘biomarker’ used in environmental toxicology (Chapter
4). Commonly used petroleum biomarkers include certain iso-alkanes and complex cyclic alkanes.

Naphthenes, or cyclo-alkanes, have the general formula CnH2n and can exist in multiple fused ring
structures with single or multiple alkyl substitutions. Low molecular weight (LMW) cyclo-alkanes (≤ C6)
are volatile and can cause toxic effects (e.g., US-EPA 1994; Sikkema et al. 1995) because they can
partition into biological membranes. As with iso-alkanes, cyclo-alkanes tend to be more resistant to
biodegradation than n-alkanes of the same molecular weight (Head et al. 2006). Diamondoids (based on
the adamantane structure; Table 2.1) are examples of multicyclic aliphatics, having cage-like structures
with or without alkyl side-groups (Wei et al. 2007). They are found in significant concentrations in
bitumen and are less susceptible to biodegradation than their polycyclic aromatic hydrocarbon (PAH)
counterparts (Government of Canada 2013).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 55
Table 2.1 Major chemical classes comprising petroleum, and structures of some example molecules
Name of chemical Example of class or Structure of example compound(s) Properties and relevance of
class or fraction fraction class or fraction
Usually the most abundant
SATURATES chemical class in petroleum;
highly water-insoluble; typically
are biodegradable and generally
non-toxic
n-Alkanes n- Hexadecane n-Alkanes <C6 are light liquids
or gases and may be toxic; C6-
C18 are readily biodegradable;
>C20 are waxes and more
difficult to biodegrade
iso-Alkanes Pristane iso-Alkanes are more resistant to
biodegradation than n-alkanes;
some, such as pristane and
phytane, are used as short-term
biomarkers for aerobic
biodegradation
cyclo-Alkanes and Cyclohexane; More resistant to biodegradation
Diamondoids Adamantane than n-alkanes; diamondoids
may be used as robust
biomarkers during
biodegradation (de Araujo et al.
2012)
Hopanoids and 17α(H),21β(H)-Hopane Many are non-biodegradable
Steroids (30αβ) and therefore are used as long-
term biomarkers for forensics
and biodegradation assessment

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 56
Name of chemical Example of class or Structure of example compound(s) Properties and relevance of
class or fraction fraction class or fraction
The examples shown below are strictly hydrocarbons; however, A hydrocarbon class that is
AROMATICS some aromatics having heteroatoms (i.e., S-, N- and O- generally more toxic than
heterocycles) fractionate with the aromatic hydrocarbons and saturates; many isomers and
are often considered together functionally. Examples of such homologous series are possible
aromatic heterocycles are shown in the resins fraction below. when aliphatic groups are
attached to the aromatic rings.
Monoaromatics BTEX: Benzene, The BTEX series is volatile,
Toluene, Ethylbenzene more water-soluble than other
and Xylene isomers hydrocarbons, acutely toxic
(ortho-xylene shown) and/or carcinogenic, and
relatively biodegradable under
aerobic and anaerobic
conditions
Polycyclic 2-Methylnaphthalene and Large number of isomers are
aromatic Benzo[a]pyrene possible due to multiple
hydrocarbons positions for alkyl side chains;
(PAHs) and some PAHs are toxic and/or
alkylated series carcinogenic; many persist in the
(alkyl PAH) environment because they resist
biodegradation; benzo[a]pyrene
is a US-EPA ‘Priority Pollutant’
Naphtheno- Acenaphthene and a Combination of saturated and
aromatics; triaromatic steroid unsaturated cyclic hydrocarbons;
aromatic steroids acenaphthene is a US-EPA
‘Priority Pollutant’; triaromatic
steroids have been used for
forensic attribution of
environmental contamination

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 57
Name of chemical Example of class or Structure of example compound(s) Properties and relevance of
class or fraction fraction class or fraction
RESINS The structures of individual resin compounds have not been A solubility class related to
determined, but they contain one or more S-, N- and/or O- asphaltenes (below), but of
heteroatoms; examples of aromatic S-, N- and O-moieties that lower molecular weight, lower
may form part of resin structures are shown below. The aromatic content and greater
individual chemicals below are commonly considered to be part polarity; poorly detected and
of the Aromatics fraction (despite being heterocycles rather resolved by most GC-based
than hydrocarbons). methods
Examples of S- Moieties may include: Organic sulphur moieties such as
containing Dimethylbenzothiophene these may contribute to total
moieties that may and 2-Methyldibenzo- sulphur content of sour crude
be constituents of thiophene oils
resin molecules

Examples of O- Moieties may Organic oxygen-containing


containing include:Benzofuran; groups in resins may be
moieties that may Benzo[b]naphthofuran generated from PAHs by
be constituents of photooxidation or partial
resin molecules biodegradation

Examples of N- Moieties may include: N-containing groups contribute


containing Benz[a]acridine;Dibenzo[ to polarity of the resins fraction
moieties that may a,i]carbazole
be constituents of
resin molecules

ASPHALTENES The structures of individual asphaltene molecules have not A solubility class of complex
been determined; they are related to resins (above) but have high molecular weight (HMW),
higher molecular weight, greater aromatic content and are less polar compounds that are
polar than resins. water-insoluble; contribute to
heavy oil viscosity; non-
biodegradable

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 58
Name of chemical Example of class or Structure of example compound(s) Properties and relevance of
class or fraction fraction class or fraction
Hypothetical asphaltene Numerous models of asphaltene
molecules, after Hoff and structures have been proposed of
Dettman (2012) and Boek different size, aromaticity and
et al. (2010) degree of condensation; different
subclasses of asphaltenes with
different chemical properties
have been proposed recently

MINOR ORGANIC COMPONENTS


Petroporphyrins C28 Etio vanadium Vanadium and nickel are
(organometals) porphyrin; common metals coordinated in
Nickel deoxophyllo- centre of the porphyrin structure;
erythroetioporphyrin (D the core derives from
PEP) (bacterio)chlorophyll molecules
transformed over geological time
by pressure and heat

Naphthenic acids Ethylcyclopentane-3- Heterogeneous class of O-


propanoic acid; containing polar compounds
Diamondoid acid associated with acute toxicity to
aquatic life; may be products of
partial biodegradation of
hydrocarbons; many resist
biodegradation

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 59
Fused complex alicyclic hydrocarbons include hopanoids and steroids that are derived from microbial,
animal and plant sterols and terpenes. Some of these compounds are extremely resistant to biodegradation
and therefore are ideal oil biomarkers that persist for thousands or even millions of years (Peters et al.
2005). Others, such as cholestane and possibly hopane, are less stable and may slowly biodegrade (Prince
and Walters 2007).

Waxes (linear alkanes >C20), which can represent a substantial proportion of certain crudes, may
precipitate out of petroleum at low temperature and can also form coatings on surfaces, affecting the
interfacial properties of the oil (Hollebone 2015). They are non-toxic but resist biodegradation due to low
bioavailability in water.

2.1.1.2 Aromatics

Aromatic hydrocarbons are cyclic, planar,


unsaturated compounds having structures based
Monoaromatics are the most water-soluble of
on a single benzene ring or multiple fused
the hydrocarbons and thus the most mobile in
benzene rings. They can have one or more alkyl
the water phase by diffusion. They may cause
groups in various isomeric arrangements. The
acute toxicity because of their ability to partition
most common (alkyl)-monoaromatics in
into biological membranes. PAHs having two or
petroleum are the BTEX series (benzene, toluene,
more fused aromatic rings are less volatile and
ethylbenzene and the three xylene isomers; Table
commonly more persistent than monoaromatics.
2.1). Benzene, ethylbenzene and toluene are on
the US-EPA list of toxic ‘Priority Pollutants’
(http://water.epa.gov/scitech/methods/cwa/pollut
ants.cfm). They are also light enough to be volatile, presenting flammability and breathing hazards.

PAHs have two or more fused aromatic rings and may additionally have alkyl side-groups (alkyl PAH).
Many of the smaller PAHs are of intermediate biodegradability that is inversely proportional to the
number of fused rings in the structure. Several ‘parent’ PAHs (i.e., unsubstituted PAHs such as
benzo[a]pyrene; Table 2.1) are on the US-EPA list of ‘priority pollutants’ because of their toxicity,
carcinogenicity (typically via inhalation) and/or resistance to biodegradation (van Hamme et al. 2003),
even though these parent PAHs typically represent <15% of the total PAH (TPAH) in the oil. The bulk of
PAHs have one or more alkyl substituents in various positions (e.g., 2-methylnaphthalene; Table 2.1),
generating many homologous series of alkyl PAH with different physical and chemical properties and
sometimes very different biological effects and biodegradability (Wang et al. 1998). The chronic toxicity
of oil is attributed primarily to alkyl PAHs having 3–5 rings (discussed in detail in Chapter 4).

Naphthenoaromatic structures are combinations of saturated and unsaturated rings, with or without alkyl
groups. Their concentrations, along with the naphthenes, tend to be greater in heavier oils (Speight 2014)
and contribute to the overall resistance of these oils to biodegradation. An example is acenaphthene
(Table 2.1), which is also a US-EPA ‘priority pollutant’.

Some heterocycles (e.g., those shown in Table 2.1 as moieties potentially found in resin structures),
including dibenzothiophenes, furans, carbazoles and their alkyl-substituted homologues, fractionate with
the aromatics even though they are not hydrocarbons. Some of these aromatic heterocycles are
biodegradable, whereas some isomers resist enzymatic attack (Kropp et al. 1996) or are only partially
oxidized, yielding dead-end products that are more water-soluble than the parent compound and have
different toxicity.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 60
2.1.1.3 Resins

Unlike saturates and aromatics that can be defined by their


Resins are not hydrocarbons chemical structures, petroleum resins (and asphaltenes, Section
since their chemical structures 2.1.1.4) are defined on the basis of their solubility in hydrocarbon
include elements other than solvents rather than on discrete structures (Appendix A, Figure
carbon and hydrogen, i.e., the A1), with the resins fraction being soluble in n-pentane or n-
heteroatoms S, N and/or O. heptane but insoluble in liquid propane (Speight 2004). The
structures of individual petroleum resins have not been
determined, although they all have one or more heteroatoms that
may be incorporated into ring structures that are saturated, partially saturated or aromatic, and/or long-
chain paraffin residues, with or without heteroatoms. Resins are chemically related to the asphaltenes
(Section 2.1.1.4), but are smaller (molecular weights estimated at 500 – 1,000 daltons [Da]; Bertoncini
2013), have lower aromaticity and are more polar2 than asphaltenes. In the petroleum industry, resins are
considered the smallest of the polar compounds in oil (versus hydrocarbons that are non-polar). Light
paraffinic crude oils may comprise >97% hydrocarbons and <3% resins, whereas heavy oils and bitumen
may be only ~50% hydrocarbons with the remainder being resins, asphaltenes, naphthenic acids and
metals (described below). In general, the resins constitute a greater proportion of petroleum mass than the
other non-hydrocarbon components, although this is not apparent by GC methods because the resins
either tend to be poorly resolved, contributing to the UCM ‘hump’ in chromatograms (Appendix A), or
not to be resolved at all by GC.

The polarity of resins is contributed by the presence of heteroatoms in diverse moieties within the resin
structure, such as those shown in Table 2.1, i.e., S-containing aromatic thiophenes, benzothiophenes,
dibenzothiophenes, naphthobenzothiophenes and aliphatic alkyl sulphides; N-containing pyrroles, indoles
and carbazoles; and/or O-containing hydroxyl, ester, carboxylic acid, carbonyl, furan and
sulfoxide/sulfone functional groups. These heteroatoms make resins more polar than hydrocarbons
(Speight 2004) and in some cases more toxic because of their water solubility (Melbye et al. 2009).
Because of the dominant effect of these moieties on the resins fraction, the resins are also sometimes
called the ‘Polar’ or ‘NSO’ fraction of petroleum.

Resins typically resist biodegradation, although some may be partially oxidized by microbes to generate
products of greater polarity and water solubility and thus potentially greater mobility, toxicity and
environmental persistence. In fact, some resins may represent the dead-end products of incomplete
biodegradation of organic matter over geological time, which are resistant to further enzymatic attack, or
that may have been generated from hydrocarbons in spilled oil by poorly-understood photooxidative
processes (Section 2.4.1.3).

2.1.1.4 Asphaltenes
Asphaltenes are the most complex, most
Asphaltenes are not hydrocarbons because they
diverse and highest molecular weight
contain S-, N- and O-heteroatoms, like the resins, but
components of petroleum. They are also the
they have greater average molecular weight (>400
least susceptible to biodegradation.
Da, mean weight ~1,200 Da [Gray 2015]). This
Asphaltenes comprise a small proportion of
chemical class (Speight 2004) represents myriad
light crude oils and greater proportions of
individual chemicals and is defined on the basis of
heavier oils and bitumens, conferring increased
insolubility in n-alkanes, such as pentane and
viscosity and density, among other properties.
heptane, but solubility in toluene (Appendix A,
Figure A1). Thus, petroleum can be separated into

2
Polarity refers to the distribution of electric charge across a molecule. Simplistically, polar compounds dissolve better in water
(a polar solvent) than do non-polar compounds.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 61
two major fractions simply by diluting it in 40 volumes of n-pentane (Speight 2004) or similar solvent,
whereupon the asphaltenes precipitate and the nC5-soluble fraction (known as the maltenes or de-
asphalted oil) remains dissolved. The proportion of asphaltenes in a crude oil affects its properties by
causing it to adhere to and affect the wetting characteristics of surfaces it attaches to, including pores and
channels in sediments, earning this class the nickname ‘the cholesterol of crude oil’ (Boek et al. 2010).

Despite considerable research (e.g., Strausz et al. 1992; Peng et al. 1997; Sheremata et al. 2004; Dettman
et al. 2005; Hoff and Dettman 2012), the structure of individual molecules within the asphaltene class has
remained controversial due to their HMW, complexity, heterogeneity and propensity to intimately
associate with other petroleum components like resins (Speight 2004). Asphaltene molecules are
proposed to exist as nanoaggregates (Boek et al. 2010) and/or as colloidal micelles in the bulk petroleum,
with resins forming the outer layer and preventing asphaltene aggregation (Petrova et al. 2011). Such
properties and interactions make chemical characterization technically difficult. Asphaltenes are neither
resolved nor detected by conventional GC but instead are subjected to alternative analyses, such as gel
permeation chromatography, nuclear magnetic resonance and high-temperature simulated distillation
(Appendix A). Speight (2004) considered that the predominant chemical groups in asphaltenes are
aromatic (sometimes >50% of the total carbon) with saturated hydrocarbon linkages. Hoff and Dettman
(2012) have isolated and defined different subclasses of asphaltenes having varying degrees of
condensation and differing alkyl substitutions (Table 2.1) and therefore different properties. Yang et al.
(2015) have further inferred and differentiated structural features of asphaltenes that stabilize oil:water
emulsions through interfacial activity (Yang et al. 2014). Thus, the structures and properties of
asphaltenes are gradually emerging with advances in analytical techniques but consensus has yet to be
achieved.

Recommendation: Continued fundamental research is needed into the composition of the


high molecular weight, polar components of oil, and standardized analytical methods for
characterizing these fractions should be developed and published. Characterization of the
resins and asphaltenes classes is presently incomplete because of current analytical
technical limitations preventing comprehensive, detailed monitoring and prediction of the
behaviour and fate of many spilled oils, particularly heavy oils and bitumen blends that are
rich in these fractions.

Because of their size and complexity, asphaltene molecules are extremely resistant to biodegradation
(despite a small number of questionable literature reports purporting to demonstrate asphaltene
degradation). Their stability in the environment is shown by their preservation for thousands of years in
ancient artifacts and their modern use in asphalt roads and roof shingles.

2.1.2 Minor Components of Crude Oils

Some of the minor constituents of crude oils


In addition to SARA fractions, petroleum may include organometallic compounds, such as
contain small and variable concentrations of petroporphyrins (Caumette et al. 2009) containing
non-hydrocarbon components derived from vanadium (V), nickel (Ni), iron (Fe) or copper (Cu)
biological or geological sources, such as derived from plant and microbial sources and
metals, organometals, sulphur, naphthenic dissolved in the bulk oil, as well as inorganic salts
acids, mineral particles and /or water. of these metals and others in colloidal suspension
(Speight 2014). Metalloporphyrin structures (Table
2.1) are often associated with the asphaltenes;
therefore, metals are more abundant in heavier
petroleum fractions and in heavy bituminous crudes at concentrations up to hundreds of parts per million
(Fingas 2015b).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 62
Whether metal atoms that are strongly associated with their organic ligands in petroleum actually become
water-soluble and bioavailable after oil spillage into water is a matter of debate (Joung and Shiller 2013).
Thus, reports of environmental impacts of heavy metals from oil contamination vary. For example,
concentrations of Ni and V in Saudi Arabian coastal sediments contaminated by massive volumes of
Kuwait oil during the 1991 Gulf War were only slightly elevated above background values, and may have
resulted from oil combustion as well as spillage (Fowler et al. 1993). In contrast, tonnes of Cu, Ni and V
were estimated to have been released to the Atlantic Ocean during fuel oil leakage from the sinking of the
Prestige oil tanker (Santos-Echeandía et al. 2008). Seawater concentrations of Cu and Ni (but not V) were
elevated in the upper water column (0-50 m depth), associated with upward movement of escaping fuel,
and in deep waters that had longer contact time with the oil (Santos-Echeandía et al. 2008). Similarly,
Wise Jr. et al. (2014) detected increased concentrations of chromium (Cr) and Ni in floating oil and tar
balls associated with the Deepwater Horizon (DWH) blowout in the Gulf of Mexico and elevated
concentrations in sperm whale tissue samples collected in the Gulf versus control samples. In contrast,
Joung and Schiller (2013) found that samples of surface waters near the DWH blowout site did not show
unusual concentrations of Ni or several other metals, but deep water samples collected near the well had
higher concentrations of cobalt (Co). Elevated barium (Ba) concentrations detected in deep water were
likely associated with the drilling mud used to contain the blowout rather than the oil plume itself. Thus,
the magnitude and effects of metal contamination from individual spills require additional study, whereas
the accumulation of metals in chronically hydrocarbon-contaminated water and sediments (e.g., those
impacted by refineries or shore-to-ship oil transfer) is well-known (e.g., González-Macías et al. 2006).

Inorganic S present in elemental form (S0) or as gaseous H2S may be a minor component by mass (often
<1%), but H2S is toxic to humans even at parts per million levels. H2S is also a major concern regarding
corrosion to infrastructure (e.g., pipelines and surface handling facilities) and explosion potential in
confined spaces, including rail cars. Depending on the water content of the oil and its pH, H 2S may exist
as a gas or may be dissolved in the water phase. H2S may represent only a fraction of the total S in sour
crudes, which can contain significant quantities of organic S compounds (Table 2.1) that contribute to
total S. The organic S compounds have different chemical properties than H2S, and are more problematic
in refineries than in the environment. Oils with a significant proportion of total S (usually ≥1%) are called
‘sour’ crudes, as opposed to ‘sweet’ crudes with ≤0.5% S (Section 2.3.6). Sweet oils are generally more
valuable than sour, as the presence of S imposes a price penalty at the refinery, and H2S concentrations
must be reduced before transportation for safety reasons.

Naphthenic acids (NAs; Table 2.1) comprise a large, diverse group of alkyl-substituted cyclic and
noncyclic carboxylic acids, historically defined as having the chemical formula CnH2n+ZO2, where z
reflects the number of rings in the structure (Clemente and Fedorak 2005; Johnson et al. 2011). Thus, they
share similarities with the resins fraction, being composed of both hydrocarbon and heteroatom moieties
and being polar. NAs have also been defined by the extraction methods used to recover them from
samples and by their infrared (IR) absorbance spectra, but the range of possible structures has been
controversial, currently precluding a definition accepted by both industry and academia (Grewer et al.
2010; Rowland et al. 2011). Whereas NAs were previously considered to be saturated compounds with a
single acidic group, this class has expanded to include aromatic NAs (Jones et al. 2012) and NAs with
multiple oxidized functional groups (i.e., carboxylic acids plus hydroxyl groups). Definition of NAs has
become even more nebulous (Grewer et al. 2010) with the development of sophisticated analytical
methods (such as ESI FT-ICR-MS and multi-dimensional GC-MS; Appendix A) that detect more isomers
and analogues than previously recognized. NAs have been used industrially as components of wood
preservatives and mixtures are commercially available. However, application of new analytical methods
has shown that the component NAs in those mixtures differ significantly from naturally occurring suites
of NAs in petroleum-impacted waters (Grewer et al. 2010), throwing into question some of the early
research on NA biodegradation conducted using commercial mixtures.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 63
NAs are ubiquitous, but have been associated particularly with biodegraded oil reservoir fluids and oil
sands process waters in western Canada where they may represent the accumulation of dead-end products
of incomplete microbially-mediated oxidation of cyclo-alkanes (naphthenes; Table 2.1) (Rowland et al.
2011) over geological time. This view is supported in part by NA persistence in the environment, as many
resist further biodegradation (Whitby 2010). Current environmental interest in NAs arises from their
reported acute toxicity to aquatic life (Allen 2008; Zhang et al. 2011), contribution to Total Acid Number
(TAN; the petroleum acidity parameter), and accretion of precipitates (scale) that increase corrosion of
pipelines, refinery units and other oil-handling infrastructure. Due to their acidic group(s), they are more
polar than hydrocarbons, and this water solubility makes them more mobile in the environment than the
oil itself. In fact, petroleum that has been ‘water-washed’ in the reservoir or the environment may become
depleted in NAs as they dissolve into the water fraction. Conversely, spilled oil having high NA contents
may be a source of toxicity in aquatic environments as the NAs dissolve in water. However, attribution of
toxicity to specific NA compounds within a complex mixture of NAs is currently in flux due to the
expanding definitions of NA structures, discussed above. Further discussion of NA toxicity in oil sands-
associated environments is available (Gosselin et al. 2010).

Crude oils may also contain small amounts of minerals, such as clays, and emulsified water originating
either from the geological formation (connate water) or from water injected into the reservoir during oil
recovery (produced water). The great majority of entrained or emulsified water is removed from the oil to
meet pipeline specifications (typically ≤0.5% water plus sediments).

It is clear that the chemical composition of crude oils is extremely complex and variable, and the
structures of many oil components, including the UCM, are currently cryptic. The limitations of resolving
and identifying individual molecules in the resins and asphaltenes fractions, particularly, are slowly being
addressed with renewed scientific interest and development of new techniques. However, improved
analytical methods for chemical characterization are still needed to gain a better understanding of the
persistence and behaviour of these components in spilled oil. Additionally, the quality of data in different
oil property archives varies according to methodology. Development and adoption of standardized
methodology with data quality indices would allow for comparisons among studies.

Recommendation: Additional analytical research is needed over the long-term to address gaps in
fundamental knowledge of asphaltenes, resins and naphthenic acids, and to populate oil-property
databases with verified data. Such knowledge provides insight into criteria that affect
emulsification, photooxidation, biodegradation, toxicity, adhesion, etc., particularly for heavy oils
and bitumens that have high proportions of such chemical classes.

2.2 Bulk Properties of Oil

Complex mixtures, such as petroleum, exhibit complex chemical and physical properties requiring suites
of analytical methods to discern important differences between oils. Several properties are discussed
below; they are interrelated because all are affected by oil composition. Appendix B (Table B1)
comprises a summary of major oil properties, associated analytical methods and the relevance of those
properties to oil spill cleanup (discussed fully in Chapter 7).

2.2.1 Density, Specific Gravity and API Gravity

Density is the mass of a given volume of oil, variously expressed in units of g/mL or g/cm3 or kg/m3. Two
measures of density are commonly applied to petroleum: specific gravity and API (American Petroleum
Institute) gravity. Specific gravity is the density of a substance compared to that of water (and as a ratio is
dimensionless). Because most oils are lighter than water, they float on it. For example, fresh water has a
density of 1.0 g/mL and sea water a density of 1.03 g/mL at 4 °C. Therefore, an oil of specific gravity
<1.0 (e.g., gasoline at 0.8) will float on water at that temperature and one with specific gravity >1.0 will

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 64
sink at that temperature. As the temperature increases, the density (and thus the specific gravity) of oil
decreases (Nmegbu 2014). Thus, bitumen and certain heavy fuel oils may have densities greater than
water at some temperatures and may submerge (Section 2.4.2.3).

API gravity (measured in degrees; ° API), devised by the American Petroleum Institute, is an inverse
measure of density that is used to compare different petroleum types (Table 2.2; Figure 2.1; Appendix C,
Table C1). Fresh water is assigned a gravity of 10° API at 15 °C; ‘light’ crude oils have API gravity of
>31.1°; ‘medium’ crudes are 22.3-31.1° API; ‘heavy’ crudes <22.3° API (or, more stringently, 10-15°);
and bitumens 5-10° API (Speight 2014). API gravity can also be considered a rough indicator of crude oil
quality, since more valuable light oils have higher ° API values.

Oils of API >10° will float on fresh water at 15 °C whereas bitumen, defined as <10° API (Speight 2014),
will sink. However, the sinking or floating of oil on water is additionally affected by salinity (e.g.,
seawater density 1.03 g/ml), by temperature and by interactions with particles (Section 2.4.2.3).
Weathering of spilled oil, particularly diluted bitumen, can have a profound effect on the oil’s density
because evaporation (Section 2.4.1.2) of the LMW components (‘light ends’) leaves residual oil of higher
density that may no longer be buoyant. Thus, the decrease in specific gravity (increase in ° API) of spilled
oil at higher temperatures may gradually be overcome by faster evaporation rates, resulting in an oil that
is more dense.

Figure 2.1 Relationship among viscosity, density and API gravity over a range of petroleum types. Adapted from
Speight (2014).

2.2.2 Viscosity and Pour Point

Viscosity, defined as resistance of a liquid to


deformation by shear or flow (WSP 2014), is also Oil viscosity, like density, is determined in
informally described as resistance to flow, or the large part by its chemical composition. Higher
‘thickness’ of a fluid. The lower the viscosity, the proportions of ‘light’ components, such as
more easily it flows. Two types of viscosity may be LMW alkanes and aromatics, contribute to
reported: 1) dynamic viscosity (resistance to shear, lower viscosity; whereas, heavier components,
where adjacent layers move parallel to each other but such as asphaltenes and resins, increase
at different speeds), expressed as centipoise (cP) or petroleum viscosity.
milliPascal second (mPa s); and 2) kinematic
viscosity (the ratio of the dynamic viscosity to the
density of the fluid) reported in centiStokes (cSt) or m2/s. The significance of viscosity in oil spill
response is discussed in Chapter 6.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 65
Gasoline has a dynamic viscosity of 0.5 mPa s, light crude viscosity may be 5-50 mPa s, whereas
‘conventional’ crude oils have viscosity up to 10,000 mPa s (similar to that of molasses) and ‘ultraheavy’
oils and bitumen usually have viscosity >50,000–100,000 mPa s (Speight 2014), similar to peanut butter.
The relationships between viscosity, density and API gravity are shown in Figure 2.1. Increased viscosity
decreases oil spreading on water (Section 2.4.1.1), dispersion (formation of oil droplets in water; Section
2.4.2.2) and the efficacy of chemical dispersants during remediation (Chapter 6).

As shown in Figure 2.2, kinematic viscosity is not only a function of the average molecular weight of the
oil, it also declines exponentially as temperature increases (Nmegbu 2014). Thus, temperature impacts the
extent of oil spreading on surfaces and penetration of shoreline sediments. Temperature also defines the
oil’s pour point - the temperature below which an oil will not flow but rather starts to solidify or gel.
Bitumen, for example, does not flow in the reservoir and must be heated or, alternatively, bitumen must
be diluted with light oils for transport (Section 2.3.7). The interrelatedness of viscosity, pour point and
temperature then becomes particularly important when considering the release of heavy oil or diluted
bitumen from a heated pipeline or flow of an oil spilled onto ice or cold water.

Dilbit

Light crude
VISCOSITY (cSt)

Medium Heavy crude

Heavy crude

TEMPERATURE (°C)

Figure 2.2 Relationship between temperature and kinematic viscosity (reported in centiStokes [cSt]) for four oil
types (Table 2.2). Adapted from Tsaprailis (2014)

2.2.3 Volatility and Flash Point

High concentrations of light, volatile


The ‘light’ components of petroleum that hydrocarbons contribute to the flammability
contribute to decreased viscosity and density potential of an oil by lowering the flash point, the
(e.g., LMW saturates and aromatics) also tend to lowest temperature the fuel must experience to
evaporate (volatilize) readily. Volatility is achieve ignition in the air. Notably, these light
particularly important when considering diluted hydrocarbons, particularly the monoaromatics,
bitumen blends, which contain substantial also correlate with acute toxicity to life forms (as
proportions (≥30%) of light petroleum products discussed in Chapter 4). Thus, vapours from an oil
that pose a flammability hazard, even though the spill can pose a hazard to spill responders, as well
majority of the oil mass (the bitumen) is as to wildlife exposed to the fumes.
considered non-flammable.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 66
After volatilization of some or most of the light components through weathering processes in open
conditions (Section 2.4), the oils will have decreased flammability and acute aqueous toxicity.
Temperature is obviously a primary factor affecting volatility and flash point, with higher temperature
speeding volatility losses from spilled oil, but the thickness of the oil layer (which will also respond to
temperature via viscosity changes) will also affect the rate of evaporation.

2.2.4 Water Solubility

In general, the saturates are highly water-


insoluble compared with aromatics of a similar Most petroleum hydrocarbons are virtually
carbon number (e.g., the aqueous solubility of insoluble in water. Only the LMW aromatics (e.g.,
naphthalene, a C10 PAH, is approximately 600- BTEX), very small saturates and small polar
fold greater than the C10 saturate n-decane). compounds (some resins) have appreciable
Monoaromatics are the most water-soluble of the solubility in water, typically in the range of parts-
hydrocarbons, whereas (alkyl-) PAHs are poorly per-million (ppm). Water solubility makes these
soluble (e.g., at parts-per-billion concentrations) molecules potentially more toxic because they are
yet retain their toxic properties. For example, the more ‘bioavailable’ to aquatic life if they
aqueous solubility of benzene (Table 2.1), the subsequently partition into biological membranes.
most soluble BTEX compound, is 1.8 g/L water
at 15 °C; solubility of n-hexane is 0.0013 g/L at
20 °C; naphthalene is 0.031 g/L at 25 °C; and n-hexadecane (C16) and cyclohexane (C6) (Table 2.1) are
considered water-insoluble. Small aromatic heterocycles and resins, being polar, may be more soluble
than their hydrocarbon counterparts. For example, whereas the hydrocarbon benzofuran (Table 2.1) is
considered water-insoluble, its S-containing analog benzothiophene is soluble at 0.13 g/L at 25 °C and
indole, its N-containing analog, is soluble at 3.6 g/L at 25 °C. However, large resin compounds and
asphaltenes, despite their polarity, are water insoluble and the resins in particular tend to partition to the
interface between oil and water.

High molecular weight components of bitumen, such as asphaltenes and most resins, are thought not to be
toxic because they have extremely low solubility in water and do not readily cross biological membranes.
However, diluents added to bitumens contain LMW compounds that can dissolve in water and contribute
to acute aqueous toxicity before they are lost by weathering. In addition, heterocycles (e.g.,
dibenzothiophenes; Table 2.1) and 3-to-5-ringed alkyl PAHs detected in dilbit are derived from bitumen.
Their concentrations in dilbit are similar to those of conventional crude oils and are thought to be the
components causing chronic toxicity to fish embryos (Madison et al. 2015); see Chapter 4.

A measure of water solubility that is particularly pertinent to biological effects of petroleum is the
octanol-water partition coefficient, which describes the concentration of a chemical in the solvent octanol
(C8H15OH, a mimic for biological lipid membranes) versus its concentration dissolved in water, in a two-
phase system at equilibrium for a given temperature. This ratio is usually expressed as the logarithm of
the ratio (i.e., as log Kow). The greater the Kow value, the less polar and less water-soluble the chemical.
Thus, Kow helps predict the propensity of a petroleum constituent to dissolve into water and subsequently
partition into a biological lipid membrane, an important factor in toxicity (Sikkema et al. 1995). K ow and
toxicity are discussed in detail in Chapter 4.

2.2.5 Surface Tension and Interfacial Tension

At liquid-air interfaces, surface tension is the measure of attraction (cohesion) between the surface
molecules of a liquid rather than to molecules in air (adhesion). Information on the surface tension of
different crude oils at various temperatures is available in Environment Canada’s Oil Properties Database
(http://www.etc-cte.ec.gc.ca/databases/oilproperties/). In the case of two immiscible liquids such as oil
and water, the force between their surfaces is referred to as interfacial tension. The higher the interfacial

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 67
tension, the less the oil will spread on water; whereas, if the interfacial tension is low, the oil will spread
evenly without help from wind and water currents. Because higher temperature reduces interfacial
tension, oil is more likely to spread in warmer waters than in cold waters. However, interfacial tension has
less effect on oil spreading than viscosity.

2.2.6 Adhesion

Oil adhesion is the property of petroleum sticking to a surface, whether it is the rock formation
comprising the oil reservoir, the production equipment or environmental surfaces after a spill. It is not
included in the standard suite of oil properties reported by petroleum producers and is rarely discussed in
the oil spill literature, but might be a predictor of oil spill cleanup needs if further measured and calibrated
against a variety of oil types and their behaviour in the environment. A semi-quantitative gravimetric
method has been proposed (Jokuty et al. 1995) in which a stainless steel needle is dipped in oil that is then
allowed to drain, and the mass of oil adhering to the needle after 30 minutes of draining is weighed to
produce a measurement with units of mass per surface area (g/m2). However, the method has been neither
adopted as a standard nor vetted. The relationship between evaporative weathering and adhesion is
discussed in Section 2.4.1.2.

2.3 Examples of Oil Types

This section introduces a selection of oil types that are relevant to petroleum transport in Canada (Table
2.2) and/or to this Report. Within each oil type, one or more specific crude oils or blends is presented to
highlight the differences between oils in terms of their chemical and physical properties, known or
predicted behaviour and fate in the environment, timelines of weathering processes, and expected residues
and spill response strategies (discussed in detail in later chapters).

With few exceptions this Report does not consider refined products, such as jet fuel, gasoline or diesel,
nor liquefied natural gas or biofuels. The exceptions include products used to dilute bitumen for
transportation and selected refined heavy fuel oils, for comparative purposes. The information presented
in Table 2.2, and Appendix C (Table C1) has been collated from many different sources and is
representative but not comprehensive. Two major online sources of information for oils transported in
Canada are the Environment Canada Oil Properties Database (http://www.etc-
cte.ec.gc.ca/databases/oilproperties/) and the Crude Quality Inc. industry website (crudemonitor.ca).
However, both are incomplete in scope and compositional information. The Environment Canada website
hosts a valuable catalogue of various chemical and physical data for 450 oils produced and transported
globally but, regrettably, it lacks entries for many of the emerging unconventional oil types transported in
Canada (e.g., dilbit, synbit, shale oil). The crudemonitor.ca website provides monthly analyses of oils
currently transported inter-provincially in Canada, including unconventional oils, but publishes chemical
composition data only for the light ends (≤C8) and general parameters, such as TAN, metals and sulphur
content (and High Temperature Simulated Distillation curves [HTSD; Appendix A]) for heavy oils and
blends. These data are relevant for achieving pipeline transportation specifications but not necessarily for
oil spill response. Full spectrum analyses of heavier components that are predominant in heavy oils and
bitumens and highly relevant to environmental cleanup are not available on this website.

Recommendation: Online public databases with compositions and properties of oils transported in
Canada should be expanded to provide support for oil spill preparedness, response decisions and
for the safety of first responders to a spill. To enable collection of relevant chemical data, expanded
and standardized analytical methods must be developed, particularly for characterizing bitumen
blends, such as dilbit and synbit.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 68
Table 2.2 Properties of some oil types commonly transported in Canada and/or relevant to this Report.a
Oil type and examples Region of origin Relevance
Natural condensates and refined products used as diluents for transport
Sable Island condensate Nova Scotia 40° API; Naturally occurring sweet ultra-light crude
oil, often produced in conjunction with natural gas;
predominantly volatile hydrocarbons
Fort Saskatchewan Alberta 68° API; Refined products, predominantly LMW
condensate blend aliphatics (C5-C6), with minor BTEX; used as diluent
for oil sands processing and for bitumen transport
Cold Lake Diluent Alberta 69° API; Natural gas condensate used as a bitumen
diluent
Naphtha Alberta 57° API; Refined product, predominantly ≤C12
saturates and monoaromatics used as diluent for oil
sands processing and for bitumen transport
Southern Lights US 80° API; US condensate shipped to Alberta;
Condensate Diluent predominantly C5-C12 saturates
Suncor Synthetic Crude Alberta 33° API; Refined product derived from partially
Oil (SCO) upgraded bitumen and used as diluent in synbit
Light crude oils
Alberta Sweet Mixed Alberta 36° API; Blended aggregate of light, sweet
Blend (ASMB) conventional crude oils shipped from Alberta; used
Reference #4 as a benchmark and laboratory reference oil by
Environment Canada
Macondo (MC252) Gulf of Mexico, 35° API; Oil released during DWH blowout; rich in
USA LMW saturates that are gases or light liquids at
atmospheric pressure
Norman Wells Northwest 38° API; Light crude from Canadian Arctic; wells
Territories drilled from artificial islands in Mackenzie River
Statfjord Norway 38° API; North Sea offshore light oil
Arabian Light Saudi Arabia ~32° API; Sample from St. John, NL, refinery
West Texas Texas, USA 36-41° API; Benchmark oil for North American oil
Intermediate (WTI) market pricing
Medium crude oils
Alaska North Slope Alaska, USA 29° API; Conventional oil from Arctic; spilled from
(ANS) Exxon Valdez oil tanker into Prince William Sound,
AK, in March 1989
Atkinson Point Beaufort Sea, 24° API; Oil from an onshore discovery well on
Canada Tuktoyaktuk Peninsula, NWT; no current production

Prudhoe Bay-A Alaska, USA 29° API; Alaska North Slope crude; Used as
(PBCO) reference oil by US-EPA in 1995

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 69
Oil type and examples Region of origin Relevance
Gullfaks and Troll Norway 29° API; typical North Sea oils; used for oil spill
research in Arctic regions e.g., Svalbard
Shale oils (Tight oils)
Bakken North Dakota and 42° API; unconventional light crude produced by
Montana (USA), fracking; may contain H2S (23,000 ppm in an
Saskatchewan exceptional case); flammable agent in the 2013 Lac-
and Manitoba Mégantic QC, derailment and explosion
Waxy crude oils
Hibernia and Terra Newfoundland 34° API; Offshore Atlantic oils transported to
Nova mainland by tanker ship
Sour crude oils
Midale Saskatchewan, 31° API; 2.4% S; Benchmark for Canadian medium
Manitoba sour crude oils
BC Light British Columbia 41° API; sour light crude spilled into Pine River, BC,
in 2000
Western Canadian Alberta 22° API; 3.1%S; A blend of heavy sour crude oils,
Blend (WCB) deliberately undiluted with butane that is co-
produced with the crudes (to reduce explosion
hazard) but instead diluted with condensate or light
crude for transport
Heavy oils and diluted bitumen (See Table 2.3 for definitions of various bitumen blends)
Bunker C fuel oil (No. International 11° API; Common refined fuel oil class, often diluted
6 Fuel oil) including with lighter petroleum for transportation; heavy
HFO 7102 component of oil spill in Lake Wabamun AB, 2005
Cold Lake Blend (CLB) Alberta 23° API; Oil sands bitumen diluted with lighter
dilbit petroleum; highest volume dilbit type transported in
Canada; tested in wave tank trials for sinking and
emulsification; one of the dilbits in the Kalamazoo,
MI, oil spill in 2010
Albian heavy synthetic Alberta 20° API; Partially upgraded heavy crude oil,
crude (AHS) transported as dilsynbit; used as refinery feedstock
Access Western Blend Alberta 23° API; sour heavy oil (3.9% S) produced by
(AWB) SAGD; tested in wave tank trials (Government of
Canada 2013) for sinking and emulsification;
composition of pipelined oil changes with season
Wabasca Heavy Alberta 16° API; Heavy oil spilled near Mayflower, AR, in
2013
Western Canadian Alberta 19-22° API; Considered the benchmark for heavy,
Select (WCS) acidic (high TAN) crudes blended with bitumen
streams to maintain TAN <1; diluted 10-50% with
sweet synthetic crude oil and/or condensates; a

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 70
Oil type and examples Region of origin Relevance
component of the Kalamazoo, MI, dilbit spill in 2010
Synbit Blend Alberta 21° API; Blend of several synbits plus heavy oils;
evaporation behaviour differs from dilbit because of
lower concentration of light ends in diluent
Athabasca/Cold Lake Alberta 8-10° API; Density > 1.0; must be diluted or heated
undiluted bitumen for transport

Orimulsion™ Venezuela 8° API; A proprietary emulsion of Orinoco bitumen


with 26% fresh water and surfactant, density > 1.0;
must be maintained above 30 °C to prevent gelling;
limited production at present
a
Additional data including composition, specific gravity, sulphur content, adhesion and dispersability, where
available, are given in Appendix C, Table C1. Note that the composition of several oils, particularly blended oils
and dilbits, changes seasonally to maintain pipeline specifications for viscosity.

2.3.1 Condensates and Refined Products Used as Diluents for Bitumen

Condensates are rich in LMW saturates and, if


Condensates are conventional naturally-occurring spilled, would experience rapid evaporation,
ultra-light crude oils, often co-produced with losing a high proportion of the spilled mass of
natural gas (methane). For example, Sable Island oil. Because of the high saturate content, after
Condensate is produced from natural gas fields evaporation and dissolution a weathered
(e.g., Thebaud and Venture) off Nova Scotia and condensate would typically be highly
shipped to the mainland (Point Tupper, NS) by biodegradable, given appropriate conditions,
pipeline. and little residual oil would be expected to
persist in an open water spill.

Historically, natural condensate from western


Canadian gas fields was used to dilute bitumen and heavy oils from the Athabasca and Cold Lake regions,
but more recently the diluents are blends of light oils and/or refinery fractions, including naphtha and
synthetic crude oil that is itself derived from bitumen (Segato [no date]; Crude Quality Inc. 2015).
Although the diluents are predominantly composed of LMW materials far lighter than the bitumen, it
should be noted that most diluents also contain appreciable proportions of the same HMW compounds
found in bitumen that may not be detected by conventional GC (GC-FID or GC-MS), but are revealed
using HTSD (Appendix A).

2.3.2 Light Crude Oils

Light crudes with API gravity >31.1° (Speight 2014) are conventional oils found throughout the world.
Examples include Arabian Light, WTI (used as a pricing benchmark in the oil market), Statfjord from the
North Sea and Norman Wells from northern Canada. Alberta Sweet Mixed Blend (ASMB), a major crude
shipped from Alberta, has been used as a reference oil by Environment Canada for chemical composition,
biodegradability and emulsification studies. Perhaps the most infamous light oil is Macondo (MC252)
from the DWH oil well in the Gulf of Mexico that suffered a catastrophic blowout in 2010 (Section
2.4.2.1, Section 2.4.2.3 and Chapter 8), injecting not only liquid oil into deep seawater, but also light
saturates that are gases at atmospheric pressure (methane, ethane, propane and butane).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 71
2.3.3 Medium Crude Oils

Oils between 22-31° API are considered ‘medium’ crudes (Speight 2014). These conventional crudes
include many of the ANS oils, Prudhoe Bay crude oil from the Arctic, and Norwegian North Sea oils
produced from offshore wells. The latter oils have been used for research projects in the Arctic to
determine the behaviour and fate of oil spills onto Arctic beaches. ANS oil was spilled in the highly
publicized 1989 grounding of the Exxon Valdez in Prince William Sound, AK (Chapter 8), where up to
25% of spilled oil mass was lost by weathering and another ~30% was estimated by some researchers to
have biodegraded within a few weeks of the spill (Atlas and Hazen 2011).

2.3.4 Shale Oils (Tight Oil or Light Tight Oil)

Shale oils 3 are unconventional liquid crudes that have only recently entered the North American oil
markets in substantial volumes. As opposed to being newly ‘discovered’, many shale oil reservoirs were
already known but were inaccessible pending recent technological advances afforded by horizontal
drilling and hydrofracturing (‘fracking’) to release liquid oil that otherwise cannot flow because of ‘tight’
pores in the formation. One large known reserve is the Bakken formation straddling North Dakota,
Montana and parts of southern Saskatchewan and Manitoba. Although Bakken oil is generally considered
‘sweet’ (i.e., low organic S and low H2S), occasional samples have had high H2S concentrations
(American Fuel & Petrochemical Manufacturers 2014). Bakken shale oil was the flammable product in
the Lac Mégantic, QC, rail disaster in 2013 and H2S was implicated as a possible contributor to the
explosion.

2.3.5 Waxy Crude Oils

These crudes contain the typical suite of petroleum fractions but also significant proportions of waxes
(HMW linear paraffins, typically >C20) and large naphthenes (cyclo-alkanes; Table 2.1). When present at
high concentrations in petroleum, waxes may crystallize (‘freeze’) and form micro- or macroscopic
deposits in the reservoir or infrastructure, depending on temperature and pressure. Because HMW
saturates are less biodegradable than those of ≤C18, spills of waxy oils may persist in the environment
longer than non-waxy oils, although the wax residues have low chemical toxicity to aquatic life. Oils
produced from the Hibernia and Terra Nova fields off the coast of Newfoundland are considered waxy
crude oils.

2.3.6 Sour Crude Oils

A crude is considered ‘sour’ if it has more than 0.5 wt% total S content (>1% in some markets), as
opposed to ‘sweet crude’ that has <0.5 wt% S. Most S in petroleum is included within aromatic or
saturated structures in resins or asphaltenes (e.g., in heterocycles; Table 2.1) or is S0, in which case it is
not considered hazardous but is removed in the refinery at considerable expense. However, where oils
contain highly toxic and potentially explosive H2S gas, the H2S concentration is reduced to safe
concentrations before transport. The amount of H2S varies considerably with the crude oil source (Hess
2012). Thus, both conventional and unconventional oils can be ‘sour’, regardless of the concentration of
H2S, with the unconventional oils more likely to contain organic S compounds that may resist
biodegradation. Examples include the conventional medium sour crude Midale, and the WCB and AWB
dibits; the latter was a component of the dilbit spill in Kalamazoo, MI (Chapter 8).

3
Shale oil (‘tight oil’) should not be confused with oil shale. The latter is a sedimentary rock that harbours kerogen, a solid
immature organic material that can be thermally processed at surface facilities into a shale oil that differs from the natural shale
oil.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 72
2.3.7 Heavy Oils, Bitumen and Diluted Bitumen Products

Heavy oils are defined as being <22° API or, more stringently, 10-15° API, with bitumens being semi-
solids having density >1.0 g/cm3 or 5-10° API (Speight 2014). Some heavy oils are refinery products and
represent the ‘heavy ends’ that remain after
distillation of lighter fractions like gasoline and
Heavy oils and bitumen are at the end of a kerosene from lighter oils. Bunker C fuel oil
continuum of increasingly heavy, viscous oils (also called No. 6 fuel oil or residual fuel oil;
with increasing proportions of HMW resins and Table 2.2) is one such refined product that can
asphaltenes and decreasing proportions of light be blended with lighter fractions to make an oil
molecules. for combustion or used undiluted to make
asphalt.

The distinction between heavy oil and bitumen is blurred by their similar chemical and physical
properties and obscured by the working definitions and colloquial terms used by industry. Thus, different
entities may assign the same oils to different categories. As noted by Winter and Haddad (2014), Exxon
and CrudeMonitor.ca both identified the Wabasca Heavy Crude that spilled in Mayflower, AR, in 2013
(Chapter 1) as a ‘diluted heavy crude’, but the Canadian Government, Battelle (acting for API) and
Penspen (acting for Canadian Energy Pipelines Association, CEPA) initially labeled it a ‘diluted bitumen’
(later corrected). The WCS blend that spilled in Kalamazoo, MI, in 2013 (Chapter 8) was called ‘heavy
oil’ by Enbridge, but ‘dilbit’ by environmental agencies, such as Environment Canada (Government of
Canada 2013). According to Yang et al. (2011), bitumen products have unique chemical markers that
distinguish them from conventional crude oils. These fingerprints include the distribution of PAH, which
is generally skewed towards the smaller multi-ringed structures (e.g., naphthalenes; Table 2.1) in
conventional crudes versus larger PAHs (e.g., chrysenes) in bitumens, and have different distribution
profiles within the alkyl PAH isomer series. However, in practice the two types of oil may be conflated in
definition.

Regardless of nomenclature, it is generally accepted that both heavy oil and bitumen represent former
conventional crude oils that have been extensively biodegraded and thermally altered in situ over
geological time during uplifting or migration of the petroleum (Strausz et al. 2010; Fustic et al. 2012).
Biodegradation depleted the lighter saturate fraction of the crude oil first, leaving poorly-degradable,
complex, heavy molecules as residues. In addition, the products of incomplete microbial oxidation over
geological time likely contributed to an increased resins fraction, yielding an extremely viscous, heavy
petroleum (bitumen and extra-heavy oil) that cannot flow to production wells under normal reservoir
conditions. This concept of liquid oil biodegradation to semi-solid bitumen over geological time is
supported by the chemical compositions of heavy oils and bitumen, which have small proportions of light
compounds known to be biodegradable (such as n-alkanes, iso-alkanes, BTEX and LMW PAH) and
correspondingly higher proportions of poorly or non-biodegradable components (resins, asphaltenes and,
of course, biomarkers like hopanoids, steroids and diamondoids; Table 2.1). Canada and Venezuela have
the world's largest deposits of these highly viscous extra-heavy oils. However, Venezuela’s Orinoco Belt,
which is sometimes described as oil sands, has lower viscosity than the Canadian oil sands bitumen and
can be transported as a 70:30 oil:freshwater emulsion (Orimulsion®; McGowan 1990) rather than being
heated or diluted with light hydrocarbons. Whether natural or refined, extra-heavy oils contain much less
biodegradable material than lighter crude oils.

There is a variety of diluted bitumen blends (Table 2.3). The most common diluents for bitumen transport
include condensate (ultra-light oils partitioned from natural gas wells) or naphtha (a refined product),
often at a ratio of 30% diluent to 70% bitumen for transport by pipeline (Crosby et al. 2013; Dew et al.
2015), whereas railbit for transport by rail tanker car is diluted half as much (i.e., ~15% diluent and 85%
bitumen; Fingas 2014-2015). Synbit is a blend of bitumen and synthetic crude oil, a product of partial
upgrading of bitumen that is heavier than condensate, at a ratio of approximately 50:50 (Crosby et al.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 73
2013). Synbit has different properties than dilbit because of the heavier nature of the diluent. Dilsynbit is
bitumen diluted with synthetic crude oil plus diluent.

In addition to seasonal changes in dilbit blends,


Whereas some heavy oils may be suitable for the composition and proportions of diluents may
shipping without alteration, bitumen does not flow differ in blends transported by inter- versus
unless it is heated and/or diluted. Therefore, intra-provincial pipelines. This is because
bitumen is diluted with various light petroleum producers must conform to regulations when
products to make it less viscous for different types shipping oil outside the province, but may use
of transport (rail, ship, pipeline). specific blends to move bitumen between their
own facilities. Notably, although the
CrudeMonitor.ca database provides information
on oils that are transported inter-provincially, it does not cover intra-provincial pipelines or those internal
to an operator’s facilities, which may have different specifications. For example, a paraffinic nC5-nC6
solvent used for bitumen extraction from oil sands ores by Shell Albian Sands in northern Alberta is also
used as a diluent to pipeline the bitumen to the upgrading facility near Edmonton, but the solvent is too
expensive to use as diluent for shipping to the U.S.; instead it is recovered and returned by pipeline to the
extraction facility (NRC 2013). Another example is the bitumen:water emulsion that spilled from an intra-
facility pipeline at Nexen’s Long Lake operation in July 2015 (Chapter 8). It is also important, as noted
previously, to be aware of the limited petroleum characterization published by Crude Quality Inc.
(crudemonitor.ca), specifically alkanes ≤C6 and BTEX contents. These light hydrocarbons are important
for safety and viscosity reasons during transport, but represent only a small proportion of heavy oils and
bitumen blends in which the predominant heavy components are more relevant to oil behaviour and post-
spill cleanup. Thus, there is a need for more internet-accessible oil composition data for emerging and
unconventional oils to assist with planning, preparedness and response to spills.
Table 2.3 Definitions of diluted bitumen products. Data from Fingas (2014-2015; 2015d), Dew et al. (2015) and
Crude Quality Inc. (2015)
Product Description
Bitumen, Neatbit Undiluted extremely heavy oil extracted from oil sands. Must be heated to
be shipped
Diluent Any light petroleum used to dilute bitumen for transportation by pipeline or
rail; traditionally a condensate or ultra-light crude oil but now often a
refinery cut such as naphtha a
Synthetic crude oil (SCO) A liquid product made by partial upgrading or refining of bitumen; used as
a diluent in synbit
Dilbit Bitumen diluted with ~30% diluent, such as condensate or naphtha, for
pipeline transportation
Railbit Bitumen diluted with ~15% diluent (i.e., half as much diluent as dilbit),
typically for transport by rail tank car
Synbit Bitumen diluted ~50% with SCO
Dilsynbit Bitumen diluted with SCO plus another diluent, usually a condensate
(currently Albian Heavy Synthetic is the only dilsynbit transported)
Lightened Dilbit or C4/C5 Bitumen blended with a diluent supplemented with LMW alkanes such as
enhanced Dilbit C4 (butane) and C5 (pentane) b
a,
For some examples of condensate sources, see ISCO report #457, p. 8 (Fingas 2014-2015)
b,
Condensate is often in short supply in western Canada, in which case C4 and C5 may be used to replace a portion
of the diluent while maintaining pipeline viscosity specifications. Potential problems with this supplement are that
asphaltenes may precipitate during transport (Appendix A), and volatility and flammability increase (Section 2.2.3)

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 74
Blends, such as dilbit and diluted Bunker C, may exhibit ‘bimodal’ properties that are non-linear with
conventional crudes and therefore their behaviour can be difficult to predict based on chemistry alone. In
fact, diluted bitumen may have the most variable composition of any transported petroleum product
(Fingas 2015d) because of the composition and proportions of various diluents (natural condensates and
ultra-light oils, butane-enhanced condensate, naphtha, synthetic crude, etc.) that are used to achieve
pipeline transportation specifications of <300 mPa s and specific gravity <0.94 (Tsaprailis 2014; Fingas
2014-2015). Seasonal differences in diluent proportions and compositions (formulated to achieve
regulated viscosity at ambient temperature) make chemical definition of diluted bitumens even more
convoluted. This compositional variability gives the blends new properties that may not fully correlate
with volume proportions and further confounds prediction of their behaviour in the environment.

2.4 Weathering of Oil Spilled in Aquatic Environments

Weathering is a general term encompassing the


changes in petroleum properties brought about
Interactions of complex non-biological and
by physical, chemical and biological processes
biological processes lead to different fates of oil
when oil is exposed to environmental conditions
spilled in the environment. All of these are
such as in aquatic systems (Figure 2.3). These
influenced by oil chemistry, environmental
combined processes, along with the original oil
conditions and time, making prediction
chemistry and the time elapsed since the spill,
complicated and unique for each oil spill.
affect the behaviour, fate, chemical composition
and mass of residual oil after a spill event.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 75
Figure 2.3 Overview of processes affecting the fate and behaviour of oil spilled in freshwater and marine environments. The insets show processes at
different scales. Adapted from Daling et al. (1990), Tonina and Buffington (2007, 2009), McGenity et al. (2012), AOSRT (2014), NRC (2014), WSP
(2014) and Dew et al. (2015). Additional details on oil interactions with ice can be found in Figure 2.10.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 76
Weathering processes occur at different rates and with different onset times, resulting in progressive
changes in oil composition and behaviour after the spill (Figure 2.4). Some weathering processes, such as
evaporation, begin immediately and are most significant in the short-term (within hours or days); others
such as biodegradation occur after a delay or more slowly (over months or years). Therefore, gross
weathering rates are not constant following a spill and are generally highest immediately after the spill.
Moreover, weathering processes are not constant in all areas of a spill site. Oil at the surface of a
waterbody will experience certain processes more severely than oil below the water surface, beneath ice,
on the shoreline or at the edges of the spill compared to the thicker slick centre (as discussed below). As a
corollary, certain environmental factors are key to the rate of weathering, such as temperature, wave
action (energy), sunlight, suspended sediment in the water and microbial activity (as discussed in
Chapters 3 and 6).

Figure 2.4. Time of onset and relative importance of weathering processes over time after an oil spill onto water.
The onset and magnitude of effect will vary with temperature and for different oils (note the time scale, which
emphasizes the early onset of most processes). Figure adapted from AOSRT (2014).

It is also important to note that weathering is influenced greatly by the type of oil spilled. For example,
light crude oils with high proportions of LMW hydrocarbons spread on water more readily and therefore
are more susceptible to evaporation at the surface, whereas heavier oils have lower proportions of volatile
hydrocarbons and are more likely to sorb to suspended sediments and subsequently sink (Section 2.4.2.3).

General short-term physical and chemical weathering effects on oil (Table 2.4) are organized below
according to processes occurring at the water surface (fresh or marine water), in the water column, and
along the shoreline (riverbank or beach) or in underlying sediments. Notably, discussion of some
processes is skewed by the availability of data for certain well-studied oil spills (e.g., DWH blowout and
Exxon Valdez Oil Spill [EVOS]). Specific processes that depend on environment type and local factors at
the impacted sites, including biodegradation and remedial intervention, are discussed in Chapters 3 and 6.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 77
Table 2.4. Effect of natural weathering processes on oil properties (NRC 2014).
Oil property Natural weathering process Effect on oil property
Viscosity Loss of LMW components by evaporation and/or Increased viscosity
dissolution and/or biodegradation
Formation of water-in-oil emulsions, including Increased viscosity
‘mousse’
Specific gravity Loss of LMW components by evaporation and/or Increased specific gravity
dissolution and/or biodegradation
Volume of oil at Loss of LMW components by evaporation and/or Decreased volume
surface dissolution and/or biodegradation
Loss of small droplets by dispersion Decreased volume

Emulsification and/or biodispersion Increased volume


Potential toxicity Loss of LMW components by evaporation and/or Decreased acute toxicity
dissolution and/or biodegradation
Formation of photooxidation products at surface, or Increased toxicity
of partially oxidized metabolites from incomplete
biodegradation

2.4.1 Weathering Processes at the Water Surface

2.4.1.1 Spreading

When oil is spilled onto water and is allowed to spread unhindered, it moves away from the source where
the oil layer is thicker (a ‘slick’), forming a thin ‘sheen’ at the edges (Figure 2.3), e.g., the familiar
‘gasoline rainbow’ of 5-10 m thickness to nearly invisible sheens of <1 m. As it thins, the slick may
form patches or ‘ribbons’ (AOSRT 2014). Thus, spreading increases the spill area and decreases the
average thickness of the oil layer. In the case of a sheen, the area of oil contamination can appear to be
very large, but the actual mass of oil involved can be far smaller than in a slick. Spreading is influenced
by the oil viscosity, water and air temperature, and wind, wave and/or current action; in turn, spreading
affects evaporative losses (discussed below).

However, spreading is typically non-uniform, thicker ‘windrows’ (streaks of floating oil) separated by oil-
free water or sheen are commonly observed as the spill progresses (Simecek-Beatty and Lehr 2007).
These are understood to arise from mixing forces (Langmuir circulation) in the surface water (discussed
in Chapter 5). The action of such vortices (Langmuir cells) producing patchy oil slicks will affect the
efficiency of mechanical recovery of floating oil (Chapter 8).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 78
2.4.1.2 Evaporation, Aerosolization and Atmospheric Re-deposition

Evaporation can account for 75% of mass lost


from condensates and ultra-light oils and 20-
Among the significant weathering processes
30% of losses from light oils, but ≤10% from
occurring at the water surface, evaporation is
heavy oils (NRC 2003), reflecting the different
usually the most immediate and rapid and has the
concentrations of volatile hydrocarbons in these
greatest effect on the mass of spilled light and
different oil types. Saturates <C15,
medium crude oils.
monoaromatics (BTEX) and some small PAH
are particularly subject to evaporative loss
(AOSRT 2014) and 50% of hydrocarbons ≤ C16 were lost within 1 hour of an experimental oil spill by a
combination of evaporation and dispersion/dissolution (Gros et al. 2014). The thickness of the oil layer
also affects the rate of evaporation, with sheens evaporating more quickly than thick slicks and typically
following an exponential curve (Fingas 2015e). Thus, evaporation can significantly reduce the total spill
volume of a light oil. It also dynamically changes the composition of the surface oil which, subsequently,
changes behaviour of the residual oil. For example, because volatile hydrocarbons, such as BTEX, tend to
be the most acutely toxic, evaporation can reduce the acute toxicity of the spilled oil, but volatilization
creates a potential explosion hazard above the spill and potential breathing hazards for oil spill
responders. Simultaneously, with the loss of ‘light ends’, the residual oil becomes enriched in the HMW
compounds and therefore becomes more viscous and dense, as well as becoming enriched in alkyl PAH
that increase the chronic toxicity of the residual oil (Chapter 4). When the oil is unable to spread due to
confinement or low wind/wave action, such as on a pond, a ‘skin’ of resins and asphaltenes can form on
the surface, decreasing further evaporation of lighter components.

There is currently debate about the influence of surface winds over an oil spill. Fingas (2015c) pointed
out that the air boundary layer above an oil spill theoretically can regulate the rate of evaporation, but that
evaporative losses from light oils (e.g., ASMB, diesel, gasoline) in laboratory trials appeared to be limited
more by slick thickness (affecting diffusion of volatile molecules to the oil surface) than by simulated
wind speed. A previous report (Fingas 2004) indicated that temperature and time were greater factors in
oil evaporation than surface wind velocity or oil layer thickness for a wide range of crude oils in which
the bulk of hydrocarbons are >C10. In contrast, Gros et al. (2014) determined that wind speed strongly
influenced evaporation (as well as dispersion and dissolution) very early in an experimental spill of a
Norwegian crude in the North Sea.

Recommendation: More information is needed for early evaporation and dissolution processes in
actual oil spills to resolve present debate.

Whereas evaporation from a light or medium crude can be substantial (>30-50%), heavy oils experience
far less evaporative loss. AWB and CLB dilbits lost only 15-18 % mass under quiescent conditions in an
outdoor flume tank (King et al. 2014), and under laboratory conditions the latter lost a maximum of
17.4% mass at 22 °C (Waterman and García 2015). Synbit blends should experience even less
evaporative loss than dilbits because the diluent (synthetic crude) has higher average molecular weight
than the condensates or naphtha used in dilbit (Dew et al. 2015; Fingas 2015e). Diluted bitumen blends
show bimodal behaviour, with light diluent components evaporating at a rate that decreases with time as
the spilled oil becomes more viscous, slowing losses of volatile hydrocarbons. For example, elevated
benzene levels measured in the air after dilbit spilled into the Kalamazoo River, MI (Crosby et al. 2013)
(Chapter 3) represented evaporative losses from the diluent fraction. Dilbit that is initially buoyant
(specific gravity <1.0) becomes denser (≥1.0) during evaporation, increasing the potential for sinking
through the water column and subsequently contaminating sediments (Winter and Haddad 2014).
Therefore, the early period of a diluted bitumen spill may be very important for efficient recovery and
cleanup (Winter and Haddad 2014).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 79
Yarranton et al. (2015) recently showed in
laboratory tests that evaporation of diluent from Evaporative losses from dilbits are substantially
Cold Lake Winter Blend dilbit was comparable less than from light crude oils, but have a greater
regardless of whether it occurred from the effect on physical properties. As dilbit weathers, it
surface of a glass slide or fresh water; i.e., exhibits bimodal behaviour as diluent volatilizes
contact with quiescent water had no additional and bitumen dominates the chemistry of the
effect on evaporative losses and, furthermore, air weathered oil. Current questions include whether
flow had little effect on evaporation rate. They blending bitumen with a diluent yields a
determined that ambient temperature (5 °–25 °C) homogeneous fluid equivalent to a conventional
and dilbit film thickness (1.6–5.4 mm) were key heavy oil and, conversely, whether loss of diluent
parameters and that evaporation rate was limited restores dilbit to the original bitumen composition
by diffusion of the light diluent components and properties.
through the dilbit film, which, in turn, correlates
with changing density and viscosity of the
product during weathering. In contrast, the evaporation rate of the light reference oil ASMB was limited
by convective mass transfer of volatile components. This highlights the concern that existing evaporative
models of dilbit spills may not adequately predict evaporative behaviour of dilbit, and that further testing
using a range of diluted bitumen products is required for validation or refinement of evaporation models.
Additionally, these laboratory observations highlight the potential for sinking of diluted bitumen products
that reach densities >1 g/cm3 by evaporative weathering, and may inform prediction of adhesion of the
weathered product to surfaces or suspended sediments.

There is debate whether 100% of the diluent component of bitumen blends can be lost by natural
evaporation (Fingas 2015e) or whether the residual bitumen/heavy oil will retain some of the diluent
components as intimately blended constituents, conferring novel properties on the partially weathered oil
(Winter and Haddad 2014). This is particularly important for predicting if weathered dilbit, synbit, etc.,
will float or sink in water (Section 2.4.2.3). It seems plausible that some higher molecular weight
components of the diluent would be retained in the weathered oil, but at concentrations too low to
significantly change the physical behaviour of the residual oil compared with the original bitumen or
heavy oil stock. Although research has been initiated recently into the evaporative behaviour of various
bitumen blends under actively mixed conditions (King et al. 2015a), these data are not yet published and
further scrutiny is warranted. The observation of evaporative mass losses of <20% after rigorous
weathering at environmentally-relevant temperatures for dilbits nominally comprising ≥30% diluent
suggests that a substantial proportion of diluent remains intimately associated with bitumen.

The effect of evaporation on adhesion by dilbit is significant during the initial weathering phase when
diluent is being lost at the fastest rate. For example, AWB dilbit showed a rapid increase in adhesion
during the first 24 hours of natural weathering on seawater in open-air flume tanks (Figure 2.5). The
effect on intermediate fuel oil (IFO 180) adhesion was less and after 24 hours there was little change in
adhesion (or composition) of either oil.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 80
Figure 2.5 Effect of progressive natural weathering on adhesion. Unpublished data for Access Western Blend
(AWB) dilbit and an intermediate fuel oil (IFO 180) provided by Fisheries and Oceans Canada. Natural weathering
was achieved in an open-air seawater flume tank, and adhesion of the recovered oil was determined using the semi-
quantitative method of Jokuty et al. (1995).

Hollebone (2015) measured adhesion for several oils that experienced progressive evaporation in the
laboratory. A range of semi-quantitative adhesion values for selected oils is given in Appendix C, Table
C1 and a comparison of adhesion by selected light, medium and waxy crudes, CLB winter blend dilbit
and Bunker C fuel oil during controlled weathering is shown in Figure 2.6. Adhesion affected by
evaporative loss follows a continuum, with near-linear responses by light crude oils having high
proportions of saturates (Hollebone 2015) versus exponential responses by heavy oils with high resins
and asphaltene contents, consistent with observations by Jokuty et al. (1995). As expected, the sole dilbit
sample showed bimodal behaviour, with little initial change in adhesion followed by an exponential
increase similar to that of the heavy oils.

Recommendation: Further study on a range of oils under varied environmental conditions is


needed to resolve the current uncertainty around evaporative losses experienced in surface spills,
including the effect of oil layer thickness and surface winds, the role of photooxidation and
formation of weathered ‘skin’. Field work especially, but also lab work, are essential to determine
the extent and kinetics of diluent evaporation from various bitumen blends, and the resultant
behaviour of such blends due to weathering processes.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 81
Bunker C
1200
CLB Winter
1100
Sockeye Sour
1000
Sockeye
900
Maya B
800 Platform Gail
Adhesion (g/m2)

700 Lucula

600 Hondo

Hebron M-04
500
Hibernia
400
Terra Nova
300
Gullfaks
200
ANS middle
100 ASMB
0 West Texas Int.
0 5 10 15 20 25 30 35 40
Evaporative Mass Loss (%)
Figure 2.6 Correlation between adhesion and evaporation of selected oils. Data compiled from Hollebone (2015)
and unpublished data from Environment Canada (dilbit values); trend lines were drawn in Excel v.14.4.6.
Evaporative loss was achieved using rotary evaporation in a laboratory and adhesion was measured using the semi-
quantitative gravimetric method of Jokuty et al. (1995). Blue symbols, light crude oils; red, medium crude oils;
green, waxy crude oils; black, heavy crude oils; purple, dilbit (Cold Lake Blend-Winter); grey, Bunker C.

Aerosolization is a recently proposed dispersal route for atmospheric transport of crude oil spilled onto
water. Laboratory studies have shown that bubble bursting (e.g., due to ebullition of gases from a shallow
subsea blowout) and/or wave action (e.g., simulated white caps) can create hydrocarbon aerosols by
mechanisms similar to those that generate ocean spray (Ehrenhauser et al. 2014). Likewise, hydrocarbon
aerosols are formed by the impact of raindrops on an oil slick (Murphy et al. 2015). In both cases, the
presence of dispersant, such as Corexit 9500 used in the DWH spill (Section 2.4.2.1), enhanced the effect.
Notably, this dispersion route could also significantly increase evaporation from air-borne droplets.

Atmospheric re-deposition of the aerosolized oil (SL Ross 2012) onto water or land could distribute the
hydrocarbons beyond the extent of the main spill. Anecdotally, during the 1984 Uniacke G-72 surface
blowout off Sable Island, NS, which ejected a plume of condensate into the atmosphere, responders
reported ‘hydrocarbon rain’ near the spill site (K. Lee, pers. comm. 2015); a similar phenomenon
occurred during the 1982 onshore blowout of a sour condensate well near Lodgepole, AB (J. Foght, pers.
comm. 2015). However, literature on the magnitude and impacts of atmospheric hydrocarbons on water
and land is sparse (NRC 2003), warranting further study.

2.4.1.3 Photooxidation

In a poorly understood free-radical-generating process, aromatic hydrocarbons (particularly PAHs and


including aromatic N-, S- and O-heterocycles) react with oxygen in the presence of sunlight, yielding
oxygenated products that are more water-soluble and usually more resistant to biodegradation than the
parent compounds. Whereas removal of PAH from the oil is beneficial, as some PAHs contribute to its
potential carcinogenicity and embryotoxicity, the increased mobility and persistence of the photooxidized

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 82
products in the water column are likely detrimental. The full suite of photooxidation products is currently
unknown, although they are known to have impacts on aquatic life (Chapter 4). Notably, the susceptibility
of PAH to photooxidation is the opposite of biodegradation potential: the smaller the PAH the more
biodegradable and less susceptible to photooxidation. Conversely, the more aromatic rings, the less
biodegradable but more susceptible to photooxidation.

It is likely that photooxidation is most important relatively early in the weathering process. Recent
laboratory studies with fresh and weathered oils from the Heibei Spirit spill (a mixture of light and heavy
crudes; Yim et al. 2012) used ultra-high resolution GC to show that photooxidation preferentially
increased the proportion of oxygenated compounds in fresh oil versus weathered oil (Islam et al. 2005).
The N-containing resins (Table 2.1), particularly compounds with secondary and tertiary amine groups,
such as phenylamines, were most vulnerable to photooxidation. S-containing resins, such as aromatic
dibenzothiophenes, were also photooxidized, with unknown environmental consequences for toxicity and
persistence of the weathered oil.

Photooxidation is obviously affected by sun angle and season, particularly at higher latitudes in the
Arctic. However, even at lower latitudes the process appears to be slow, accounting for <0.1% of total
losses per day (WSP 2014). Although photooxidation may not directly achieve significant mass losses, it
can have indirect effects on subsequent weathering processes. For example, oxidized products may
contribute to emulsification (Fingas 2015e) and crust or ‘skin’ formation on the surface of an oil slick
(Bobra and Tennyson 1989) that decreases evaporative losses, as discussed above. Unexpectedly,
ultraviolet irradiation of light Macondo oil from the DWH oil spill affected formation of microbial flocs
in seawater compared with fresh Macondo oil (Passow 2014). This ‘photo-chemical aging’ appears to
have promoted subsequent sinking and sequestration of some of the oil in the Gulf of Mexico seabed
(discussed in detail in Section 2.4.2.3).

It appears that there may be only a small window of time for oil spill response to reduce the effects of
photooxidation on other weathering processes, but further research into the products and magnitude of
photooxidation is required, especially for emerging, unconventional oils and blends.

2.4.1.4 Emulsification

Formation of emulsions (Figure 2.7) is important for several reasons. By incorporating up to 60-80%
water, stable emulsions increase the effective volume of the spilled oil up to two to five fold and increase
the oil’s viscosity by up to 1,000-fold (Fingas and Fieldhouse 2015). This decreases evaporation of
volatile components and reduces oil spreading. Oil entrained in mousse emulsions resists dispersion with
chemicals and biodegradation due to its viscosity, low bulk surface area and hindered nutrient
replenishment (AOSRT 2014). Emulsions tend to move from floating on the water surface to being
submerged in the water column, where physical recovery may be hampered. All of these behaviours
affect cleanup and remediation response options (Chapter 6).

The type of emulsion formed, whether water-in-oil (w/o) or oil-in-water (o/w), depends on the
environmental conditions (temperature, mixing) and on the mass and chemical composition of the oil.
Some crude oils form w/o emulsions rapidly but soon revert to two discrete oil and water phases; others,
such as heavy fuel oils, form w/o emulsions poorly or only slowly. Some w/o emulsions can be stable for
months or years (Fingas and Fieldhouse 2015), likely stabilized by sub-fractions of asphaltenes (Yang et
al. 2014, 2015), waxes acting synergistically with asphaltenes (Kokal 2002), resins (WSP 2014), or even
by bacterial cells (Dorobantu et al. 2004) that prevent coalescence of the oil by aligning at the oil:water
interface (Figure 2.7). Eventually the emulsions may separate into oil and water again by natural
processes, including additional weathering, oxidation and/or freeze-thaw action (Fingas 2014).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 83
In light of the proposed roles of asphaltenes and
resins in stabilizing emulsions, it is both When sufficient mechanical energy (e.g., wind
interesting and unexpected that bitumen and and/or wave activity) is applied to mix oil and
severely weathered diluted bitumen that have a water, unstable or stable emulsions may be
high content of both classes do not appear to form generated; the latter may take the form of oil
stable emulsions (Fingas and Fieldhouse 2015), droplets dispersed in the bulk water phase or,
although lightly weathered bitumen blends can conversely, water droplets in a bulk oil phase
incorporate water droplets given sufficient commonly known as ‘chocolate mousse’.
mixing energy. During outdoor wave-tank tests
where two weathered dilbit products (AWB and
CLB, Table 2.2) were applied to seawater at an average temperature of 8 °C with and without artificial
sediment (fine clay minerals), breaking waves caused the oil to form unstable w/o emulsions and large
submerged droplets that readily coalesced into a surface slick (Government of Canada 2013). This
observation was confirmed by another study of fresh and weathered Cold Lake dilbit and ASMB light oil
(Table 2.2) agitated at 15 °C in an indoor wave tank with river water containing natural floodplain
sediment (Zhou et al. 2015). The ASMB immediately dispersed into the water column whereas the dilbit
formed an unstable emulsion that subsequently separated and remained floating for eight days or became
stranded on the apparatus ‘beach’; no formation of mousse was observed with the dilbit. Poor emulsion
formation by heavy oils and weathered dilbits may occur because high viscosity limits dispersion and
therefore the amount of water incorporated, and such water is only transiently entrained not actually
emulsified. It is notable that this unanticipated behaviour was revealed only by experimentation rather
than from first principles, highlighting the importance of research on unconventional oils. Further studies
on the fundamental mechanisms of dilbit interactions with water are underway to help address this
knowledge gap (e.g., Dettman and Irvine 2015), including the effects of emulsification on adhesion
properties (Section 2.2.6).

Recommendation: Laboratory and field trials are needed to determine the conditions under which
fresh and weathered diluted bitumen blends will form stable or unstable emulsions in fresh water
and seawater. A variety of bitumen blends would be essential to investigate under various agitation
conditions and salinities to inform modeling and response to spills.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 84
Figure 2.7 Emulsification of oil and water. Upper panel: Generation of stable and unstable water-in-oil (w/o) and
oil-in-water (o/w) emulsions in seawater. Lower panel: Stabilization of o/w and w/o emulsions in artificial fresh
water by cells of hydrocarbon-degrading bacteria. Lower panel adapted from Dorobantu, L.S., Yeung, A.K.C.,
Foght, J.M. and M.R. Gray. 2004. Stabilization of oil-water emulsions by hydrophobic bacteria. Applied and
Environmental Microbiology 70(10): 6333-6336.Copyright © American Society for Microbiology.

2.4.2 Weathering Processes in the Water Column

2.4.2.1 Dissolution

Evaporation and dissolution are interrelated and competing processes. In the former process, petroleum
constituents are moved to the atmosphere from the surface of the slick, and in the latter, they are diluted
in the water column from the underside of the slick. Both processes reduce the potential acute toxicity of
the residual oil while potentially increasing the chronic toxicity by enriching the remaining oil in alkyl
PAH (Chapter 4). However, evaporation typically accounts for far greater mass losses of spilled oil than
dissolution.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 85
Hydrocarbons typically have very low solubility in water
Dissolution is important because many (Section 2.2.4), which decreases with molecular size.
of the most water-soluble components Monoaromatics and the lightest saturates are the most
of petroleum (e.g., BTEX) are also the water-soluble, whereas larger saturates and PAH have
most volatile and most acutely toxic. lower solubility, and asphaltenes (and most resins) are
considered water-insoluble. For this reason, heavy oils and
bitumen do not readily dissolve in aqueous solution (the
water soluble fraction is <10 ppm) and dissolution does not significantly reduce the total mass of heavy
oils spilled onto water. However, LMW components of the diluents in dilbit and synbit will have some
aqueous solubility, in which case dissolution of even a small fraction of a very large dilbit spill may
represent a large mass of dissolved chemical. Dispersion of the oil as small droplets with large ratios of
surface area to volume would hasten and increase such dissolution.

The relative importance of dissolution increases in deep subsurface spills, such as the DWH blowout,
where the long travel time for oil droplets to rise or travel laterally through the water column increases the
opportunity for small molecules to dissolve rather than evaporate upon reaching the water surface (FISG
2010). Furthermore, subsurface application of the chemical dispersant Corexit 9500 to the DWH blowout
(Chapter 6) decreased oil droplet size, increased the total oil:water surface area to facilitate partitioning
into the water phase, generated a subsurface plume of droplets <100 m diameter that moved horizontally
with deep currents in the Gulf of Mexico, and prolonged oil contact with the water (Thibodeaux et al.
2011). This ‘continuous molecular extraction’ of oil droplets by the water column, which did not reach
equilibrium, preferentially depleted the oil of the most soluble monoaromatics, some resins and even
PAHs, alkyl PAHs and light alkanes <C8, and enriched the alkanes ≥C9 and heavier compounds in the oil
droplets. This phenomenon may have decreased the acute toxicity of dispersed Macondo oil and thereby
increased its biodegradability (Section 2.4.2.5 and Chapter 6).

Recommendation: There is a need for the development of new dynamic (non-steady state) models
describing dissolution of hydrocarbons in deep seawater and turbulent rivers, especially for diluted
bitumens where the fate of such hydrocarbons is not yet understood. Similarly, toxicity testing of
different water-soluble fractions (with and without dispersant addition) is needed for
unconventional oils and blends that exhibit bimodal behaviour.

2.4.2.2 Natural dispersion

Natural dispersion of oil in the water column (as opposed to addition of chemical dispersants as a cleanup
intervention; Chapter 6) occurs when the mechanical action of waves or turbulence detaches oil droplets
from the slick and forces them into the water column. Likewise, turbulent flow likely leads to oil
dispersion in rivers having steep gradients, high-velocity flows and/or boulder/cobble substrates, and
during offshore subsurface well blowouts like the DWH (Zhao et al. 2014). Depending upon droplet size,
depth and energy of the system, droplets may remain dispersed (i.e., suspended in the water column) or
may resurface with or without coalescing with other droplets (Figure 2.8). Droplets <20 m (0.02 mm)
may remain stably dispersed in water without resurfacing for a long time (Fingas 2014b), as observed
with the DWH oil (Valentine et al. 2014).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 86
oil slick
re-surfacing
unstable
dispersion
coalescence
stable dispersion
current drift

sinking

Figure 2.8 Natural dispersion of oil into the water column.

Dispersion affects oil behaviour in several ways. First, stable dispersion removes oil from the surface,
reducing evaporation and photooxidation. Second, the greater surface area afforded by small oil droplets
makes the oil more available for dissolution (Section 2.4.2.1) and biodegradation (Brakstad et al. 2015;
Prince 2015; Section 2.4.2.5). In fact, the rate of oil biodegradation has been suggested to correlate with
oil droplet size distribution (Venosa and Holder 2007). Moreover, many oil-degrading microbes are
attracted and attach to the oil–water interface (Kang et al. 2008a,b; Abbasnezhad et al. 2011a,b) where
they may: (a) produce biosurfactants to further increase droplet formation; and/or (b) stabilize oil droplets
by preventing coalescence, as described for emulsions (Dorobantu et al. 2004; Figure 2.7). Oil droplet
size is a factor in toxicity to aquatic organisms, as zooplankton may ingest tiny droplets and thus pass the
hydrocarbon up the food chain, and smaller droplets (with a higher ratio of surface area to volume)
facilitate dissolution of hydrocarbons associated with toxicity (Figure 2.3 and Chapter 4).

Although dispersion of conventional crude oils is relatively well understood (see Daling et al. [2003]),
considerably less is known about dispersion of bitumen blends in fresh water, estuarine (brackish) water
or seawater (Dew et al. 2015). Recently, dispersion of CLB dilbit in filtered seawater was conducted in a
wave tank under spring-time and summer temperatures (~8.5 °C and 17 °C, respectively), with or without
addition of fine mineral particles (King et al. 2015b) and with or without chemical dispersant. With
breaking waves, the dilbit showed poor natural dispersion (only 6% dispersion effectiveness) in the
presence and absence of suspended fine mineral particulates. Addition of chemical dispersant alone
efficiently produced droplets (45-59% dispersion effectiveness), but the droplet size distribution differed
from that of chemically-dispersed conventional crude oils. Additional studies of dilbit dispersion are
underway in Government of Canada laboratories, but data are not yet available for review.

Recommendation: More data are required for natural and chemically-enhanced dispersion of
unconventional oils and diluted bitumen blends, especially in fresh water.

2.4.2.3 Submergence, sinking, sedimentation and re-suspension of oil

Multiple processes can alter oil density to influence submergence, sinking and sedimentation, including:
increased density due to evaporation and/or dissolution of light components (Section 2.4.1.2 and Section
2.4.2.1); emulsification (Section 2.4.1.4); and interactions with particles. The latter process can be further
delineated by considering inorganic particles (e.g., minerals and detritus suspended offshore, or onshore
sediments if the oil spills over land before entering the water body) or organic particles (e.g., flocs of
microbial biofilms, biological detritus and/or phytoplankton mucus).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 87
The density of many crude oils, including diluted
bitumen products, is less than that of both fresh Oil may submerge through dispersion or
water (~1 g/cm3) and sea water (~1.03 g/cm3) and emulsification and eventually sink through the
therefore these oils will float when spilled. water column to become entrained in the
However, bitumen by definition has a density >1 underlying sediments (marine or freshwater) in
g/cm3 and, if diluent evaporates from dilbit during a a process called ‘sedimentation’. If disturbed,
spill, the residual bitumen will become more dense sunken oil may become re-suspended to enter
and may become submerged or sink. Even so, oil the water column again, resurface and/or be re-
droplets of density ~1.0 can be neutrally buoyant distributed in the subsurface.
and remain suspended in the water column,
especially if present in turbulent water such as in a
rapidly flowing river or high wave energy system. In rivers or at beaches, suspended oil droplets could be
entrained into porous gravel by hyporheic flows (Figure 2.3), creating a potential for contamination of
sediments and pore water with free-phase oil or dissolved hydrocarbons (reviewed in more detail in
Chapters 3 and 5). In addition to evaporation, other environmental processes, such as photooxidation
(Section 2.4.1.3), can affect oil buoyancy (NRC 1999; Government of Canada 2013). As water
temperature generally decreases with water depth, the oil density will increase, possibly causing it to
remain submerged or sink, and temperature cycling between day and night may cause oil to alternatively
submerge and float. Emulsification can also cause floating oil to become submerged (but typically not to
sink).

Interaction of oil with non-oil particulates changes


the properties of the aggregates, causing normally
OMAs form when oil interacts with and buoyant oil to sink, for example, or liquid oil to
adheres to inorganic particles. A more general become semi-solid. ‘Oil-Mineral Aggregates’
term, OPAs, includes OMAs plus oil associated (OMAs) can form when oil interacts with
with organic particles. inorganic materials, such as clay minerals
suspended in the water column (Lee 2002), or
with soils or shoreline sediments during overland
flow into a waterbody. The term oil-particle aggregates’ (OPAs) collectively encompasses OMAs as well
as oil associated with organic material, such as detritus and living microbial cells (Fitzpatrick et al. 2015;
Waterman and Garcia 2015), and combinations of these particles (Figure 2.9). OPAs may have neutral
buoyancy and float at or below the water surface, or may sink and become entrained in sediments. The
latter fate would potentially place the oil in anaerobic conditions where biodegradation is slower (Chapter
3), although shallowly buried oil may be re-suspended by wave action or by ebullition of gases from the
sediments. If the OPAs remain suspended, the additional reactive surface area afforded by the minerals or
organic matter may enhance bacterial attachment (Figure 2.3) and aerobic biodegradation (Chapter 6).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 88
A

Figure 2.9 Oil-particle aggregate (OPA) formed by interaction of weathered dilbit with natural Kalamazoo River
sediments. A, appearance of an OPA under ultraviolet light; yellow areas are oil. B, the same OPA observed under
visible light, showing oil plus sediment particles. Scale bars in bottom right are 50 m. Images from Lee et al.
(2012).

OPAs can be generated by overland spills where oil initially contacts sediments on shorelines or
riverbanks before entering the water column and sinking. Alternatively, oil that strands on beaches,
interacts with shoreline sediments, then re-enters the water column with accumulated mineral particles
may sink. Another natural mechanism for generating OPAs in the water is association of floating or
submerged oil droplets with suspended particles in the water column, especially near the shore or in
estuaries where the suspended sediment load is high. In fact, incorporating as little as 2–3% mineral
content into the oil phase is sufficient to cause some oils to sink (NRC 1999). The degree of interaction
between oil and minerals depends on oil viscosity and the mineral type, but may begin with formation of
an oil monolayer around the mineral followed by further accumulation of oil on the hydrocarbon surface
(Omotoso et al. 2002).

The importance of submergence, OMA or OPA formation and sedimentation in the recovery and overall
fate of spilled oil was highlighted in the DWH subsurface blowout in which chemical dispersants were
injected into the escaping plume of light Macondo oil (Table 2.2). The large initial subsurface plume of
suspended dispersed oil droplets (<100 m diameter) at depths of 900-1,200 m was undetectable a few
months after the discharge and was initially presumed by some researchers to have been rapidly
biodegraded by indigenous bacteria (e.g., Hazen et al. 2010) that had been enriched by natural oil seeps in
the Gulf (Figure 2.3). However, two independent analytical methods (hopane biomarkers and natural
abundance radioisotopes) applied to sediment cores collected from the vicinity of the blowout have
recently revealed elevated concentrations of hydrocarbons patchily distributed in a thin sediment layer
overlaid by cleaner sediments from post-spill deposition (Valentine et al. 2014; Chanton et al. 2015; Joye

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 89
2015). This sequestered oil within a 3,200 km2 footprint on the Gulf seabed may account for 2-14% of the
total volume of discharged Macondo oil (Valentine et al. 2014; Chanton et al. 2015; Joye 2015).

Two complementary mechanisms are proposed to have promoted this sedimentation pattern (Thibodeaux
et al. 2011). One is the formation of OMAs as the circulating subsurface oil plume interacted with
suspended mineral particles near the continental shelf, the concentrations of which may have further
increased when flow from the Mississippi River was diverted to flush Macondo oil out of shorelines
(Passow 2014). Sinking of these OMAs to the seafloor (Figure 2.3) is proposed to have led to a ‘bathtub
ring’ of oily sediments around the Gulf’s continental shelf (Valentine et al. 2014).

A second mechanism of oil submergence, sinking and sedimentation, presumed to be biologically-driven,


was also observed by analysis of Gulf sediments and in laboratory simulations. In fact, this phenomenon
may account for a large proportion of oil ‘missing’ from initial oil budgets for the DWH spill (e.g., FISG
2010; Joye 2015) and may be more important than OMA sedimentation in this particular spill. Creation of
an oily ‘footprint’ or ‘fallout plume’, predominantly of microbial origin (Passow et al. 2012; Passow
2014; Valentine et al. 2014) may have formed as follows: even in uncontaminated ocean waters, marine
phytoplankton and zooplankton near the sea surface naturally secrete sticky polymers that aggregate
particulate and dissolved organic material into >0.5 mm flocs called ‘marine snow’ (Azam and Malfatti
2007; Kleindienst et al. 2015). These flocs slowly settle to the seafloor over months or years to form new
sediments enriched in recalcitrant organic matter. During and briefly after the DWH event, unusually
elevated concentrations and sizes of such flocs (mm to cm diameter) were observed (Brooks et al. 2015).
These are proposed to represent blooms of hydrocarbon-degrading bacterial communities attached to
weathered Macondo oil droplets at or near the sea surface (Figure 2.3). Additional organic material in the
flocs may be mucus produced by oil-stressed phytoplankton. These large, dense, oil-rich organic flocs
rapidly sank over days or weeks while incorporating additional submerged oil in the subsurface plume
(and possibly mineral particles), creating a transient ‘dirty blizzard’ of oil-laden marine snow (Passow
2014; Valentine et al. 2014; Brooks et al. 2015). It is possible that some degree of aerobic oil
biodegradation occurred during sinking of the OPAs, but the fallout area created an ‘oily footprint’ on the
Gulf seafloor. Thus, both surface and subsurface oil are thought to have been transported to the seabed in
microbe-rich aggregates.

Weathering of Macondo oil at the surface, including photooxidation, affected floc formation (versus fresh
Macondo oil that did not promote floc formation) and may have enhanced the rate of sinking of the oily
marine snow (Chanton et al. 2014; Passow 2014). The role of Corexit 9500 dispersant in floc formation
appears to be concentration-dependent (Passow 2014). Furthermore, dispersion of Macondo-like oil with
Corexit 9500 may lead to enrichment of specific Colwellia subtypes capable of growing on the dispersant
itself and indirectly influencing floc formation (Kleindienst et al. 2015). These authors have recently
claimed, based on laboratory experiments, that Corexit 9500 may suppress oil biodegradation by
indigenous hydrocarbon-degrading microbes in deep Gulf waters, although divergent opinions have been
expressed regarding the effects of chemical dispersants on oil biodegradation in the open ocean (e.g.,
Prince 2015).

These types of hydrocarbon-rich sediment layers were gradually overlaid by clean sediments and post-
spill marine snow through natural deposition processes, sequestering the oiled sediment layer (Romero et
al. 2015). The long-term fate of this buried light Macondo oil is not yet known. However, it is possible
that the sequestered oil will be biodegraded (even under anaerobic conditions in the sediment layer)
because it is a light oil rich in alkanes, and because dissolution during initial rise of droplets to the surface
(Section 2.4.2.1) and subsequent surface weathering (evaporation, photolysis) may have reduced the oil’s
toxicity. Also, attachment of active hydrocarbon-degrading microbes during floc formation may have
‘inoculated’ OPAs to promote biodegradation during and after sedimentation; however, this possibility
remains to be investigated.

Formation of OPAs likely contributed to sedimentation of a diluted Bunker C-class heavy fuel oil spilled
into Lake Wabamun, AB, by a train derailment in 2005 (Chapters 4 and 8). During the overland flow of

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 90
Even the relatively simple, well-characterized Macondo light crude exhibited unexpected
sinking behaviour when dispersed. The submergence and sedimentation of chemically
complex heavy oils and diluted bitumen blends is even more unpredictable, especially since
dispersion of both fresh and weathered dilbits is not yet understood. Furthermore, there are
very few published reports on formation of OPAs from heavy oil or bitumen products,
especially in fresh water. This hampers understanding and modeling of dilbit submergence,
incorporation into sediments, re-emergence and transport in rivers. The role of microbial flocs
in sedimenting heavy residues into shallow receptors (e.g., river sediments and lakebeds) is
unknown, as this mechanism was not recognized until very recently from DWH studies.

this heavy oil into the lake the lighter diluent components evaporated, allowing the weathered oil to
interact with shoreline material and subsequently with suspended organic particles in the water column. In
addition to forming tar balls (Section 2.4.2.4), some oil sank, impacted the lake sediments, and continued
to resurface over subsequent years (Hollebone et al. 2011). Floating, sinking, submerged, sedimented and
refloating oil, including the production of sheens, were all observed in the year following the spill
(Hollebone et al. 2011).

Currently there is considerable debate regarding the fate of weathered and unweathered diluted heavy oils
and bitumen blends: will they float on the surface, submerge, sink or sediment with OPAs? An early
literature review (NRC 1999) found that only 20% of heavy oil spills resulted in a significant portion of
the products sinking or being submerged in the water. However, a more recent review (Winter and
Haddad 2014) concluded that dilbit has a greater potential for sinking than conventional oils. First,
evaporation of the diluent could increase the oil density, causing the residual oil to submerge or sink.
Second, in theory, the majority of weathered dilbit would be available for interaction with suspended
sediments as OMAs because so little of the bitumen itself will evaporate or biodegrade, further increasing
the possibility of sedimentation and sequestration. It is not yet fully known how adhesive properties
(which change with evaporative losses; Figures 2.5 and 2.6) affect OPA formation by diluted bitumen
products in the field.

Field experience with the highly-publicized overland spill of dilbit into the Kalamazoo River, MI, in 2010
(Chapters 6 and 8) indicated that the dilbit, with an initial density of 0.96-0.98 g/cm3 and near neutral
buoyancy, initially floated. Interaction with soil particles before entering the river and subsequently with
elevated suspended sediment loads in the water due to high rainfall events preceding the spill enabled a
considerable proportion of the dilbit to become submerged in the river water and eventually sink (Crosby
et al. 2013). This phenomenon contaminated the river sediments at natural collection points. The oil-
sediment aggregates were stable for at least two years (Lee et al. 2012), and at three years post-spill were
estimated to represent 20-30% of the initial oil volume (US-EPA 2013b).

This case supports the statement by Crosby et al. (2013) in a report from the US National Oceanographic
and Atmospheric Administration (NOAA) that: ‘Little research is currently available regarding the
behavior of oil sands products spilled into water, and how they weather in the environment. Most tests
have been conducted in the laboratory, so predicting the actual behavior of oil sands products for a range
of spills is difficult.’

Recommendation: Additional study is needed to examine OPA formation by dilbits and heavy oil
samples that have experienced various stages of weathering (including photooxidation, which may
be important for heavy products that are enriched in PAH), using various types and loads of
suspended sediments at various temperatures and salinities.

Some research is beginning to emerge to address this gap. Waterman and Garcia (2015) performed
laboratory studies with CLB dilbit (Table 2.2) and Kalamazoo River sediments to examine the kinetics of
OMA formation. The dilbit was artificially weathered (without UV exposure) over six days to achieve
different mass losses (0%, 9.9% or 17.4% loss; density range 0.932–0.993 g/ml), then added to the
surface of tap water containing uncontaminated Kalamazoo River sediments of different mesh sizes. With

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 91
conventional gyrotory mixing at 21-22 °C the dilbit formed stable small-diameter (<200 m) spherical
droplets coated with particles (similar to those shown in Figure 2.9 and in Stoffyn-Egli and Lee [2002]).
However, larger (>1 mm) irregularly-shaped ‘solid’ OPAs formed with unconventional mixing intended
to simulate sinking, re-suspension and droplet coalescence, possibly from folding of the sediments into
the oil rather than simple attachment to the surface of droplets per the mechanism proposed by Stoffyn-
Egli and Lee (2002). Some OMAs of each type were positively buoyant (floated in tap water), whereas
others sank, depending on the relative proportions of oil and incorporated sediment. Thus, the behaviour
of diluted bitumen is not predictable without experimental data to inform predictions.

Other recent laboratory studies in seawater found that weathered dilbit achieved densities >1 g/cm3 and
furthermore adhered to fine sediments (kaolin clay) added to the water, forming OMAs that sank in
saltwater, as well as forming floating tar balls (Government of Canada 2013; reviewed by Dew et al.
2015). Additional details about floating and sinking behaviours of AWB and CLB dilbits in saltwater can
be found in a Government of Canada (2013) report. The authors noted that bitumen is predicted to sink in
fresh water after extensive weathering, but the behaviour of dilbit during weathering in the marine
environment is currently poorly delineated. The relevance of using laboratory conditions to simulate oil
behaviour in the environment has been questioned, as some studies have failed to use UV irradiation (to
cause photooxidation) and/or to consider microbial interactions or temperature effects on density and
sinking (e.g., weathered bitumen may float at 15 °C but sink at 5 °C; Section 2.2.1). As reviewed by Short
(2013), other studies used unrealistically thick dilbit slicks or inappropriately slow wind speeds. To begin
to address the need for open-air studies, experiments were performed on two fresh (unweathered) dilbit
samples (AWB and CLB; Table 2.2) in outdoor wave tanks containing filtered seawater at ~19 °C (King
et al. 2014). In only six days the AWB dilbit weathered sufficiently for droplets to detach from the slick
and sink in the upper 10 cm of brackish water. This sinking was not enhanced by OMA formation
because the seawater had been filtered to remove particles >5 µm. The density of CLB dilbit increased
more slowly over 13 days, likely because of its naturally higher concentration of alkyl PAHs, which are
more resistant to weathering (King et al. 2014).

Despite increased interest and research activity, such as laboratory and field tests that are currently
underway in government laboratories (the results of which are not yet available for review), the
conditions under which dilbit will float, submerge or sediment constitute a knowledge gap that warrants
further investigation, particularly for freshwater systems. For example, the magnitude of microbially-
driven sedimentation is currently cryptic, with or without photooxidation, OPA formation and different
combinations of other weathering processes. If this sedimentation process is as significant for dilbit
spilled in shallow water and freshwater systems as it appears to have been in the light Macondo oil
subsurface spill, it is a major process that should be incorporated into oil spill models. The bimodal
evaporation kinetics of dilbit and the longer time needed for onset and action of microbial processes to
affect sinking versus physical processes (Figure 2.4) suggest that experiments on unconventional oils
should be conducted for much longer time periods than trials with conventional crude oils.

When oil sinks and becomes sequestered in sediments, it is less likely to be quickly or extensively
biodegraded (Section 2.4.2.5) because the sediment pore spaces are usually anaerobic below the interface
and nutrient replenishment is hindered. The fate of sequestered OPAs is largely unknown and requires
study over long time periods (years) to account for the slow rates of anaerobic biodegradation. In the case
of sequestered weathered dilbit and heavy oil, it is possible that the oil becomes virtually chemically-inert
to benthic organisms, although it may still have physical impact.

Recommendation: The long-term biodegradation of sedimented and sequestered oil in anaerobic


marine and freshwater sediments requires long-term study (i.e., over years to decades) to provide
reliable information for modeling and oil spill budgets.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 92
2.4.2.4 Formation of tar balls, tar patties and tar mats

Tar balls are typically discrete, roughly spherical conglomerations of oil <10 cm in diameter, tar patties
>10 cm in diameter, and tar mats ≥ 1 m in diameter (Fingas 2015b) that may be partially or completely
submerged (Warnock et al. 2015). Pelagic tar floats or is
shallowly submerged (and can travel considerable Weathered oil that has acquired a surface
distances with waves and currents, where it can become ‘crust’ through weathering, emulsification,
stranded on the shore), whereas benthic tar residues are photooxidation and/or sediment interaction
immobilized by sinking to the sea floor. The impact of may form spherical tar balls, flattened tar
marine tar residues on aquatic life is less than that of mats or patties (collectively, marine tar
fresh oil in the water column, although tar balls may residues) that may float, submerge or sink,
serve as a reservoir of PAHs that partition into depending on the oil type and severity of
surrounding water (Martin et al. 2014), and tar residues weathering.
that strand on beaches may release sheens of nearly-
fresh oil if physically disrupted.

Marine tar residues can be produced from ‘chocolate mousse’ emulsions (Section 2.4.1.4) and, due to
entrained water, may harbour microorganisms within the structure, as well as at the surface. However, the
small surface area:volume ratios of tar residues, as well as the presence of a crust, hampers
biodegradation of the oil (Warnock et al. 2015). The crust may be enriched in asphaltenes and resins as
these are the most polar compounds in oil and are prone to alignment at the oil:water interface; therefore,
heavy oils are more likely to form tar balls and mats. If suspended, the balls and mats may intermittently
release a sheen of fresh oil from breaches in the crust (Hollebone et al. 2011). Small tar mats may persist
for decades, as noted after the 1979 Ixtoc I blowout in the Gulf of Mexico and the Prestige oil tanker
sinking off Spain in 2002 (reviewed by Warnock et al. 2015). In 2013 an enormous tar mat estimated at
18 tonnes was discovered in the Gulf of Mexico as a consequence of the DWH subsurface spill in 2010. It
consisted of 15% oil and 85% sand, shells and water (Buskey 2013).

2.4.2.5 Biodegradation

Oil spill biodegradation processes and


Biodegradation is a natural process in which living limitations are discussed in detail in Chapter
organisms and/or their enzymes break down organic 3 and bioremediation in Chapter 6. Briefly,
material to produce simpler chemical compounds oil susceptibility to biodegradation primarily
such as organic acids, alcohols and/or gases. depends on: chemical composition (ratio of
Bioremediation is the application of biodegradation biodegradable alkanes and aromatics to
processes to help clean up a polluted site. Petroleum recalcitrant resins and asphaltenes); physical
biodegradation in aquatic environments is primarily state (surface area available at the oil:water
achieved by bacteria. interface for microbial attachment or
dissolution of light hydrocarbons and hence
also dependence on spreading, dispersion
and emulsification state of the oil);
temperature (biochemical activity being slower at low temperatures); nutrient availability (particularly
soluble nitrogen and sometimes phosphate, as discussed in Chapter 6); and available electron acceptors
for redox reactions (aerobic or various anaerobic conditions). Regarding the latter, oil degradation is
generally faster and more efficient under aerobic conditions, such as in a well-aerated water column or in
the uppermost layer of sediment, but may also occur slowly in buried sediments under anaerobic
conditions (Section 2.4.3.1). The range of susceptible hydrocarbons under anaerobic conditions appears to
be smaller than with aerobic degradation (e.g., reviews by Foght 2008; Mbadinga et al. 2011) and the
enzymes, pathways, end products and residual oil composition also differ.

Oil biodegradation also requires competent microbes. Hydrocarbon-degrading bacteria are ubiquitous,
and although they may be present in a pristine environment in only small numbers, they can flourish after
an oil spill because they have an advantage over microbes that cannot utilize hydrocarbons. Environments

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 93
that are routinely exposed to oil spills or natural hydrocarbon seeps, such as in the Gulf of Mexico,
usually have indigenous microbial communities that are enriched in hydrocarbon-degraders and thus can
respond more quickly to an incursion of oil. However, the widespread synthesis of alkanes by marine
cyanobacteria may help maintain a baseline level of competent hydrocarbon-degrading bacteria even in
pristine waters (Lea-Smith et al. 2015).

Notably, because biodegradation and physical weathering processes, such as evaporation and dissolution,
target many of the same compounds (light hydrocarbons) but at different rates, it is often important to
discriminate between biological and non-biological changes to oil chemistry. This can be done by the
judicious selection of reference samples and the use of biomarkers (e.g., Prince et al. 1994; Blenkinsopp
et al. 1996; Wang et al. 1998). As with chemical weathering, biological weathering is greatly influenced
by oil composition, and each oil has different susceptibilities to aerobic and anaerobic biodegradation
under specific environmental conditions. For this reason, surveys of oil biodegradability are useful for
predicting oil fates. An example is the Environment Canada report by Blenkinsopp et al. (1996) in which
a suite of oils transported in Alaska was tested in a laboratory for relative susceptibility to biodegradation
by a standardized microbial consortium under cold freshwater conditions, with extended incubation times.

Recommendation: To complement catalogues of chemical and physical properties of oils, a


biodegradability survey should be conducted at various “reference sites” for conventional and
unconventional oils commonly transported in Canada. These surveys should consider aerobic
biodegradation potential under fresh- and saltwater conditions at cold and moderate temperatures,
and anaerobic biodegradation in water-saturated sediments. Similar studies conducted under
optimum laboratory conditions with fresh and weathered oils can be used to delineate the
maximum biodegradation rates to be expected under actual environmental conditions.

The paucity of research on biodegradation of heavy oils and diluted bitumen must be addressed. It is
certain that the majority of the bitumen in the blends is highly recalcitrant to biodegradation (Section
2.1.1.3 and 2.1.1.4). Even if some of the constituent bitumen chemicals were altered, would that change
be detectable using current analytical methods (Appendix A) that are inadequate for characterizing
asphaltenes and resins? Theoretically, the diluent components of bitumen blends should be relatively
biodegradable (and resolvable by chromatography), but few blends have been subjected to controlled
biodegradation experiments either aerobically or anaerobically to ascertain what proportion of the diluent
is degradable in fresh or weathered dilbit. This uncertainty arises from the question raised in Section
2.3.1.2: does all diluent evaporate from diluted bitumen under natural conditions, leaving only recalcitrant
bitumen? If so, there will be very little light hydrocarbon mass to be biodegraded in weathered dilbit. If
not, is any of the residual diluent ‘bioavailable’ to microbes? It is possible that diluent chemicals are
intimately associated with the resins and asphaltenes and therefore not available for microbial attack.
Additionally, the diluent may be protected from microbial attack because the high viscosity of the
weathered bitumen affords only a small surface area for microbial access.

Recommendation: Biodegradation of different bitumen blends and diluted heavy oils should be
tested under a range of laboratory incubation combinations and analyzed using complementary
suites of techniques to determine whether and how much biodegradation of constituents can be
achieved. Comparison to multiple control conditions will be essential for rigorous interpretation of
the results.

2.4.3 Oil Interactions with Shorelines and Sediments

Oil can impact shorelines and sediments via different routes. Overland spills will first impact the soil or
beach so that the oil may be partially weathered by the time it reaches the water, as well as having
incorporated mineral and organic matter (as discussed in Section 2.4.2.3). Oil that is spilled onto water
and is subsequently stranded on the shoreline or riparian zone (the interface between a river or stream and
the land) or buried in sediments will have different fates, including interaction with vegetation and
possible re-mobilization.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 94
2.4.3.1 Sorption, penetration and sequestration

Oil is adhesive, particularly when weathered (Fingas 2015b; Figure 2.6) and may attach to suspended
particles (Section 2.4.2.3) and surfaces. If OPAs sink and are buried by additional sediment, the oil may
be protected from remobilization, and biodegradation rates will be limited in instances where the water in
sediment pore spaces is not adequately replenished with oxygen or nutrients. That is, oil buried in
sediments, such as found in mudflats or deep ocean, may persist virtually unchanged.

Oil that remains floating or submerged in the water column can reach the shore and adhere to sand,
cobbles, bedrock and shells, as well as man-made structures like piers and jetties; the same behaviour
might be expected in the riparian zone. Once onshore in the intertidal zone, a light oil may penetrate by
gravity into beach sediments due to its low viscosity, where it may sorb to the minerals and/or be retained
(sequestered) in small pools, with reduced exposure to further weathering. This was observed after EVOS
at many cobble beaches in Alaska. Relatively fresh oil can still be found in patches below the surface
cobble and gravel more than 25 years after the spill (Atlas and Hazen 2011). This persistence is due in
part to poor access of oxygen and nutrients to the buried oil, limiting remobilization and biodegradation
of the oil (discussed in Chapter 3), as well as limited photooxidation, emulsification and other weathering
processes (Boufadel et al. 2010).

Viscous heavy oils or highly weathered oils are less likely to penetrate deep into intertidal sediments, but
may be forced to depth by wave action on high-energy beaches. Weathered oil that is thrown above the
tidal zone will continue to experience physical and chemical weathering and may form an ‘asphalt
pavement’, as observed on Chedabucto Bay beaches, persisting decades after the 1970 Arrow spill in
Nova Scotia (Advanced Technology and Continental Shelf Associates 1990), and after the Baffin Island
Oil Spill field experiment in the Canadian Arctic (Owens et al. 1987).

Although it is generally assumed that spills of light oils, such as condensate, in the open ocean would
cause little or no harm due to rapid losses by evaporation, coastal studies in Nova Scotia have shown that
condensate entrained within intertidal sandy beach environments can persist for extended periods (Strain
1986; Lee and Levy 1989). The implication is that evaporation of the diluent portion of dilbit that is
sequestered in beach sediment may be retarded if unweathered dilbit is buried.

On low-energy beaches, cobble and boulder


Recent field observations and laboratory tests ‘armour’ (Figure 2.3) prevents the sediments
suggest that ‘armoured beaches’ (those that have from being mixed by wave action and
boulders and/or cobbles underlain by fine sediments) therefore protects the sequestered oil beneath
may retain viscous heavy oils and heavily weathered (Harper et al. 2015). This oil, retained in the
oils, with the oil penetrating only to the surface of upper fine sediments, would be alternately
the packed sediment but not further. replenished with oxygen and nutrients by
tidal action and might undergo limited
evaporation and dissolution, but the low
surface area:volume ratio could slow biodegradation rates.

The presence of aquatic ‘biofilms’ (layers of living or recently living plants, animals and microbes) on
rocks and structures could potentially reduce the adherence and sequestration of oil, particularly if the
living material is highly hydrated (e.g., wet algae). The prevalence of algae on exposed bedrock
shorelines is a component of ongoing baseline surveys in Canada that may yield observations about
biofilm protection of substrata from oiling (Laforest et al. 2015).

There is much interest in the fate and behaviour of oil that enters aquatic areas having extensive
vegetation (e.g., wetlands, estuaries, riparian zones) and impacts the underlying sediments. In this case,
rates of submerged oil biodegradation may be retarded by anaerobic conditions in saturated sediments or
by diauxic growth conditions (i.e., utilization of elevated levels of readily-degraded organic detrital
matter by microbes, in preference to recalcitrant or toxic hydrocarbons). Alternatively, oil biodegradation

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 95
may be stimulated by microbial associations with plant roots and the rhizosphere (soil surrounding roots
and root hairs) in the process of ‘phytoremediation’ (Figure 2.3; Chapter 6). The latter interactions could
also potentially contribute to biodegradation of components of the stranded residual oil in the sediments.
Toxicity of dilbit spills to aquatic vegetation represents a current knowledge gap that has only recently
begun to be studied (Dew et al. 2015).

2.4.4 Perspective

Based on review of the literature and on interviews conducted by the Expert Panel, it is clear that there is
currently insufficient information about the chemical composition and environmental behaviour of
emerging petroleum types, including diluted bitumen products, synthetic crude oils and perhaps shale
oils. The question was raised by the Panel and others whether unconventional oils, such as diluted
bitumens, are sufficiently different from conventional oils to be relegated to distinct classes for regulatory
or spill response purposes. Based on current analytical capabilities, heavy oils and bitumen appear to fall
along a recognizable chemical continuum spanning condensates to bitumen. However, in the case of
diluted bitumen blends it is their physical behaviour in the environment, such as bimodal weathering, that
currently appears to distinguish them from conventional petroleum and to confound extrapolation of their
properties from the two components: diluent and bitumen. This lack of knowledge stems from insufficient
information from laboratory tests, wave tanks or field trials to use in comparing bitumen blends with
conventional oil behaviour, particularly in fresh water, in cold conditions and over prolonged time in the
environment. For example, the basic question of how much diluent will evaporate from a particular
diluted bitumen under environmental conditions has only begun to be addressed in laboratory and field
trials. Emulsification of diluted bitumens is beginning to be tested under a variety of conditions, but much
of the work to date has focused on seawater emulsification, with little work on fresh water. Similarly, new
dynamic models are needed for predicting dissolution and dispersion of diluted bitumen, and there is a
paucity of controlled trials showing the effects of particulates on dilbit sinking and sedimentation
behaviour, whether accumulated from soils in overland spills or from interaction with suspended solids
and microbes in the water column. Some of these parameters are currently being tested in wave tank
studies with natural seawater, but the data are not yet available for the Panel to review. The US National
Academy of Sciences (NAS) has commissioned a report on dilbit to be released shortly after this Report;
the information in the NAS document may begin to address some of these gaps.

Recommendation: Each oil spill is unique due to the convergence of individual oil compositions
with specific suites of environmental parameters. Therefore, the environmental behaviour of
unconventional oils, blended heavy oils, bitumens and diluted bitumens needs to be investigated
under a wide range of relevant environmental and climatic conditions, including long-term, large-
volume outdoor trials. Online availability of such data would help inform first responders,
remediation personnel and modelers. To conduct such research requires sustained funding and
rapid access to funds set aside for investigating ‘spills of opportunity’.

2.4.5 Oil Interactions with Ice

Petroleum resource development in the Arctic raises the potential of oil spills in cold, pristine, ice-
impacted waters (Lee et al. 2011a). After having been given conditional regulatory approval in May 2015
and spending $7 billion to begin oil exploration in the Chukchi Sea off Alaska (Davenport 2015), Royal
Dutch Shell PLC halted exploration in October 2015 citing reasons of insufficient quantities of oil, low
crude prices, ‘high costs associated with the project, and the challenging and unpredictable federal
regulatory environment in offshore Alaska’ (Katakey and Zhu 2015). In response the US Department of
Interior cancelled the 2016 and 2017 auctions of natural gas and offshore oil leases (Kaufman 2015).
Even with this recent turn of events, the likelihood of increased shipping traffic through navigable areas
of the Northwest Passage brings greater risk for spillage of fuel oils, at the least. Furthermore, the concept
of exporting Alberta bitumen from Churchill, MB, via tanker ships (Welch 2014) highlights the need for
oil spill planning, preparedness and response in Arctic waters. Therefore, a brief summary of oil
interactions with ice is presented here as a special case affecting oil behaviour, in addition to site-specific

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 96
information presented in Chapter 3. Complete discussion of oil behaviour in/on ice and description of the
different types of ice on water are beyond the scope of this Report; the reader is directed to
comprehensive reviews (Lee et al. 2011a;
Considering that large-scale oil spills, such AOSRT 2014; NRC 2014) and a list of ice
EVOS in Alaska and DWH in the Gulf of definitions (Fingas 2015a). In addition, the
Mexico, held surprises even for professionals in longstanding, well-recognized Arctic and Marine
oil spill modeling, clean-up and response, the Oilspill Program (now the AMOP Technical
potential for a ‘Coldwater Horizon’ scenario in Seminar on Environmental Contamination and
the Arctic Ocean is a daunting prospect. Response; http://ec.gc.ca/amop/) organized and
sponsored by Environment Canada has published
a large body of peer-reviewed physical, chemical
and biological research, although much of it is not easily accessible. The reader is also directed to papers
arising from the 1980-1983 Baffin Island Oil Spill (BIOS) project, several of which are published in a
special issue of the journal Arctic (for an overview see Sergy and Blackall 1987). The BIOS project
focused on oil spill countermeasures for Arctic near-shore and shoreline oil spills, particularly the use of
chemical dispersants.

Certain weathering processes discussed in Sections 2.4.1–2.4.3 (e.g., spreading, evaporation, dispersion,
emulsification, biodegradation) are slower or may be eliminated seasonally at high latitudes under ice
(e.g., photooxidation, formation of tar balls), whereas others may be enhanced (e.g., sinking). Some of
these effects of ice are due to the increased viscosity and density of oil at low temperatures, to the lack of
sunlight to form oxidized crusts on tar balls, or to the lack of open water. Generally, oil evaporation on
water varies with the degree of ice coverage, wave height and air temperature (Brandvik and Faksness
2009). As ice forms, oil can remain on the ice surface (Figure 2.10), pool beneath it (Payne et al. 1990) or
migrate upward through brine channels (Petrich et al. 2013) and fissures in the ice to the floe surface. In
spring, if oil under the ice is warmed by sunlight, the lower albedo of the oil versus overlying snow and
ice may allow the oil to surface through the ice and float on melt pools. Heavy or emulsified oil may be
too viscous to migrate upwards and may remain pooled at the water–ice interface until disturbed by ice
movement or melting. The smaller oil surface area and low temperatures decrease dissolution and
dispersion into the underlying water and reduce evaporation into the atmosphere, thus retaining toxic
LMW hydrocarbons. Conversely, evaporation may be the major process affecting oil stranded on the ice
surface or absorbed into snow (Lee et al. 2011a). Oil trapped in pack ice will tend to move with the ice as
it is driven by wind and currents, but spreading of oil on top of ice and snow is retarded, creating thicker,
localized slicks. Spreading of oil under ice is much slower than on the water surface and the extent is
largely dictated by the roughness of the under-ice surface; the ice:water topography may slow oil
spreading (reviewed by Lee et al. 2011a). Emulsification may be reduced as ice dampens wind and wave
action and as the oil is confined in thicker slicks within smaller open water areas (Brandvik and Faksness
2009). However, heavy fuel oils may remain suspended in ice–water slush (AOSRT 2014), rapidly
becoming encapsulated in layers of ice during freeze-up (Dickins 2011) and isolating the oil from the
underlying water, further reducing weathering. Interaction of oil with suspended minerals and organic
particulates to form OPAs (Section 2.4.2.3) may still occur provided sufficient mixing, as observed in a
field trial in the St. Lawrence River (Lee et al. 2011b).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 97
Figure 2.10 Summary of interactions of oil with ice and snow. Additional detail can be found in AMAP (1998)
Figure 10-5 and accompanying text.

Cold marine waters are higher in dissolved oxygen and often also in dissolved nutrients, which should
promote biodegradation, but the rates are likely slower in the Arctic than in more temperate water because
enzyme processes and mass transport (discussed in Chapter 3) are slower at low temperatures.
Nonetheless, biodegradation of two North Sea oils by microbes from Arctic Ocean water and sea ice has
been demonstrated in the laboratory (Deppe et al. 2005), and versatile hydrocarbon-degrading bacteria
have been recovered from seawater exposed to an experimental oil spill in the North Sea (Chronopoulou
et al. 2015). Interestingly, diverse cold-adapted microbial communities, including species known to
degrade hydrocarbons, can be found in biofilms up to several centimetres thick at the bottom of sea ice in
contact with the water column along with pooled oil, where they may be able to biodegrade oil
components (Greer et al. 2015). Recently, biodegradation of chemically- and physically-dispersed ANS
oil was demonstrated at near-freezing temperature (-1 °C) by microbes in seawater from the Chukchi Sea,
in the absence of additional nutrients (McFarlin et al. 2014). Arctic lakes, however, may present
additional biodegradation constraints due to limited concentrations of nutrients and major ions.

The need for expanded testing of oil–ice interactions and oil behaviour in Arctic conditions merits special
attention. Specifically, significantly more research is needed to understand the behaviour and fate of oil
spilled into cold, deep seawater, including the rates and extents of physical, chemical and biological
processes. Most published observations of oil–ice interactions have been made with first-year ice, but
long-term studies of oil distribution and effects on multi-year ice are lacking. This is a time-sensitive
knowledge gap, as increased shipping and the possibility of offshore drilling in Arctic waters could
impact Canadian Arctic shorelines and seabeds. Likewise, future development of Canadian onshore
Arctic resources and pipeline transport will demand investigation of oil effects on cold freshwater rivers
and lakes and their porous shorelines.

Recommendation: Additional information is needed to describe and model the fates of petroleum in
and under sea ice and freshwater ice. Most of the limited research to date has been conducted on
first-year sea ice, but studies with multi-year ice are needed.

The facilities to conduct such oil–sea ice interactions may soon be available at the newly announced
Churchill Marine Observatory (Hudson Bay) to be constructed in Canada’s only Arctic deepwater port
(http://news.gc.ca/web/article-en.do?nid=996359). A purpose-built ‘Oil in Sea Ice Mesocosm’ (OSIM;
(http://news.umanitoba.ca/new-research-facility-to-open-in-churchill/) coupled with an Environmental
Observing system will enable investigation, detection and mitigation of oil spills in Arctic sea ice.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 98
2.5 Summary of Identified Research Needs and Recommendations

2.5.1 Research Needs

2.5.1.1 Short Term (High Priority)

Conduct research on the behaviour of unconventional crude oils, particularly diluted bitumen
blends, spilled into marine and fresh water

The behaviour of unconventional oils and bitumen blends currently cannot be predicted with confidence.
This knowledge gap affects risk assessments, spill response planning and cleanup decisions. Accelerated
research is needed in the form of laboratory experiments and field trials to define the behaviour of diluted
bitumen and unconventional crude oils (e.g., shale oils) under various seasonal and site conditions
relevant to Canada.

Information needed about the effects of various weathering processes on such oils includes:

 Evaporation, particularly the bimodal behaviour of various bitumen blends;


 Emulsion formation, particularly of weathered diluted bitumen in fresh water;
 Susceptibility of dilbit spills to chemical dispersion;
 Sinking behaviour, including interactions with suspended particulates; and
 Susceptibility of fresh and weathered heavy oils and diluted bitumen products to biodegradation.

Combinations of environmental conditions, such as water temperature, salinity, mixing, aeration and
sediment load, must be considered under laboratory and mesocosm (e.g., wave tank) test facilities over
time scales of hours to days. Some trials are underway, but many relevant combinations of products and
conditions exist and should be explored until general rules of oil behaviour can be discerned.

Information generated from such studies must be made readily accessible to first responders, regulators,
remediation personnel, modelers and oil spill scientists.

2.5.1.2 Medium Term

Perform baseline physical, chemical and biological environmental surveys, including natural
microbial communities and natural environmental variability.

To discern the impact of an oil spill we must have an understanding of the natural level of variability of
sites under pre-spill conditions. Each oil spill is unique, not only to the specific oil product, but also to
environmental and climate factors. It is thus essential to collect baseline physical, chemical and biological
information for Canadian environments that are potential receptors of conventional or unconventional oil
spills, particularly for sensitive ecosystems.

Of the factors reducing residual oil concentrations in the environment following a spill, biodegradation is
the ultimate process for oil removal from the aquatic environment over the long-term. Therefore, the
baseline surveys should include the composition of natural microbial communities in Canadian aquatic
environments, particularly for sensitive ecosystems that may be impacted by spilled oil. Knowing the
presence (or paucity) of natural hydrocarbon-degrading microbes in an ecosystem would help selection of
cleanup strategies and predict the duration and success of cleanup efforts.

Information generated by the surveys must be made readily accessible to first responders, regulators,
remediation personnel, modelers and spill scientists to support monitoring needs, the selection of response
strategies and predictions of impact and recovery.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 99
Develop a ‘Biodegradation Index’ that correlates with chemical composition of oils.

Twenty years ago Environment Canada was commissioned by the Alaska Department of Environmental
Conservation to determine the susceptibility of nine crude oils and oil products commonly transported in
Alaska to biodegradation under cold, freshwater conditions (Blenkinsopp et al. 1996). This type of study
should be conducted using representative oils transported in Canada.

The study should include aerobic biodegradation potential under fresh- and saltwater conditions at cold
and moderate temperatures, and anaerobic biodegradation in the presence of sediments. If conducted
under optimum laboratory conditions, the results would delineate the maximum biodegradation to expect
under actual environmental conditions. Furthermore, data generated from the screen should be used to
construct a ‘Biodegradation Index’ correlating oil composition with susceptibility to biodegradation as a
reference tool for oil spill remediation.

Augment online catalogues of crude oil composition and properties

Existing online catalogues, such as Environment Canada’s Oil Properties Database and CrudeMonitor.ca,
are incomplete in scope and in compositional information. Expanding and augmenting such catalogues
would be useful for oil spill response planning and preparedness, particularly for diluted bitumen blends,
and for the safety of first responders and the public.

2.5.1.3 Long Term

Support for fundamental petroleum chemistry research.

Fundamental understanding of oil composition and properties is lacking in several areas. These
knowledge gaps can be addressed primarily in laboratory studies. Needs include:

 Development of validated and standardized methods to enable detailed chemical characterization


of poorly-characterized oil components, including alkyl PAH, asphaltenes, resins, naphthenic
acids and unresolved complex mixtures. Although this information may not have direct practical
importance for oil transportation per se, it would provide insight into the fate and behaviour of
spilled oil, including emulsification, photooxidation, biodegradation, toxicity, etc. This is a
particularly critical gap for heavy oils and bitumen blends that have large proportions of these
chemical classes, where such information should be used to refine oil spill models;
 Development of validated and standardized analytical methods and a framework to enable
comparison of conventional and unconventional crude oils. This would help determine whether
all transported oils simply reflect a continuum of chemical composition, or whether diluted
bitumens represent a novel oil class. A forthcoming report by the US NAS may address part of
the latter question, but the need for robust analytical methods remains; and
 Development of refined and standardized methods for measuring physical weathering processes,
including evaporation, dissolution and oil:particle interactions. Such information should be used
to improve predictive spill models and inform mass balance budgets post-spill.

Apply research results to address the longstanding remediation question ‘How clean is clean?’

Data obtained from laboratory studies, large-scale open-air experiments, long-term field research and
baseline data surveys should be compiled and interpreted with the intention of determining acceptable
endpoints of spill remediation based on chemical and biological indices. Long-term monitoring of follow-
on effects after experimental spills and spills of opportunity would inform regulatory guidelines for
cleanup endpoints.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 100
2.5.2 Operational Preparedness Needs

Arctic preparedness: There is urgent need for expanded documentation of oil–ice interactions and
oil behaviour in Arctic conditions.

This is a time-sensitive knowledge gap, given predictions of commercial tanker traffic through the
Northwest Passage, the potential for Arctic Ocean oil exploration and production, and increasing transport
fuel to Arctic communities. The facilities to conduct such oil-sea ice interactions may soon be available at
the newly announced Churchill Marine Observatory in Hudson Bay. It is also important to note that most
published observations of oil–ice interactions have been made with first-year ice, but long-term studies of
oil distribution and effects on multi-year ice are lacking.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 101
CHAPTER 3: EFFECT OF ENVIRONMENT ON THE FATE AND BEHAVIOUR OF OIL

Abstract

Following Chapter 2, which describes oil properties in terms of its chemistry and composition, this
chapter is divided into three main topical areas, namely: 1) key environmental factors affecting spilled oil
fate and behaviour; 2) oil impacts on different aquatic environments; and 3) importance of oil type in
relation to the environment. Under each main section, up to nine subsections discuss in detail how the
biological, chemical and physical characteristics of each environment bring about change to the oil and
the ramifications these changes have on the ecology of the biosphere. Section 3.1 outlines how
microorganisms degrade the different hydrocarbon classes in oil and how temperature, dissolved oxygen,
nutrient supply, salinity and pH affect and change the behaviour of the contaminating oil. Section 3.2
describes how oil impacts a wide diversity of environmental systems, such as freshwater ponds, lakes and
rivers; estuaries, including bordering wetlands; seawater, both surface and deep sea; and Arctic environs,
including permafrost, ice and snow. Section 3.3 concerns the three main types of oil (light, medium and
heavy) and how they differ in their behaviour in relation to the various environments.

Research recommendations based on the discussion of knowledge gaps were consolidated and prioritized
at the end of the chapter. In summary, research is needed to further our understanding of how the Arctic
environment affects oil spilled into that harsh but sensitive ecosystem and how the spilled oil affects the
biology and ecology of that environment. Oil interactions with permafrost and spring melt are important
unknowns in terms of access to sites for spill response, rapid spreading of oil during freshet, slow rates of
weathering, and how hydrocarbons interact with ice and suspended particulate matter. Research is also
needed to further our understanding of the effects of spilled oil on permafrost ecosystems and how best to
develop appropriate response strategies that mitigate the damage without causing further harm. In
addition, little has been done to advance our knowledge of how to deal with subsurface blowouts in ice-
covered marine environments. Given that Arctic drilling will likely ramp up in the future, improved
methods will be needed to detect and monitor the behaviour of spills on, in and under ice and within the
water column in the event of such a subsurface deep sea blowout. While the National Energy Board has
identified the need for additional precautionary measures (e.g., same season relief well capacity), further
research is needed to aid in development of oil detection and response strategies in these important
environments. Finally, although microorganisms represent the ultimate cleanup of oil-impacted
environments, microbial activity has been reported to be slower in temperatures near freezing. However,
this premise has recently been challenged, as relatively fast oil degradation rates were observed in the
5 °C deep sea of the Gulf of Mexico during the Deepwater Horizon (DWH) spill (Hazen et al. 2010).
Confirmation of this finding would greatly advance our knowledge of the role microorganisms might play
in Arctic spill cleanup in the future.

Introduction

Whereas Chapter 2 described the major physical and chemical properties that influence oil behaviour,
Chapter 3 discusses what happens to oil spilled in various aquatic environments. Types of waterbodies
that can be affected by oil spills include rivers, streams, mud flats, wetlands (including bogs, fens and
marshes), salt marshes, offshore (surface, water column and deep ocean), reservoirs, permafrost, muskeg,
ponds, lakes and groundwater.

With or without human intervention, spilled oil is affected by microbes naturally present in the
environment that are in turn influenced by prevailing physical and chemical conditions at the spill site.
Therefore, this chapter provides detailed discussion of key environmental factors that affect petroleum
biodegradation, including indigenous microbial species, temperature, dissolved oxygen concentration,
presence of nutrients, salinity (marine, estuarine and freshwater environments), pH and the type of
environment impacted, such as shoreline type (e.g., lacustrine, riverine, marine and estuarine) and

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 102
location (e.g., intertidal, supratidal, subtidal). Other factors include non-biological weathering processes,
oil type, amount spilled and components in oil causing short-term and long-term impacts. As discussed in
Chapter 2, each type of oil has distinct physical and chemical properties that will influence the way it will
spread and break down, the hazard it may pose to aquatic and human life, and the likelihood that it will
pose a threat to natural and man-made resources. All these factors are discussed below in the context of
aquatic environments.

3.1 Key Environmental Factors Affecting Spilled Oil Fate and Behaviour

Petroleum and its precursor materials have been present in the Earth’s biosphere for tens of millions of
years, and certain microbes have adapted to utilizing such high-energy compounds for growth. Even in
the present time, some algae synthesize alkanes (Lea-Smith et al. 2015), calanoid copepods synthesize
pristane from ingested chlorophyll (Short and Harris 1997), many plants synthesize waxes (Baker 1982),
and petroleum naturally seeps from the earth’s crust (such as in the Gulf of Mexico seabed, the Santa
Barbara, CA, seafloor, at the sea bottom off the
Hydrocarbons have been ubiquitous on this planet coasts of Baffin Island, NU, and the Queen
for millennia. Over this time, diverse microbes Charlotte Islands, BC, and in the Athabasca
have adapted to consuming hydrocarbons under River, AB, as it cuts through oil sands
myriad different environmental conditions, deposits). Thus, oil-degrading microbes are
typically cooperating as microbial communities to ubiquitous, and diverse microbes having the
degrade complex oils. ability to degrade hydrocarbons will always
occur in aquatic ecosystems, although in
pristine sites they comprise a small proportion
of the total microbial community. Oil incursion provides a competitive advantage for oil-degrading
microbes compared to those that cannot utilize hydrocarbons. Oil degraders will thrive while the
hydrocarbons are present and their proportions in a contaminated site will be much greater than in a
pristine site, while overall microbial biodiversity decreases during active degradation. As the degradable
components are depleted, the oil degraders will decrease in numbers and diversity will again increase.

Natural attenuation is the result of petroleum biodegradation by oil-degrading microbes naturally


present in the biosphere and is unassisted by humans. These natural rates of oil degradation may be
enhanced by bioremediation, which is a human intervention practice that accelerates the rate of oil
decomposition by microbes either through biostimulation (a process used to accelerate biodegradation by
supplying the indigenous oil-degrading communities with nutrients or other growth needs, as discussed
below, or bioaugmentation (addition of exogenous oil-degrading bacteria to augment the natural
populations to degrade the contaminating hydrocarbons). Bioremediation is normally used as a polishing
step after oil that can be removed by other means has first been removed, so that the microbes have only a
small mass of oil to degrade. These processes are discussed in greater detail in Chapter 6. Regardless of
intervention, microbial hydrocarbon degradation is still subject to physical and chemical conditions in the
impacted environment.

3.1.1 Presence of microbial species in oil-impacted water, sediments, shorelines and wetlands

Much progress has been made in recent years in our knowledge of hydrocarbon-degrading microbial
communities, especially in marine ecosystems. This research has identified a group of aerobic
microorganisms called obligate hydrocarbonoclastic (hydrocarbon-degrading) bacteria (OHCB), which
have been shown to play a key role in eliminating petroleum hydrocarbons contaminating marine waters
(natural seeps as well as accidental spills). According to Yakimov et al. (2007), when hydrocarbon
contamination occurs in marine ecosystems, successive blooms occur in indigenous bacterial genera
belonging to the OHCB group (Alcanivorax, Cycloclasticus, Marinobacter, Thallassolituus and
Oleispira, among others), which were present at undetectable levels prior to the spill. The most notable
genera are Alcanivorax, which contains several key enzymes that confer the ability to metabolize alkanes

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 103
for carbon and energy acquisition (Schneiker et al. 2006), and Cycloclasticus, which degrades polycyclic
aromatic hydrocarbons (PAHs) (Table 2.1). Microbes, such as Alcanivorax borkumensis, are exquisitely
adapted to biofilm formation (Figure 2.3), oil solubilization, degradation of iso-alkanes (Table 2.1) that
are considered recalcitrant, and tolerance of oil-induced stress (Schneiker et al. 2006; Sabirova et al.
2008). In addition to OHCB, many generalist (non-obligate) hydrocarbon-degraders, such as
Pseudoalteromonas, are also prevalent in marine water (e.g., in the North Sea; Chronopoulou et al. 2015).
OHCB have not yet been reported in marine sediments, shorelines and wetlands, although these
environments do harbour diverse generalist oil-degrading microbes. The difference in sediments and soils
may be due to anaerobic conditions versus aerated/aerobic conditions in surface water (Section 3.1.3),
resulting in the presence of different predominant microbes.

Some bacterial species are specialists in degrading saturated hydrocarbons and others in aromatics (Foght
et al. 1990). Thus, a suite of different microbes (a ‘community’) is required to achieve maximum oil
biodegradation. Some microbes possess complete biodegradation pathways and therefore can mineralize
hydrocarbons, oxidizing them completely to carbon dioxide (CO2) and water without requiring a partner
microbe. Mineralization is the preferred outcome for oil spill cleanup. However, some microbial species
have incomplete pathways and are only able to partially oxidize (‘transform’) the hydrocarbon, especially
if it is complex (e.g., an alkyl PAH). In the latter case, the transformation product from the first attack
may subsequently be mineralized by other species in a combined or concerted effort by a microbial
community, or it may resist further degradation and accumulate. These partially oxidized products may be
more or less toxic than the original substrate, depending on the chemical properties of the transformed
product. Furthermore, some bacteria have the ability to exchange genetic material (DNA) so that species
previously unable to degrade hydrocarbons may acquire that ability by obtaining some or all of the genes
encoding the requisite enzymes. Thus, the products of microbial oil biodegradation can vary depending on
the composition of the oil, the microbial community and gene transfer among species.

The enrichment of the aforementioned bacterial genera is affected by temperature, salinity, oxidation-
reduction potential and other physical/chemical factors. The rapid development of these classes of
microorganisms can be accelerated by the use of biostimulation procedures, such as the addition of
nutrients (for example, nitrogen and phosphorus) that frequently exist only in limiting quantities in nature
(discussed in greater detail in Chapter 6).

Many of the constituents of petroleum oils have been known to be biodegradable for decades.
Biodegradation of oil is one of the most important processes involved in weathering and the eventual
removal of petroleum from the environment, particularly for the non-volatile components of petroleum.
Numerous scientific review articles have covered various aspects of this process and the environmental
factors that influence the rate of biodegradation (Zobell 1946, 1973; Atlas 1981, 1984; NRC 1985; Leahy
and Colwell 1990; Johnsen et al. 2005; Head et al. 2006; Yakimov et al. 2007; Singh et al. 2012;
Gutierrez et al. 2013). Two comprehensive U.S. EPA reports on oil spill bioremediation covering all
aspects of microbial interactions with spilled oil and providing guidelines on how to implement
bioremediation in the field were published in 2001 and 2004 and are still relevant today (Zhu et al. 2001,
2004).

Microorganisms capable of degrading petroleum hydrocarbons and related compounds are ubiquitous in
marine, freshwater and soil habitats (Head et al. 2006; Yakimov et al. 2007; Atlas and Hazen 2011), and
degradation mediated by indigenous microbial communities is the ultimate fate of most of the
hydrocarbons that contaminate the environment (Leahy and Colwell 1990; Foght 2008; Prince 2010;
Atlas and Hazen 2011; Kostka et al. 2014). Hundreds of species of bacteria, yeasts and fungi have been
shown to degrade hydrocarbons ranging from methane to compounds of over 40 carbon atoms and in
temperatures spanning psychrophilic (-20 °C to +10 °C) to thermophilic (41°C – 122 °C) (Zobell 1973;
Floodgate 1984; Jordan and Payne 1980; Cooney 1984; Margesin et al. 2013).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 104
Despite their ubiquity in nature, hydrocarbon degrading microorganisms are distributed in relation to the
historical exposure of the environment to hydrocarbons. Those environments with recent or chronic oil
contamination, in particular the Gulf of Mexico with its myriad natural petroleum seeps, will have a
higher percentage of hydrocarbon degraders than unpolluted areas. Notably, no single strain of bacteria
has the metabolic capability to degrade all the components found in crude oil (Karrick 1977; Leahy and
Colwell 1990; Vinas et al. 2002). In nature, biodegradation of a crude oil typically involves a succession
of species within the consortia of microbes present. Microorganisms classified as non-hydrocarbon
utilizers may also play an important role in the eventual removal of petroleum from the environment.
Degradation of petroleum involves progressive or sequential reactions, in which certain organisms may
carry out the initial attack on the petroleum constituent, producing intermediate compounds that are
subsequently utilized by a different group of organisms, in the process that results in further degradation.

Petroleum is a complex mixture of many thousands of compounds consisting mostly of carbon and
hydrogen. As mentioned in Chapter 2, these compounds have been divided into four major groups,
abbreviated as SARA (saturates, aromatics, resins and asphaltenes). In general, the saturates fraction is
the most biodegradable with the exception of the fused complex alicyclic ring compounds, such as the
hopanes and steranes. The aromatic compounds, especially the PAHs, are of intermediate
biodegradability (which is inversely proportional to the number of fused rings), whereas the polar
fractions (i.e., resins and asphaltenes) have been long thought resistant to biological degradation (Pineda-
Flores et al. 2004; Naveenkumar et al. 2010).

With respect to saturates, the predominant mechanism of n-alkane degradation involves terminal
oxidation to the corresponding alcohol, aldehydes or fatty acid functional group. Branched alkanes are
less readily degraded in comparison to n-alkanes. Methyl branching increases the resistance to microbial
attack because fewer alkane degraders can overcome the blockage of beta-oxidation (NRC 1985). Highly
branched alkanes, such as pristane and phytane, which were earlier thought to be resistant to
biodegradation, have also been shown to be readily biodegradable, although at lower rates than normal
alkanes. Cycloalkanes, however, are more resistant to biodegradation. Complex alicyclic compounds,
such as hopanes and steranes (see Table 2.1), which belong to the class of pentacyclic triterpanes, are
among the most persistent compounds of petroleum spills in the environment (Wenger et al. 2002; Peters
et al. 2005). Because of their relative resistance to microbial attack, these compounds are used to
normalize the concentrations of the more biodegradable petroleum compounds as a chemical forensic
tool.

By normalizing biodegradable petroleum hydrocarbons to the relatively non-biodegradable hopane, the


variability of loss among alkane and aromatic compounds can be significantly mitigated, thereby
facilitating the calculation of rates of biodegradation of petroleum hydrocarbons in field studies.

Although the aromatics are generally more resistant to biodegradation than the alkanes, some low
molecular weight (LMW) aromatics, such as naphthalene, may actually be oxidized before many saturates
(Foght 2008). Monoaromatic hydrocarbons, such as benzene or toluene, are toxic to some
microorganisms due to their solvent action on cell membranes, but in low concentrations they are easily
biodegradable under aerobic conditions. PAHs with 2-4 rings are less acutely toxic and biodegradable at
rates that decrease with the level of complexity. PAHs with five or more rings can only be degraded
through cometabolism, in which microorganisms fortuitously transform non-growth substrates while
metabolizing simpler hydrocarbons or other primary substrates in the oil. Alkylated aromatics are
degraded less rapidly than their parent compounds; the more highly alkylated groups are degraded less
rapidly than less alkylated ones. The metabolic pathways for the biodegradation of aromatic compounds
have been the subject of extensive study (Atlas 1981; Cerniglia 1992; Prince 1993). The bacterial
degradation of aromatics normally involves the formation of a diol (two adjacent hydroxyl groups
attached to the aromatic ring), followed by ring cleavage and formation of a di-carboxylic acid. Fungi and
other eukaryotes normally oxidize aromatics using monooxygenases, forming a trans-diol.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 105
Compared to saturates and aromatics, little is known about biodegradation of resins and asphaltenes due
to their complex structures, making them difficult to analyze. Resins and asphaltenes have previously
been considered to be refractory to biodegradation. However, recent evidence suggests that asphaltene
degradation may occur through cometabolism (Leahy and Colwell 1990) or by other means (Liao et al.
2009).

Biodegradation typically affects the physical properties of oil by causing a decrease in the LMW alkane
and aromatic fractions and enrichment in the polar resins and asphaltene fractions (Liao et al. 2009). The
increase in the fraction of asphaltenes in heavy oil is attributed to the selective loss of alkanes and some
aromatic hydrocarbons. The asphaltene fraction is defined typically to contain those species that are
insoluble in low carbon number n-alkanes, such as n-hexane or n-heptane, while resins are soluble in such
solvents (Tissot and Welte 1984). In their study of biodegradation of asphaltenes, Liao et al. (2009)
concluded that the products of biodegradation, such as carboxylic acids, phenols and alcohols, may not
only contribute to the resin fraction of crude oils, but are also linked with functionalities of resins and
asphaltenes. The amount of asphaltenes increases because some resin molecules are enlarged and their
polarity increased such that they can be precipitated by hexane as newly generated asphaltenes.

One study (Uraizee et al. 1998) involved systematically altering the concentration of non-biodegradable
material in crude oil and analyzing its impact on transport of the biodegradable components of crude oil
to the microorganisms. They did this by first fractionating crude oil into its primary SARA classes, then
reconstituting the oil with increasing concentrations of the asphaltene fraction, and finally subjecting the
reconstituted oil to biodegradation in the laboratory as measured by respirometry. The authors developed
a mathematical model to explain and account for the dependence of biodegradation of crude oil through a
putative bioavailability parameter. Experimental results indicated that as the asphaltene concentration in
oil increased, the maximum oxygen uptake in respirometers decreased. The mathematically fitted
bioavailability parameter of degradable components of oil also decreased as the asphaltene concentration
increased. This was offered as one explanation for the lower biodegradation rates of heavy oils vis-à-vis
lighter oils.

Mass transfer is often overlooked as an important aspect of hydrocarbon biodegradation. If one assumes
that the aqueous solubility of PAHs limits their microbial degradation, mass transfer becomes important
in understanding and predicting the fate and rate of biodegradation in aqueous solution (Volkering et al.
1992, 1998; Bosma et al. 1997; Harms and Bosma 1997; Johnsen et al. 2005). Optimization of mass
transfer can be achieved in several ways:

 The production of biosurfactants by the degrading microbial communities, which facilitate


transmembrane diffusion of hydrocarbons into the cell by increasing the diffusion coefficient
governing the flux, and
 Exogenous addition of surfactants, such as dispersants, which tend to increase the interfacial area,
also resulting in a higher flux of substrate (PAH).

The rate at which microorganisms degrade a substrate at low concentration depends on their specific
affinity toward the substrate (Harms and Bosma 1997), defined as the ratio of the maximal substrate
uptake rate to the half saturation constant (Button 1985). High specific affinities lead to efficient low-
concentration PAH removal from the environment at the cell surface, steeper concentration gradients and
higher substrate transfer rates.

In addition to affecting diffusion as discussed above, surfactants also influence dissolution and
subsequent uptake of PAHs into the cell by attaching to the oil-water interface where micelles facilitate
PAH release from the interface (Volkering et al. 1998). Alkane degraders are known to produce
biosurfactants to increase the bioavailability of the poorly soluble hydrocarbons (Beal and Betts 2000;
Noordman and Janssen 2002). Such biosurfactants, especially those produced by Pheudomonal

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 106
aeruginosa, are commonly comprised of
rhamnolipids, which are characterized by a sugar
Hydrocarbon biodegradation can occur over a head group and a fatty acid tail. However,
wide range of temperatures, but rates generally biosurfactant production is not as common among
increase with increasing temperature to an the PAH degraders (Itoh and Suzuki 1972;
optimum for each environment. Oberbremer and Muller-Hurtig 1989; Johnsen et al.
2005), which means that biosurfactants may not be
as important for PAH metabolism as for alkanes. In
fact, because surfactants are known to interact directly with cell membranes, surfactants such as those in
dispersants may negatively affect uptake of PAHs into the cell and subsequent metabolism by disrupting
the integrity of the membrane, and it is possible this may even be exacerbated at deep sea hydrostatic
pressures (Campo et al. 2013).

Studies involving hydrocarbon-degrading communities, mostly conducted in laboratory cultures, have


been very helpful in elucidating the role of oil biodegradation in nature. However, laboratory studies can
also be misleading to our complete understanding of the behaviour and fate of oil when released into the
environment. A fuller understanding of the dynamics of in situ biodegradation affecting our ability to
predict the ultimate fate of hydrocarbons remains a challenge (Prosser et al. 2007).

3.1.2 Temperature

Temperature exerts major effects on the physical properties of oil, such as viscosity, surface tension and
density, that may influence the oil’s physical transport and environmental persistence, including
biodegradability, within the environment (Chapter 2). Temperature also influences the activity and
composition of the microbial community that responds to oil.

3.1.2.1 Physical-Chemical Properties of the Oil Influenced by Temperature

At low temperatures, the viscosity of the oil increases (Figure 2.2) and the volatility of toxic LMW
hydrocarbons declines, delaying the onset of biodegradation (Atlas 1981). This property may be
somewhat advantageous when a spill occurs in cold regions because it will allow response teams more
time to deploy treating equipment. Some hydrocarbons are more soluble at lower temperatures (e.g.,
short-chain alkanes) and some LMW aromatics are more soluble at higher temperatures (Polak and Lu
1973), altering susceptibility to biodegradation. Specific gravity of oil is dynamic in that it will increase
over time as the lighter components within the oil evaporate, dissolve or biodegrade, and it also responds
to temperature. Heavier oils, such as refined oils, dilbit and Orimulsion™ (bitumen mixed with water),
may sink and form tar balls at low temperatures or may interact with rocks or sediments on the bottom of
the water body.

3.1.2.2 Effect of Temperature on Biodegradation

Figure 3.1 shows that highest degradation rates generally occur in the range of 30–40 °C in soil
environments, 20–30 °C in some freshwater environments, and 15–20 °C in marine environments (Jordan
and Payne 1980; Bartha and Bossert 1984; Cooney 1984). These general increases with temperature
reflect the reaction rates of enzymes required for oil biodegradation, which generally double with a
temperature increase of 10 °C, to an optimum for each enzyme type. The variation among media reflects
the preferred growth temperatures of the different microbial communities that have adapted to their
environment. In addition to direct biochemical effects, increased temperatures can reduce a liquid’s
surface tension, making oil more likely to spread on warmer water than very cold water, thus affecting the
surface area of oil available for interaction with microbes.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 107
Figure 3.1 Optimum temperatures for hydrocarbon degradation rates in soil, freshwater and marine
environments. Adapted from Das and Chandran (2011).

Although biodegradation typically occurs faster and more completely at higher temperatures, in
environments where a cold-adapted microbial community has been established (e.g., one that prefers
temperatures < 15 °C, such as in perennially cold environments), degradation can also occur at significant
rates. Hydrocarbon biodegradation has been observed
at temperatures as low as 0-2 °C in seawater and -1.1
Temperature affects microbial communities °C in soil. Colwell et al. (1978) reported greater
and oil properties in complex ways, so the degradation of Metula crude oil at 3 °C than at 22 °C
rate and extent of biodegradation does not with a mixed culture in beach sand samples. Westlake
have a simple relationship with temperature. et al. (1974) found that bacteria capable of degradation
at 4 °C would metabolize oil at 30 °C, but those
populations that developed at 30 °C had a limited activity at 4 °C. Lee et al. (2011) collected samples
from the St. Lawrence River during a winter field trial study following the release of Heidrum crude oil
(25 °API; Chapter 2) in ice-infested waters, and used the samples to monitor time-series changes in oil
concentrations and composition in laboratory microcosms. Detailed chemical analysis (GC-MS with
hopane as the normalizing biomarker; Appendix A) from these studies showed that more than 60% of the
total petroleum hydrocarbons, 75–88% of total alkanes, and 55–65% of total PAHs were degraded after
56 days of incubation at 0.5 °C. McFarlin et al. (2014) recently reported on the biodegradation of Alaska
North Slope crude oil in Arctic seawater collected from the Chukchi Sea, AK, incubated at −1 °C.
Indigenous microorganisms degraded both fresh and weathered oils, in both the presence and absence of
the chemical oil dispersant Corexit 9500, with oil losses ranging from 46−61% and up to 11%
mineralization over 60 days.

Detailed oil biodegradation studies during the DWH spill with ‘genomics\ based techniques showed rapid
growth of petroleum–degrading microbes as the oil disappeared at faster than expected rates (half-life of
alkanes ranging from 1.2 to 6.1 days) under deepwater in situ conditions of 5 oC, nutrients and oxygen
(Hazen et al. 2010). A recent review by McGenity et al. (2012) presented strong evidence for oil
biodegradation in cold marine environments. Instead of Alcanivorax, which is a halophilic1 oil degrader
found mostly near the ocean’s surface and able to grow at temperatures from 4 to 35 °C, the alkane-
1
Halophiles are species that require a substantial concentration of dissolved salt to survive.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 108
degrading obligate psychrophile 2 Oleispira has been commonly associated with biodegradation of oil
spills in cold marine environments and in fact can constitute as much as 90% of the microbial community
near the oil spill. This genus has a wide global distribution and other genera are being isolated and
characterized, such as Oleibacter sp. (Teramoto et al., 2011) and microorganisms belonging to the order
Oceanospirillales (Golyshin et al. 2002; Hazen et al. 2010). Likewise, the bacterial genus Thalassolituus
is a marine OHCB that has been observed as the dominant member of an oil-degrading microbial
community in the Arctic seawater surface (Yakimov et al. 2007). Understanding presence, abundance,
activity and needs of cold-adapted hydrocarbon degraders is becoming important with the announcements
of offshore oil and gas exploration at high latitudes in regions such as the Chukchi and Beaufort Seas and
with anticipated increases in shipping in the Northwest Passage
(http://www.alaskapublic.org/2015/07/31/shells-exploratory-drilling-commences-in-the-chukchi-sea/).
However, more recent research casts doubt on future utility of transportation in the Northwest Passage as
findings by Haas and Howell (2015) show that even in today's warming climate, ice conditions must still
be considered severe (i.e., thick ice more than 100 m wide and thicker than 4 m has been observed to
occur frequently).

Recommendation: Confirmation of the finding that microbial activity was relatively fast in the 5 °C
deep sea of the Gulf of Mexico during the Deepwater Horizon spill (Hazen et al. 2010) would
greatly advance our knowledge of the role microorganisms might play in Arctic spill cleanup in the
future.

3.1.3 Dissolved Oxygen

Various microbes are able to grow with or without oxygen for their metabolism, and some can alternate
depending on the available oxygen in their environment. Oxygen has finite solubility in water (~7.5 mg/L
at 30 °C, increasing to ~15 mg/L at 0 °C for water in equilibrium with air at sea level) and, if oxygen is
depleted by microbial activity, it must be replenished to maintain aerobic metabolism. Replenishment is
readily accomplished in most shallow water columns by wind and wave action, so conditions of oxygen
limitation normally do not exist in the upper levels of the water column in marine and freshwater
environments and in the surface layer of most beach environments. Therefore, dissolved oxygen is
unlikely to limit biodegradation rates in these environments except under certain conditions, such as
strong stratification, although the physical state of the oil and the amount of available substrates also
affect oxygen availability.

Aerobic conditions are generally considered necessary for extensive degradation of oil hydrocarbons in
the environment since major degradative pathways for both saturates and aromatics involve oxygenases
(Atlas 1981; NRC 1985; Cerniglia 1992), and biodegradation of petroleum in the field has been shown to
occur substantially faster under aerobic than anaerobic conditions. Although aerobic degradation of crude
oil has been well-known for decades (e.g., Atlas 1981), mechanisms
are still being elucidated for both alkane and PAH metabolism (Rojo
The major pathways for 2009; Kanaly and Harayama 2010; Pérez-Pantoja et al. 2010; Fuchs et
removal of hydrocarbons al. 2011).
from the environment
involve aerobic conditions. There is little doubt about the importance of oxygen in determining
the rates and pathways of hydrocarbon biotransformation in
ecosystems that are dominated by anaerobic and/or anoxic conditions
including: subsurface sediments (marine, riverine and lacustrine); most fine-grained marine shorelines;
coastal and inland wetlands beneath the top several millimeters of soil; salt marshes; mud flats;
groundwater; muskeg; bogs and fens; anoxic zones of water columns; and any others where oxygen is

2
Psychrophiles are species that require constant low temperatures (usually < 10 °C) and die at higher temperatures.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 109
limiting. Bottom sediments are typically stratified, with dissolved oxygen and redox potential3 decreasing
sharply with depth, often to 0 mL L-1 and -300 mV or less, respectively—i.e., stringently anaerobic
conditions. Anoxia is primarily a result of the high affinity of aerobic micro-organisms for oxygen,
leading to its depletion in sediments where oxygen cannot diffuse quickly enough from bottom water into
pore water to replenish concentrations and maintain aerobic conditions.

Under anaerobic conditions the ultimate degradation of organic matter is generally much slower (Head
and Swannell 1999). Anaerobic degradation mechanisms involve different microbial species, degradation
enzymes and pathways, metabolites and end products than aerobic biodegradation and apparently also
impose a narrower substrate range (Schneiker et al. 2006;
Anaerobic hydrocarbon metabolism is Sabirova et al. 2006; Foght 2008). Furthermore, whereas
emerging as an important mechanism for a single species can usually accomplish complete aerobic
hydrocarbon biodegradation. hydrocarbon mineralization using well-studied pathways,
anaerobic hydrocarbon biodegradation has only recently
been recognized, many of the key microbes and their
pathways are still cryptic, some metabolites are unknown, and need for a community of microbes appears
to be more prevalent. Anaerobic oil degradation was deemed in early studies to occur only at negligible
rates, leading to the conclusion that the environmental importance of anaerobic hydrocarbon degradation
could be discounted in terms of remediation response. More recently, the biodegradation of some
aromatic hydrocarbons, such as BTEX compounds and a few PAHs (Foght 2008), n-alkanes (Mbadinga
et al. 2011) and iso-alkanes (Abu Laban et al. 2014), has been clearly demonstrated to occur under a
variety of anaerobic conditions. Studies have also demonstrated that in some marine sediments, PAHs and
alkanes can be degraded under sulfate-reducing conditions at similar rates to those under aerobic
conditions (Coates et al. 1997; Caldwell et al. 1998). Recently Gittel et al (2015) reported that anaerobic
alkane degraders are ubiquitously present in marine sediments, including sites that are isolated from deep
oil and gas formations and that are not exposed to high loads of anthropogenic hydrocarbon inputs.
Regardless of the large gaps in understanding anaerobic hydrocarbon metabolism, the occurrence of
anaerobic environments in marine ecosystems is sufficiently great that the anaerobic fate of crude oil
must be considered and the overall importance of anaerobic biodegradation of crude oil in the
environment requires further study.

It is important to note that all biodegradation of crude oil, whether anaerobic or aerobic, freshwater or
marine, aquatic or terrestrial, is selective in that some components are only partially degraded because of
their size or complexity or not modified at all (Wang et al. 1998). Thus, there will always be a residue of
original petroleum components, like asphaltenes and resins, partly degraded intermediates and other end
products of mineralization that may be generated (CO2, CH4, H2S, etc.), the mass of which will depend on
the interplay between environmental conditions, petroleum chemistry and microbial community
composition.

Recommendation: Further research is required on the mechanisms, pathways and rates of


anaerobic biodegradation of hydrocarbons, particularly with respect to circumstances where oil
has penetrated into sediment (riverine, lacustrine, estuarine or marine) where it could persist for
decades.

3.1.4 Nutrient supply and stoichiometric ratios of C:N:P

The incursion of crude oil into an environment represents a large influx of carbon that is normally
nutrient-poor. In particular, nitrogen (N) and phosphorus (P) are required for balanced growth of microbes

3
In environmental chemistry, redox potential is an electrochemical measurement or calculation to determine whether reducing or
oxidizing conditions prevail. Aerobic sites usually have positive redox values and anaerobic sites have progressively more
negative redox values.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 110
that biodegrade the oil and other microbial community members that comprise the base of the food chain.
In the 1930s, Alfred Redfield made thousands of measurements of the stoichiometric ratios of C:N:P in
oceanic plankton (Redfield 1934). These ratios were strikingly similar to the dissolved concentrations of
N and P found in the deep sea. This C:N:P relationship became known as the Redfield Ratio, which is
106:16:1, and has stood the test of time. Thus, in marine environments, nutrient limitation is generally
correlated to the low background levels of N and P in seawater. The C:N:P stoichiometry of
phytoplankton ultimately controls the nutrient ratios of the deep ocean (Arrigo 2005). These interactions
produce a self-regulating, biogeochemical system that maintains quasi-stable oceanic nutrient inventories
over both short and long timescales (Arrigo 2005). These stoichiometric ratios are important in
understanding the role microbes play in the event of a catastrophic oil spill when carbon is in great excess
to N and P. Once the oil is beached, the only way to accelerate biodegradation is to add limiting nutrients
in appropriate amounts to produce approximately the Redfield Ratio.

Freshwater systems (lakes and wetlands) range from oligotrophic4 to eutrophic5. Rivers may be nutrient-
poor at the source but generally become nutrient-rich downstream after receiving industrial and domestic
effluents and agricultural runoff (Cooney 1984). Freshwater wetlands are typically considered to be
nutrient limited, due to heavy demand for nutrients by plants. They are also viewed as being nutrient
traps, as a substantial amount of nutrients may be bound in biomass (Mitsch and Gosselink 1993). Both
freshwater lakes and wetlands may also exhibit seasonal variations in nutrient levels, which will affect the
performance of oil biodegradation. Ward and Brock (1976) found that the rate of hydrocarbon
biodegradation in an oil-contaminated lake was greatest during early spring when the nutrient content
was also high. As N and P levels declined in the summer (probably due to uptake by algal growth), oil
biodegradation also decreased.

Thus, although oil is rich in energy, carbon and hydrogen needed for microbial growth, it is poor in
essential nutrients, such as N and P. To balance the ‘diet’ of oil-degrading microbes, these nutrients can
be added in water-soluble form (such as conventional fertilizers) often, but not always, in ratios of C:N:P
approximating the 106:16:1 Redfield Ratio, such as the commonly used ratio of 100:10:1. Adding
nutrients is particularly important for beached oil, which no longer has access to dissolved N and P in the
water column. The only way to restore the C:N:P balance and accelerate biodegradation is to add the
elements that are limiting, i.e., N and P. This process, called ‘biostimulation’, is discussed in greater detail
in Chapter 6.

The largest use of biostimulation to aid in cleaning up a major oil spill was in Prince William Sound, AK,
during the 1989 Exxon Valdez incident (Atlas and Hazen 2011). Nutrients in various formulations were
used to accelerate the biodegradation of oil that contaminated the sand and gravel beaches. Field evidence
showed that biodegradation was indeed accelerated but, after a few years, both the treated and untreated
sites were equivalent in terms of residual oil remaining (Pritchard et al. 1992). It is generally true for
almost all countermeasure technologies that, eventually, nature catches up with accelerated recovery and
intervention does not necessarily result in greater total oil degradation, just faster rates. However,
biostimulation is not recommended for use in accelerating oil biodegradation on open water because
nutrients cannot be retained effectively at high enough concentrations in the water column due to dilution.
The same is true for bioaugmentation; dilution reduces the bioaugmenting culture rapidly. Use of
bioremediation for oil spill cleanup likely should be based on a Net Environmental Benefit Analysis
(NEBA) described in Chapter 4.

4
Oligotrophic environments have very low concentrations of carbon and nutrients suitable for sustaining life.
5
Eutrophic environments are the opposite of oligotrophic, having excess nutrients that can lead to unbalanced growth, such as
algal blooms, that have subsequent deleterious effects.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 111
3.1.5 Salinity

Although microbial degradation of oil has been observed over the salinity range observed in the natural
environment, changes in salinity may affect oil biodegradation by changing the composition of the
microbial population. For example, dramatic variations in salinity may occur in estuarine environments
where marine organisms mingle with freshwater species. Salinity is relevant not only to marine
ecosystems but also to fresh waters, shorelines and overland spills where brine may be a co-contaminant
with the oil.

Hydrocarbon biodegradation has been reported to be inhibited in non-saline environments and


groundwater when salinity increases (Margesin and Schinner 2001; Minai-Tehrani et al. 2009; Ulrich et
al. 2009). In contrast, most marine species have an optimum salinity range of 25 to 35 and grow poorly or
not at all at salinity lower than 15 to 20 (Zobell 1973). In a study of hypersaline salt evaporation ponds,
Ward and Brock (1978) showed that rates of hydrocarbon metabolism decreased with increasing salinity
in the range of 33 to 284. Whitehouse (1984) reported an inverse relationship between salinity and
solubility of PAHs. Thus, in general, a less saline environment may mean more soluble PAHs and better
biodegradation rates.

Microbial mats in an Arabian Gulf environment having chronic exposure to oil spills and highly variable
daily salinity and temperature fluctuations were studied by Abed et al. (2006). They reported almost
complete degradation of the PAH phenanthrene and the heterocycle dibenzothiophene at a salinity of 35,
and the best biodegradation of the iso-alkane pristane and n-octadecane between 35 and 80 salinity.
However, inhibition of hydrocarbon biodegradation by high salinity has been reported even in
environments where salt-tolerant or slightly halophilic microorganisms are dominant, such as mangroves
and intertidal microbial mats. As reviewed by Martins and Peixoto (2012), Diaz et al. (2002) assessed
hydrocarbon biodegradation in a mangrove microbial consortium immobilized onto polypropylene fibers.
They verified that alkane biodegradation increased from <40% at 0 salinity to 65% at >40 salinity and
remained there through 140 salinity, then declined back to <30% at 180 salinity, thus demonstrating an
optimum salt concentration for biodegradation by these hypersaline-adapted microbes, above which
biodegradation slows. In contrast, in studies with heavy crude oil-contaminated marine sediments, Minai-
Tehrani et al. (2009) reported that higher salt concentrations (sediment pore water having >10 salinity)
inhibited total crude oil and PAH biodegradation. Degradation of the PAHs phenanthrene, anthracene and
pyrene was greater at a salinity of 1, and fluoranthene and chrysene degradation was better at 0 salinity.
Based on the discussion above, it appears that one cannot delineate an optimum salinity range for
hydrocarbon biodegradation since such activity depends on the particular adapted microbial community
that is dominant in a given environment.

3.1.6 pH

Because seawater is slightly more buffered than fresh water, the pH of seawater is generally more stable
and slightly alkaline (Bossert and Bartha 1984), whereas the pH of freshwater and soil environments is
more variable. Organic soils in wetlands are often acidic,
as are some surface waters where the salinity is near zero,
Oil biodegradation occurs over a range such as Muskoka in Ontario where 92% of the lakes have
of pH values, but is generally optimum an alkalinity value 6 of less than 20 mg/l, which means
at near-neutral to slightly alkaline that they are susceptible to acidification through acid
conditions (pH 6.5 - 8). deposition (Muskoka Watershed Council 2007). The pH
of mineral soils is more neutral to alkaline. Most
heterotrophic bacteria (those that use organic material, including hydrocarbons, for growth) favour a
neutral pH (~6.5 – 7.5) (Leahy and Colwell 1990), while fungi are generally more tolerant of acidic

6
Alkalinity is a measure of the capacity of an aqueous solution to neutralize acid; it is not the same as pH.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 112
conditions (pH <6.5). Therefore, bacteria are the dominant hydrocarbon-degrading microbes in many
circumneutral environments, such as aquatic ecosystems, but fungal biodegradation may increase in
importance in acid-stressed environments.

3.2 Oil Impacts on Different Aquatic Environments

3.2.1 Sand and gravel shorelines

The behaviour of spilled oil on sand and gravel shoreline environments primarily depends on the
properties of the shoreline, such as the porosity of the substrate, the morphology of the shoreline (slope,
topographical variability, sinuosity, stability) and the energy of the
waves impinging on the shoreline. Differences will also occur in the Persistence of nutrients on
fate of oil along marine, estuarine and freshwater shorelines, largely marine beaches is dependent
due to differences in biological diversity and biomass, as well as on tidal and wave action.
dissolved oxygen and nutrient (N and P) replenishment, as discussed
above.

This interaction of oil and fine particles reduces the adhesion of


oil to intertidal shoreline substrates through the formation of oil-
Interactions between oil and fine mineral aggregates (OMAs; Chapter 2) that are easily dispersed
mineral particles play an by tidal action and currents. More importantly, suspended OMAs
important role in natural oil (also, described as clay-oil flocculation) enhance the availability
cleansing in marine shorelines. of oil for biodegradation (Figure 2.3), and thus oil biodegradation
rates may be accelerated by this process (Bragg and Owens
1995; Lee et al. 1997a, b).

Definitive research conducted on two beaches in Scarborough, ME (Wrenn et al. 1997a, b) clarified the
role of wave and tidal energetics and geomorphology on oil biodegradation. They pointed out that
beaches with strong wave action will tend to more rapidly lose nutrients needed to maintain or enhance
oil biodegradation and that salinity layers are created between the fresh groundwater and the tidal
seawater in both high and low wave-action beaches. These tracer studies showed that the washout rate of
nutrients from the oil-impacted intertidal zone is strongly affected by the wave activity of the
contaminated beach. Wave action in the upper intertidal zone causes nutrients from the surface layers of
the beach to be diluted directly into the water column, resulting in their immediate loss from the
bioremediation zone. Washout due to tidal activity alone, however, is relatively slow, and nutrients will
likely remain in contact with oiled beach material long enough to allow effective biostimulation on low-
energy beaches. In the two beaches studied (a high and a low energy beach), the ‘sandwiching
phenomenon’ (a layer of lower salinity water located between two higher salinity water layers) was
observed, which was confirmed by Boufadel et al. (1999) in a mesocosm study. Both the mesocosm and
field studies are helpful in providing guidelines for the application of bioremediation strategies, such as
biostimulation based on the optimal application of nutrients on marine beaches, which are discussed later
in Chapter 6.

Recommendation: The interaction between residual oil in the water with suspended particulate
matter of organic origin (e.g., phytoplankton, detritus, extracellular released products) and
inorganic origin (e.g., mineral fines) (OMA and clay flocs) on the environmental persistence of
spilled oil (sinking, oil biodegradation rates, etc) requires further study. This is a missing point on
the calculation of the oil’s mass balance following a spill.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 113
3.2.1.1 Porosity of the substrate

Higher wave exposure enhances both physical


The rate and depth of oil penetration depend removal and weathering processes. Wave-swept
primarily on the porosity of the substrate. rocky shores tend to recover from oil spills within a
matter of months, whereas marshes may act as a
petroleum sink for many years. Tidal pumping
promotes oil penetration into the sediments. Carls et al. (2003) showed that tidal pumping causes
dissolved hydrocarbons in beach pore waters to move laterally into streams cut through beach sediments.
On coarse-grained shorelines like cobble and sandy beaches, oil can penetrate more deeply and remain
longer (when it is trapped below the limit of wave action; Figure 2.3), compared to finer grained
sediments such as silt and clay. However, oil is more easily removed from coarse-grained sediments by
water flushing. Interactions of oil with tidal action, waves and shoreline substrate may also form asphalt-
like oil-sediment mats that are resistant to further biological and photochemical weathering. Thus,
porosity may control oxygen availability and hence control natural biodegradation rates. More detailed
discussion of oil interacting with armoured beaches is presented in Chapter 2.

3.2.1.2 Shoreline morphology

Steep shorelines (particularly if exposed to wind and


Shoreline morphology influences oil fate wave action) will experience significant erosional
via effects on the nature and extent of processes. Therefore, oil deposition would tend to
transport of the oil both longitudinally develop in a narrow band within the zone of wave action
along the shoreline and inland. where it would be subject to repeated cycles of re-
suspension and deposition. Steep shorelines usually have
little or no vegetative cover, thus further decreasing the
likelihood that oil will remain onshore for substantial lengths of time. Shorelines with low sinuosity (i.e.,
running in a straight line) will tend to have sandbar formation offshore, producing the potential for oil to
be deposited on the sandbars as well as along the shore itself. Shorelines with higher sinuosity (curving
with embayments) will produce lower-energy conditions on lee-sides of the curvatures, where oil
deposition would be more likely to occur. High topographical variability along a shoreline (e.g. rocky
outcrops, stream mouths, dunes and human infrastructure, such as docks, weirs and break-fronts) will
create variable exposure to wind and wave action, creating, in turn, variable oil deposition. Unstable
shorelines subject to erosion will experience changes in position and composition annually, thus affecting
the rate at which deposited oil is either re-suspended in the water column, pushed deeper and/or higher up
the shore, or transported in bulk via major bank failures.

3.2.1.3 Wave exposure

Higher wave exposure enhances both physical removal and weathering processes. Wave-swept sand and
gravel shores will often be unstable, as discussed above, resulting in a dynamic and somewhat
unpredictable sequence of deposition and suspension. The combination of wind and wave exposure with
low sinuosity and low stability would be the most likely to produce deposition/re-suspension cycles. The
ultimate fate of spilled oil under these conditions would depend upon the type of oil and rate of
weathering.

Rocky shores with high wave exposure tend to recover from oil spills within a matter of months,
particularly when the shoreline has a steep slope and vegetation cover is minimal (IPIECA 1995). In the
marine environment, more gradually-sloping rocky shores with heavy vegetation cover (e.g., Fucus, a
ubiquitous brown seaweed) and tide pools may experience some stranding of oil in the upper intertidal
zones. In freshwater lakes, rocky shores are often bare or limited to a lichen/moss vegetative cover;
therefore, oil would be less likely to strand (depending upon the density of the lichen/moss cover).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 114
3.2.2 Estuarine shorelines

Slopes are highly variable and range from gently


Estuarine shorelines are complex, dynamic and sloping (Figure 3.2) land to steep cliffs. The
diverse; therefore, the deposition and weathering waterline is not constant due to tides and storm
of spilled oil in these environments will be surges, creating a ‘shore zone’. The variability in
highly site-specific and difficult to predict. slope, energy climate, sediment composition and
other physical characteristics (including the
presence of woody debris) creates highly-variable
transport and weathering conditions for spilled oil. Rates of weathering will also depend upon the nature
of vegetative communities. Estuarine vegetation includes eelgrass beds, salt marshes and swamp forests.
Mud flats are common. Engineered structures are common along estuarine shorelines (e.g., bulkheads and
riprap revetments), adding to the complexity. Thus, there can be multiple stranding sites for spilled oil in
estuaries.

Figure 3.2 Pugwash River Estuary, Nova Scotia

3.2.3 Riverine shorelines

The behaviour of spilled oil on river shorelines will be affected by the stream order7, seasonal hydrology,
channel morphology, the tendency for hyporheic flow and the nature and extent of the riparian zone. A
spill into a lower-order stream (e.g., a small headwater stream) with a steep, rocky shoreline will behave
differently from a spill into a higher-order stream with a broad floodplain and lower slope. The timing of
the spill relative to seasonal flow will also be an important determinant. During the spring freshet, a spill
would be rapidly transported downstream and along the upper margins of the stream channel, where it
could become stranded once river flow drops. A spill that occurs during low-flow conditions (particularly
when the river is fully or partially ice-covered) may tend to remain in a smaller area; however, its ultimate
fate would depend on further transport and weathering processes once ice melts and flows increase.

7
Order is reflection of the size (and sometimes complexity) of a river or stream network, from small (brook or stream) to large
(major river)

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 115
Channel morphology will affect the deposition and weathering of spilled oil via the potential for stranding
on mid-channel bars or in side-channels. Higher-order streams will be more likely to have complex
channel morphology. Hyporheic flow (Figure 2.3) along stream margins (via surface-to-near-surface
groundwater connection) is common, and spilled oil could enter this flow system, then adsorb to particles
along the flow-path. Hyporheic flows also occur vertically across the channel, feeding water into the
hyporheic zone under the river. These downward flows may re-emerge immediately following an
obstruction or further downstream, creating transport pathways through sediments for dissolved and
particulate oil, trapping oil in sediments, and providing sources of dissolved oil to interstitial water of
streambeds and bottom waters where flows re-emerge. The width and vegetative cover of the riparian
zone along the river shoreline will also affect the transport and weathering of spilled oil. A spill that
occurs during high-flow conditions will likely enter the riparian zone and could become stranded on
vegetation or on the finer, organic substrates common to these zones. Riparian zones can contain oxbow
lakes and wetlands, which could become sinks for spilled oil.

Recommendation: More research is required on the persistence of oil in riverine, lacustrine and
mud flat environments. Such research would benefit risk assessments and aid response operations.
Research also needs to focus on mixing and physical dispersion in high gradient, high energy rivers
and hyporheic zones connected to groundwater resources.

3.2.4 Rocky cliffs and beaches

On coastlines where the land surface is at a relatively steep angle below the water table, the abrasive
action of continuous marine waves may create steep cliffs, and slopes constantly being eroded. As coastal
cliffs collapse, the shoreline recedes inland. The speed at which this happens depends, in particular, on the
strength of the surf, the height of the cliff, the frequency of storm surges, the hardness of the bedrock and,
in the Arctic, the importance of permafrost to bond shore sediments Rocky cliffs are typically not
considered a high ecological risk from oil spills unless they provide habitat to wildlife. These sites are
frequently left to self-clean by natural physical processes (e.g. scour by waves, etc.) as the oil is not
readily absorbed into rocks.

3.2.5 Lakes

Oil spilled into small lakes would likely rapidly spread to


The behaviour of spilled oil in lakes is cover a large proportion of the lake surface. Residual oil
affected by lake size, basin morphology, remaining after evaporative loss in small lakes would
exposure to prevailing winds, internal likely be deposited along the windward shoreline, but if
wind-driven lake currents (e.g., seiches) the lake is well protected with steep shores, some residual
and season of the spill. oil could remain in the narrow littoral zone immediately
offshore. Oil spilled into larger lakes would be pushed and
mixed by wind and wave action. The eventual location of
residual oil after the initial evaporative loss would depend upon the nature of the windward shoreline
(particularly slope and substrate size), the nature of the littoral zone immediately offshore (if there are
large littoral macrophyte 8 beds, the oil could be captured within those beds), and the effect of lake
currents. Seiches9 can push residual oil up the shoreline on the windward side, where it could become
stranded on vegetation or fine shoreline sediments. Oil spilled during seasonal lake turnover 10 in the
spring and fall could be subject to mixing deeper into the water column where dissolved oxygen
concentrations are low.

8
Macrophytes are aquatic plants that may be emergent, submergent or floating
9
Seiches are standing waves in an enclosed or partly-enclosed water body
10
Turnover is a seasonal phenomenon of water movement from top to bottom according to temperature (density), more common
in large, deep lakes than shallow lakes

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 116
As an example of oil spilled onto a lake, on August 3, 2005, a Canadian National Railway train derailed
near Wabamun Lake in central Alberta. The derailment released 712,500 L of Bunker C oil and 88,000 L
of pole-treating oil from the rail cars, across the shoreline and into the adjacent Wabamun Lake. Workers
removed much of the oil via vacuum trucks and by hand. Sorbent booms were ineffective in containing
the Bunker C oil. Some containment booms were deployed, but insufficient numbers were available.
Eventually, the size of the spill forced the engagement of a response organization with experience and
equipment suitable for a larger response effort. The cleanup took years. Although the heavily-used
recreational lake was re-opened a year after the spill, the Alberta government was still issuing health
warnings about tar balls (Figure 2.3) and oil sheen two years later (Goodman 2009).

To estimate the hydrocarbon-degrading capabilities of microbes naturally present in oil-contaminated


Lake Wabamun sediment, laboratory tests were conducted with spilled Bunker C oil recovered from the
lake and the reference oil, Alberta Sweet Mix Blend (ASMB), in microcosms at 22 °C for four weeks or
at 4 °C for eight weeks (Foght 2006; Wang et al. 2011). The results showed that Lake Wabamun sediment
microbes were capable of degrading a significant proportion of the ASMB reference oil at both incubation
temperatures. However, biodegradation of Bunker C was far less significant than ASMB, with smaller
masses of hydrocarbons being degraded under both incubation conditions. For example, only 126 mg
Total Petroleum Hydrocarbons per gram of Bunker C were degraded compared with 263 mg/g ASMB in
parallel ASMB cultures (fresh sediment, 22 °C for four weeks). Under the best incubation conditions,
biodegradation accounted for loss of only ~12 - 13% by weight of the Bunker C oil added to the cultures.
This shows that only a small proportion of LMW hydrocarbons is biodegradable with a high proportion of
components (including high molecular weight components, such as asphaltenes and resins) either not
readily biodegradable or not detectable by gas chromatography (Appendix A) in the spilled Bunker C oil.
This persistence is most likely due to the recalcitrance of the major constituents of Bunker C oil.

Oil spilled into the Great Lakes would be subject to high-energy phenomena similar to some marine
situations, but without tidal effects. Lake currents would be expected to play a major role in behaviour
and persistence. An offshore spill adjacent to strong alongshore lake currents could be widely dispersed in
a short period of time. Areas that are chronically contaminated with low levels of hydrocarbons (e.g.,
commercial ports and recreational marinas) might be expected to harbour larger numbers of microbes
adapted to oil degradation and therefore might be able to respond more quickly to an oil spill than
pristine areas where biodegradation could be delayed.

3.2.6 Wetlands

Wetlands, including marshes, swamps, fens, muskeg and mud flats, are sensitive to oil spills
because of their complex and rich ecosystem diversity and their ability to collect and filter runoff
from surrounding environments. Saline wetlands are common in some parts of Canada,
particularly the prairies, but the behaviour of spilled oil in these systems has not yet been
rigorously studied.

The physical and chemical characteristics of freshwater and saltwater wetlands that affect the dispersion
and weathering of oil include small areas of shallow water, finer sediments with high organic content,
high vegetation cover and high biochemical oxygen demand (leading to anaerobic conditions). Oil spilled
into these systems will not be widely dispersed by wind. Rather, it will tend to be stranded on fine
sediments or on vegetation. Transport out of the wetland would be via small stream discharge points.

Saline wetlands (Figure 3.3) are highly productive and are often ringed by broad salt flats with salt-
tolerant vegetation. They are important habitats for waterfowl. Biodegradation of residual oil spilled into
these systems would depend upon the presence and ability of salt-tolerant microbes (Section 3.1.5) to

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 117
degrade the oil. In addition, if the oil becomes entrained within the anaerobic sediments commonly found
within the wetland, the biodegradation rate of the hydrocarbons may be significantly reduced.

Figure 3.3 Aerial view of replicate plots of an experimental oil spill conducted by DFO-Canada and U.S. EPA in
2001 on a saltwater wetland in Nova Scotia, Canada.

3.2.7 Permafrost

Strategies are needed to deal with the likelihood of spills occurring in remote Arctic regions as
exploration and production activities increase. The U.S. EPA (2013) published a short report on the
impact of climate change on Alaskan permafrost11. Over the past 50 years, temperatures across Alaska
have increased by an average of 1.9 °C. Winter warming was even greater, rising by an average of 3.5 °C
and is projected to increase an additional 1.9 to 3.9 °C by the middle of this century. Thawing permafrost
associated with climate change may cause major impacts on transportation, forests, ecosystems and the
economy in the future. Over the past 50 years, thawing permafrost and increased evaporation have caused
a substantial decline in the area of Alaska's closed-basin lakes (lakes without stream inputs and outputs).
These surface waters and wetlands provide breeding habitat for millions of waterfowl and shorebirds that
winter in lower latitudes (Karl et al. 2009). Furthermore, as temperatures rise and permafrost thaws, the
softening soil can interfere with tree root systems (causing trees to sink into the ground) and infrastructure
supporting oil and gas pipelines, etc., possibly affecting pipeline integrity. With the exception of a few
recent reports (e.g., Yang et al. 2014), little is known about the impacts of oil on permafrost ecosystems
and related surface meltwaters, for example, loss of permafrost slope stability due to black-body heating
by solar radiation.

Recommendation: Research is needed to further our understanding of the effects of spilled oil on
permafrost ecosystems and how best to develop appropriate response strategies that mitigate
damage without doing further harm.

The ADAPT program (Arctic Development and Adaptation to Permafrost in Transition) was designated
to fund innovative Canadian research in the field of permafrost change. The ADAPT mission is to utilize
this research to develop a structure of “Integrated Permafrost System Science” to understand and predict
consequences of rapid changes on ecosystems and economics.

11
Permafrost is ground that remains frozen for two or more consecutive years.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 118
3.2.8 Ice-covered Arctic marine environments

The extent of sea ice is declining throughout the Arctic Ocean and adjacent seas. Some sea ice persists
from year to year (known as perennial sea ice), often getting thicker where it forms pressure ridges (e.g.,
the stamukhi zones between the 10 - 20 m isobaths). Over the past several decades, the perennial sea ice
has declined dramatically to be replaced by seasonal (first year) ice. Ocean currents and winds have also
played an important role, pushing perennial sea ice out of the Arctic basin. The average sea ice extent in
September has decreased by 11.5% per decade over the past 30 years.

Oil spills in these sensitive areas as they warm through global climate change could be devastating to the
changing ecosystems and exacerbate the effects of melting snow and ice.

A number of recent reviews have highlighted the results of landmark field studies conducted within the
Arctic by Canadian, Norwegian and American scientists on the environmental processes that affect oil
behaviour and weathering in open water and in ice, and the effectiveness of current and emerging spill
response countermeasures (Dickins 2011; NRC 2014).

In January 2012, members of the international oil and gas industry launched a major collaborative effort
to enhance Arctic oil spill capabilities under the auspices of the International Association of Oil and Gas
Producers (IOGP). This collaboration, called the Arctic Oil Spill Response Technology Joint Industry
Programme (JIP), is aimed at expanding industry knowledge and proficiencies in Arctic oil spill response
(http://www.arcticresponsetechnology.org/about-the-jip).The JIP is currently conducting a series of
research projects in laboratory conditions on six key areas of research: dispersants, environmental effects,
trajectory modeling, remote sensing, mechanical recovery and in situ burning.

Recommendation: In light of the Gulf of Mexico oil spill incident, a need has been identified to
advance knowledge of, and ability to deal with, subsurface blowouts in ice-covered marine
environments. While the National Energy Board has identified the need for additional
precautionary measures in these important environments (e.g., same-season relief well capacity),
further research is needed to aid in development of oil detection and response strategies.

Ecological exposure to oil from drilling causes adverse chronic effects on species composition and
diversity. Acute exposures from spills can be severely damaging to fisheries, marine mammals, waterfowl,
corals and coastal wildlife by affecting feeding, activity, avoidance, growth and reproduction (Semanov
et al. 1997). An important component of Arctic marine productivity is the under-ice community that
develops in response to sunlight penetrating sea ice (Chapter 4). This community would be highly
exposed to oil discharged from a subsurface blowout. Marine mammals would likewise be at high risk
given that they are forced to breathe at small holes in the ice where oil would concentrate (Chapter 4).
The primary sources of oil in the Arctic marine ecosystem are from drilling activities and spills during
transport. The risk of crude oil and/or fuel spills in the Arctic Ocean has increased with announcements in
2015 of imminent deep sea Arctic drilling in the Chukchi Sea between Alaska and Russia and the first
commercial cargo shipment navigating the Canadian Northwest Passage in 2013.

Oil spilled in the marginal ice zone and polynyas12 during high primary production periods could cause
extensive damage, especially in shallow waters important to organisms at all trophic levels in the food
chain. For example, during the summer season, runoff from the Mackenzie River forms a warm, low-
salinity ‘thermal bar’ along the coast of Beaufort Sea. Anadromous fish species congregate in these highly
productive waters to feed, grow and mature before returning to rivers in the autumn to spawn, and these
regions form crucial summer habitat for Beaufort beluga populations; these living resources and others
are depended on for food by native northerners. Onshore winds would carry surface oil into these

12
A polynya is an area of open water within the sea ice pack

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 119
shoreline zones, exposing resident species to hydrocarbons at important phases in their life cycles.
Existing response infrastructures are located at far distances, oil is less accessible when the sea is ice-
covered, and conditions may be equally challenging when open water provides conditions for large
waves, storm surges and coastal inundation. Topping this off is the slow rate of biodegradation at these
cold temperatures (Section 3.1.2), which could lead to persistence of the spilled hydrocarbons for
decades. For years many scientists have agreed that the Arctic environment will require special
precautions to minimize the risks of accidental oil spills (NRC 2014).

Uncertainty is high regarding the effect of oil in ice-covered environments. Oil tends to become trapped
between ice floes, under ice and within brine channels, resulting in a lengthening of the contamination
and exposure times (Figure 2.3). It is likely that recovery from a spill would require years for affected
littoral and benthic communities to return to health.

Drilling for hydrocarbons in the Arctic demands improved methods to detect and monitor the behaviour
of spills on, in and under ice and within the water column in the event of a subsurface blowout. In
addition, oil interactions with permafrost and spring melt are important unknowns in terms of access to
sites for spill response, rapid spreading of oil during freshet, slow rates of weathering and how
hydrocarbons interact with ice and suspended particulate matter.

In pack ice, oil accumulates at the water surface, flowing around any ice present. Oil can also collect
under the ice, which usually has rough topography associated with pressure ridges. At sea ice
concentrations less than 30%, oil behaves as in open water (Venkatesh et al. 1990), but at higher ice
concentrations, the oil drifts with the ice. The equilibrium oil thickness in slush or brash ice 13 is nearly
four-fold higher than on cold water. As a result, the oil-contaminated area at higher ice concentrations is
several orders of magnitude smaller, but the oil layer is thicker. Thus, the presence of pack ice
significantly slows oil spreading. This means that a smaller area is affected, but the impact on that area is
greater and may overwhelm the microbial community. Recent research (Greer et al. 2015) has discovered
the presence of oil-degrading microbial biofilms at the bottom of sea ice, but the prevalence and
persistence of such structures is unknown.

As it weathers, the fate of oil in ice is affected by density, viscosity, surface and interfacial tensions and
water content. As described in Chapter 2, typical weathering processes include evaporation, dissolution,
emulsification with water to form water-in-oil gels, photodegradation and dispersion. Singsaas et al.
(1994) found that the rates of evaporation, dispersion and emulsification were retarded in pack ice. They
reported that the primary factors that reduced the weathering rate when compared to open water and
temperate conditions were wave damping, limited spreading of oil bounded by sea ice and low
temperature.

Immediate surface evaporation results in losses of small alkanes (C5-C10) and monoaromatic compounds,
such as BTEX (benzenes, toluenes, ethylbenzene and xylenes). Several experiments with refined gasoline
and diesel (Serova 1992; Ivanov et al. 2005) have found nearly complete evaporation on the surface of ice
during the summer season. The highest evaporative loss occurs in the presence of uncompacted snow,
while increasing snow density and thickness reduces evaporation (Belore and Buist 1988).

To determine the importance of dissolution, Faksness and Brandvik (2008a, b) subjected six oils (five
crude oils and a heavy fuel oil) to freezing in sea ice during the winter and quantified the migration of
water-soluble components from the oil into the ice. They found that the content of water-soluble
compounds in the ice decreased with ice depth, and the concentrations changed over time, confirming that
the water-soluble components had been transported from the overlying oil through the brine channels in

13
Brash ice is an accumulation of the wreckage of other forms of ice made up of fragments not > 2 m across.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 120
the ice to the underlying water. The field experiments showed that both ice thickness and air temperature
prior to an oil spill are important for the distribution of water-soluble components in ice.

Moussification, or the formation of water-in-oil emulsions, forms by water-drop entrainment during


mixing (Figure 2.5). This is an important weathering process both in cold and warm water. Emulsification
increases the volume of the slick and reduces the biodegradation rate. The formation of stable water-in-oil
emulsions hampers various oil spill cleanup efforts by increasing volume for containment, lowering
combustibility for in situ burning, and increasing viscosity so that chemical dispersion is more difficult.

Photooxidation (Chapter 2) and biodegradation are important transformation processes that occur with
petroleum products released into the marine environment (Garrett et al. 1998; Dutta and Harayama 2000;
Prince et al. 2003). Whereas photooxidation is highly seasonal in the Arctic, biodegradation rates should
be relatively constant in Arctic sea water. Garret et al. (1998) demonstrated that photooxidation and
biodegradation operate on different aromatic hydrocarbon components in crude oils. Whereas
biodegradation and evaporation/dissolution result in the depletion of unsubstituted aromatic compounds
relative to their alkylated homologues, photooxidation selectively attacks the alkylated aromatic
compounds (Prince 1993; Prince and Clark 2004).

Finally, with respect to sedimentation (Figure 2.3), most oils do not sink as their specific gravity is less
than that of seawater. However, when suspended inorganic particulate matter interacts with oil, the
conglomerate formed may be negatively buoyant. This natural process may occur in near-shore
environments with a high suspended particle load such as that associated with glacial tills in the Arctic.
Oil sedimentation is usually considered to be disadvantageous for the overall removal of oil from the
environment because oil in the sediment phase tends to be limited by oxygen supply, and anaerobic
biodegradation of hydrocarbons is a slow, poorly understood process (Section 3.1.3), particularly when
combined with low temperatures.

3.2.9 Offshore Surface Spills at Sea and in Large Freshwater Lakes (e.g., Great Lakes).

3.2.9.1 Spreading

The spreading of oil on water is one of the most important processes during the first hours of a spill
provided that the oil pour point (the temperature at which it
At Arctic temperatures, the presence becomes semi-solid and loses its flow characteristics) is
of snow and ice can substantially lower than the ambient temperature. The principal forces
affect the spreading of oil. influencing the spreading of oil (Figure 2.3) include gravity,
inertia, friction, viscosity and surface tension. These
processes increase the overall surface area of the spill, thus
enhancing mass transfer via evaporation, dissolution and later biodegradation. Oil accumulated at the
water surface can be pushed by the wind, generally slightly to right of the wind vector in the northern
hemisphere. Wind, currents and tides, the three main contributors to transport of oil following an oil spill,
may be used to project where the oil might go and what it might encounter in its path.

3.2.9.2 Evaporation

Evaporation (Figure 2.3) is responsible for the removal of a large fraction of the oil including the more
acutely toxic, lower molecular weight components. For oil
on water, evaporation removes virtually all the normal
In terms of environmental impacts,
alkanes smaller than C15 within one to ten days. Gros et al.
evaporation is the most important
(2015) reported > 50% of ≤ C17 hydrocarbons disappearing
weathering process during the early
within 25 hours from an experimental 4,300 L North Sea oil
stages of an oil spill.
slick of < 10 km2 area and < 10 μm thickness. For oil sheen,

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 121
> 50% losses of ≤ C16 hydrocarbons were observed after 1 hour. Volatile aromatic compounds, such as
the monoaromatics benzene and toluene, can also be rapidly removed from an oil slick through
evaporation. However, these volatile oil components may be more persistent when oil is stranded in
sediments. The volatile components make up 20 - 50% of most crude oils, about 75% of No. 2 fuel oil,
and about 100% of gasoline and kerosene. As a result, the physical properties of the remaining slick
change significantly (e.g., increased density and viscosity). Major factors influencing the rate of
evaporation include composition and physical properties of the oil, wave action, wind velocity and water
temperature (Jordan and Payne 1980). Evaporation leaves behind the heavier components of the oil,
which may undergo further weathering or may sink to the ocean floor. For example, spills of lighter
refined petroleum-based products, such as kerosene and gasoline, contain a high proportion of flammable
light ends, which may evaporate completely within a few hours, thereby reducing the toxic effects to the
aqueous environment. Heavier oils leave a thicker, more viscous residue, which may have serious
physical and chemical impacts on the environment.

3.2.9.3 Dissolution

Although dissolution (Figure 2.3) is less important than evaporation from the viewpoint of mass loss
during an oil spill, dissolved hydrocarbon concentrations in water are particularly important due to their
potential influence on the success of bioremediation and the effect
of toxicity on biological systems. Light aromatic compounds,
The extent of dissolution such as benzene and toluene, are the most soluble in seawater,
depends on the composition and and the most acutely toxic. However, these compounds are also
state of the oil and occurs most first to be lost through evaporation, a process which is 10 - 100
quickly when the oil is finely times faster than dissolution, and oil contains only small amounts
dispersed in the water column. of them which tends to mitigate dissolution. Even so, the impact
of LMW aromatics on the environment is much greater than
simple mass balance considerations would imply (NRC 1985).
Dissolution rates are also influenced by photochemical and biological processes. The larger components
of oil (4-6 ring PAHs, alkanes > C20, etc.) are quite insoluble in aqueous solution, so they tend to persist
longer in the environment.

3.2.9.4 Emulsification

Emulsification (Figures 2.3 and 2.5), or formation of oil-in-water emulsions, involves the incorporation of
small droplets of oil into the water column, resulting in an increase in surface area of the oil.
Emulsification occurs by physical mixing promoted by
An emulsion is formed when two turbulence at the sea surface. The emulsion thus formed is
liquids combine, with one becoming usually very viscous and more persistent than the original oil
suspended in the other. and is often referred to as mousse because of its appearance.
The formation of these emulsions causes the volume of oil to
increase three- to four-fold, which slows and delays other
processes that would allow the oil to dissipate. Conventional oils with an asphaltene content greater than
0.5% tend to form stable emulsions that may persist for many months after the initial spill has occurred
(WSP Canada, Inc. 2014). However, as reported in Chapter 2, dilbits do not emulsify as easily as
conventional oils. Dilbits, if spilled into turbulent water, can form entrained or unstable emulsions, which
break down into water and oil within minutes or a few hours at most, once the sea energy dissipates
(Fingas 2015). Oils containing a lower percentage of asphaltenes are less likely to form emulsions and are
more likely to disperse.

Recommendation: Research is needed to further our understanding of the properties of dilbit that
preclude it from being less subject to emulsification than other crude oils.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 122
In general, oil-in-water emulsions are not stable, but water-in-oil emulsions are. Water-in-oil emulsions
containing 50 to 80% water are most common and have a reddish-brown colour and grease-like
consistency. Such emulsions are highly resistant to biodegradation. However, oil-in-water emulsions can
be maintained by continuous agitation, interaction with suspended particulates, and the addition of
chemical dispersants, discussed further in Chapter 6. Chemical dispersion may influence oil
biodegradation rates by increasing the contact area between oil and microorganisms or by increasing the
dissolution rates of the more soluble oil components. Emulsification is less likely to occur in fresh water,
even for spills of heavier oils, because significant mixing energy often does not persist long enough to
generate the stable emulsions seen at sea. Moreover, brine composition (alkalinity in particular because of
a buffering effect) is intimately tied to the pH in determining the stabilizing properties of the interfacial
films (Strasser 1968). Brines with high [Ca+2] and a high Ca+2/Mg+2 ratio form nonrelaxing, rigid films
around the water droplets, resulting in stable emulsions (Jones et al. 1978). Higher concentration of
divalent ions and high pH result in reduced emulsion stability. Where emulsions do occur, they may
separate into oil and water again if heated by solar radiation under calm conditions or when stranded on
shorelines.

3.2.9.5 Photooxidation

Oxidation is promoted by sunlight (Figure. 2.3) and the extent to which oxidation occurs depends on the
type of oil and the form in which it is exposed to sunlight.
In sunlight, oils react chemically with However, this process is very slow and, even in strong
oxygen either breaking down into sunlight, thin films of oil break down at no more than 0.1%
soluble products or forming per day. The formation of tar balls is caused by the oxidation
persistent compounds called tars. of thick layers of high viscosity oils or emulsions. This
process, also called aggregation, forms an outer protective
coating containing heavy compounds, resulting in an
increased persistence of the oil in the environment. Tar balls, which are often found on shorelines and
have a solid outer crust surrounding a softer, less weathered interior, are a typical product of this process.
Oil aggregates may exist from a month to a year in enclosed seas and up to several years in the open
ocean. They complete their cycle by slowly degrading in the water column, on the shore if washed there
by currents or (if they lose buoyancy) on the sea bottom.

3.2.9.6 Sedimentation/sinking

Since seawater has a specific gravity of approximately 1.03, only heavy oils are dense enough, or weather
sufficiently, to result in sinking in the marine environment (Figure 2.3). Sinking usually occurs due to the
adhesion of sediment particles and/or dense organic matter to
the oil. Shallow waters are often laden with suspended solids
Some heavy refined products have a
(such as planktonic organisms or clay/silt particles)
specific gravity ≥ 1.0 and so will sink
providing favourable conditions for sedimentation. This is
in fresh or brackish water.
especially evident in estuaries, which are characterized by
having relatively high concentrations of suspended
particulate matter that aid in the sedimentation of oil droplets (NRC 2003). Valentine et al. (2014)
reported a 3,200 km2 region in the vicinity of the Macondo well contaminated with excess hopane
attributed to deposition from the DWH blowout. The pattern of deposition was consistent with both a
‘bathtub ring’ of oil impinging on the continental slope and a fallout plume of dirty marine snow where
suspended oil particles sank to the sediment. Another report from the DWH spill by Gutierrez et al.
(2013) implicated the role of exopolysaccharides from enriched strains of hydrocarbon degraders in
contributing to the formation of oil aggregates during the spill (Figure 2.3). Oil stranded on sandy
shorelines often becomes mixed with sand and other sediments. If this mixture is subsequently washed off
the beach back into the sea it may then sink. In addition, if the oil catches fire after it has been spilled, the
residues that sometimes form may be sufficiently dense to sink.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 123
3.3 Importance of Oil Type in Relation to the Environment

3.3.1 Light oils (diesel, heating oil, light crudes)

These types of oil (Section 2.3.2) are characterized as being moderately volatile and leave a residue (up to
one-third of spill amount) after a few days. They contain moderate concentrations of highly toxic
(soluble) compounds, especially distilled products. They may result in long-term contamination of
intertidal resources and have the potential for acute subtidal impacts (dissolution, mixing and sorption
onto suspended sediments). Cleanup can be very effective.

3.3.2 Medium crude oils (most crude oils)

These oils (Section 2.3.3) will evaporate moderately within 24 hours, have a maximum water soluble
fraction of 10 - 100 ppm, can cause severe and long-lasting contamination of intertidal areas, and can be
highly toxic to waterfowl and fur-bearing animals. They are chemically dispersible and can be cleaned up
effectively if response is fast.

3.3.3 Heavy oils

Heavy oils (Sections 2.3.7) include bitumen and dilbit and refined fuel oils such as No. 6 fuel oil and
Bunker C. They do not readily evaporate or dissolve in aqueous solution (water soluble fraction is < 10
ppm), can contaminate intertidal areas heavily with long-lasting effects, cause severe impacts to
waterfowl and fur-bearing mammals through coating and ingestion, and can sink readily into sediments,
exerting adverse effects on benthic communities. They weather slowly and are difficult to chemically
disperse. Shoreline cleanup is difficult under all conditions.

A major report was published in 2013 by the Canadian government that provided a comprehensive, in-
depth review of the properties, composition and marine spill behaviour, fate and transport of two diluted
bitumen products from the Canadian oil sands, namely, Access Western Blend (AWB) and Cold Lake
Blend (CLB). These two products were chosen to represent all dilbits in use in Canada because they were
the highest volume products transported by pipeline in Canada in 2012 and 2013 (Government of Canada
2013). The report documented results from laboratory and wave tank studies of the properties and
composition of fresh and weathered AWB and CLB. Included were measurements of the potential of the
two products to evaporate, photooxidize, and sink in saltwater, as well as how well the dilbits behaved
when treated with dispersants. The major results of the in-depth investigation are summarized below:

 Both dilbit products floated on seawater even after exposure to visible light and mixing with
water;
 The dilbits sank when they came in contact with fine particulate matter in seawater and mixed by
high energy wave action;
 When Corexit 9500 was used in an attempt to disperse the dilbit in seawater mixed by high
energy breaking waves, little to no dispersion occurred even when fine particulates were added to
aid the dispersion; and
 The two dilbit products did not evaporate as well as conventional crudes and fuel oils, but their
behaviours were similar in other ways, such as emulsion formation and how they interacted with
sediments.

It should be noted that dispersant testing in these wave tank studies was conducted at only one
temperature, 8.3 °C. At that temperature, dispersants are less effective against heavy fuel oils than at
higher temperatures (Srinivasan et al. 2007). Government of Canada (2013) did recognize this restriction
in its conclusions: “The physical properties (e.g., density, viscosity and adhesiveness) of these products
limit the effectiveness of currently-available spill treating agents, thus restricting remediation and

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 124
potentially contributing to the persistence of the products in marine environments where seawater
temperature is < 8 °C.”

Recommendation: More research is urgently needed to determine the highest-risk combination of


environmental characteristics when dealing with dilbit. Based on current knowledge and logic, for
example, under what circumstances is it most likely that a dilbit product would sink? Under what
conditions would weathering of dilbit be the slowest?

More research is needed to further our understanding of the behaviour of dilbit and other heavy
fuel oils in a range of environmental settings, including near-shore marine, offshore marine,
estuarine, freshwater lakes, rivers and wetlands, and environmental conditions, including
combinations of climate, water chemistry and biological communities. This new information would
help to optimize the selection of current spill response strategies and to support the development
and validation of emerging technologies, including biodegradation of mousse, water-in-oil and oil-
in-water emulsions. Details on the factors that influence moussification are needed for a better
understanding of its properties and its treatment. The same needs pertain to tar balls.

Research is needed to further our understanding on why dilbit is less subject to emulsification than
other crude oils. Knowledge gained might help in developing response strategies for all heavy oils.

3.4 Summary of Research Needs and Recommendations

3.4.1 High Priority Research Needs

3.4.1.1 Arctic Ecosystems

Research is needed to further our understanding of the effects of spilled oil on permafrost ecosystems,
where challenges include limited site access, rapid spreading of oil during freshet, slow rates of
weathering, and limited knowledge on the interaction of oil with ice and suspended particulate matter.
New research would show how best to develop appropriate response strategies that mitigate the damage
without causing further harm. In addition, little has been done to advance our knowledge of how to deal
with subsurface blowouts in ice-covered marine environments. Similarly, with the recent re-invigoration
of drilling for oil over the Arctic Ocean shelves and slopes, improved methods will be needed to detect
and monitor the behaviour of spills on, in and under ice and within the water column in the event of such
a subsurface deep sea blowout. While the National Energy Board has identified the need for additional
precautionary measures (e.g., same-season relief well capacity), a number of laboratory studies have been
initiated to further research, but field trials are needed to aid in development of oil detection and response
strategies in these important environments.

3.4.2 Medium Priority Research Needs

3.4.2.1 Persistence of Oil in Non-Marine Environments

Most literature on oil persistence concerns marine environments. More information on oil persistence in
riverine, lacustrine and mud flat environments, as well as sediments, would be beneficial for risk
assessments and response operations. Although much progress has been made over the last decade in our
knowledge of oil biodegradation, anaerobic biodegradation of hydrocarbons is less understood. This is
important to oil that has sunk into the sediment, whether it be riverine, lacustrine, estuarine or marine,
where it could persist for decades.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 125
3.4.2.2 High Risk Characteristics and Treatability of Dilbit

Although the Canadian government has already begun work on the treatability of dilbit, more research is
needed to rapidly determine the highest-risk combination of environmental characteristics when dealing
with dilbit. Based on current knowledge and logic, for example, under what circumstances is it most
likely that a dilbit product would sink? Under what conditions would weathering of dilbit be the slowest?
Related to that, the need exists to further our understanding of the behaviour of dilbit and other heavy fuel
oils in a range of environments, including near-shore marine, offshore marine, estuarine, freshwater lakes,
rivers and wetlands under various combinations of climate, water chemistry and biological community
conditions. New information is required for contingency plans to optimize the selection of current spill
response strategies and to support the development and validation of emerging technologies, including
biodegradation of mousse. Details on the factors that influence moussification are needed for a better
understanding of its properties and its treatment. The same needs pertain to tar balls.

3.4.3 Long Term Research Needs

3.4.3.1 Oil-Suspended Particulate Matter Aggregates and Dilbit Emulsification

The interaction between residual oil in the water with suspended particulate matter of organic origin (e.g.,
phytoplankton, detritus, extracellular released products) and inorganic origin (e.g., mineral fines) on the
environmental persistence of spilled oil (sinking, oil biodegradation rates, etc.) requires further study.
Formation of aggregates is presently neglected in the calculation of the oil mass balance following a spill.
Research is needed to further our understanding of the properties of dilbit that preclude it from being less
subject to emulsification than other crude oils. Knowledge gained might help in developing response
strategies for all heavy oils.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 126
CHAPTER 4: OIL TOXICITY AND ECOLOGICAL EFFECTS

Abstract

Chapter 4 reviews the literature on the toxicity of oil to marine and freshwater species to identify the
primary toxic effects of oil, the mechanisms of toxicity, factors that affect measured or predicted toxicity,
the ecological impacts of oil spills, and gaps in knowledge that obstruct the understanding of oil spill
impacts and how to manage them.

Spilled oil can be rapidly lethal to fish, birds, mammals and shoreline organisms due to the readily
dissolved components of oil and the physical effects of smothering and destruction of the thermal
insulation and buoyancy provided by fur and feathers. Chronic and sublethal effects are associated with
the less soluble components of oil such as the polycyclic aromatic hydrocarbons (PAHs), and some
effects may be expressed long after brief exposures. Biomagnification of hydrocarbons in food webs is
not an issue because fish, birds and mammals can quickly metabolize and excrete petroleum
hydrocarbons, although metabolism of PAHs often creates metabolites more toxic than the parent
compound.

Data on the toxicity of oil to aquatic species underpin ecological risk assessments (ERAs) of potential oil
spills and environmental impact assessments (EIAs) of actual spills. The literature on oil toxicity is rich,
but uneven. The most commonly reported data are from laboratory tests of oil toxicity to selected aquatic
species, particularly embryonic life stages that are most sensitive to oil. However, not all data are useful.
Many studies, particularly those with algae and plants, were designed to answer highly-focused questions
about specific oil spills, and it is often difficult to draw general conclusions that can be applied to new
circumstances.

ERAs and EIAs also rely heavily on modeling to predict the acute and chronic toxicity of hydrocarbon
mixtures from chemical properties such as lipid solubility (e.g., the Target Lipid Model or TLM). The
advantages of modeling are very compelling, considering the cost and time needed to run toxicity tests for
every type of oil that might possibly be spilled, or every water sample collected during an oil spill.
However, statistical error limits about model predictions are rarely quantified or reported in ERAs and
EIAs, despite uncertainty due to untested model assumptions about the stability of oil concentrations in
water, the transformations of hydrocarbons by photooxidation or metabolic oxygenation, single or
multiple mechanisms of hydrocarbon toxicity, the presence of oil in different phases, additive interactions
and the relationship between acute and chronic toxicity.

The toxicity of oil to aquatic species depends on the extent of exposure to the toxic components of oil. In
general, light oils rich in low molecular weight narcotic compounds, such as benzene, toluene, ethyl
benzene and xylene, are more acutely lethal than medium and heavy oils. Heavy fuel oils with high
proportions of 3- to 5-ringed alkyl PAHs are more chronically toxic than light oils due to the disruption of
embryonic development by alkyl PAHs. The concentrations of polar and heterocyclic compounds in
toxicity test solutions are rarely measured, so that potential effects unique to these compounds (e.g.,
endocrine disruption) are not routinely assessed. Recent spills of diluted bitumen (dilbit) have raised
concerns about toxicity because it is often described as a ‘dirty oil’. Nevertheless, only minor fish kills
have been caused by diluted bitumen (dilbit) spills, and a single report of dilbit toxicity demonstrated that
it causes the typical signs of chronic embryo toxicity at concentrations consistent with its alkyl PAH
content. However, there are too few data on toxicity to state with certainty that dilbit is more or less toxic
than conventional oils. Dilbit does differ from other oils in its behaviour when spilled, with a more rapid
loss of volatiles during weathering and a greater propensity to sink in fresh water. Thus, impacts of spills
may represent the unique interactions between the environmental behaviour of dilbit and the exposures of
species that select sediment habitats where dilbit accumulates.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 127
ERAs and EIAs may underestimate the potential impacts of oil spills because the relationships between
oil exposure and toxicity are not always clear. Analyses of oil solutions in lab experiments and field
surveys often combine dissolved oil and particulate oil in one measurement. However, the role of oil
droplets as an exposure pathway is ill-defined and there are significant analytical challenges in
discriminating droplet and dissolved oil fractions. Oil concentrations in water change rapidly and the
spatial and temporal characterization of hydrocarbon concentrations in toxicity tests and during oil spills
is often inadequate. In coastal and freshwater ecosystems, little is known about the interactive toxicity of
oil with chemicals from municipal, industrial and agricultural effluents. Likewise, the interactions
between oil toxicity and susceptibility to infection by pathogens are not well understood. As a
consequence, the impacts associated with an oil spill may be greater than expected.

Although oil spills on land and to fresh waters are frequent in Canada, too little is known about the fate
and effects of oil spilled to freshwater ecosystems. Most information about impacts is derived from
marine spills assuming that oil fate, behaviour and effects are similar in both systems. Although the scale
of oil spills from trains and pipelines is much smaller than spills from supertankers, the relative impacts
may be as great or greater. There is less scope for dilution and degradation of oil in freshwater ecosystems
so that higher concentrations of oil in water may be maintained for longer periods. Shorelines and
sediments of freshwater systems are always in close proximity to floating oil, which may accumulate in
thick layers and there is little time to react before spilled oil contaminates sensitive shorelines or aquatic
species. Freshwater species are less able to avoid spill sites and migrating or spawning species may
congregate at high densities where oil accumulates. The fate of much of the oil spilled to freshwater
ecosystems is often unknown, which raises questions about potential interactions of surface oil with
sediments and groundwater flows out of and into streams and lakes (hyporheic flows). Contamination of
sediments contributes to the exposure and effects on aquatic species that typically inhabit sediments,
particularly the early life stages of fish that spawn in sediments.

In both marine and freshwater systems, oil spill cleanup can be as damaging as the spilled oil. Habitat
damage is caused by the removal of oiled vegetation from the water or from riparian lands, clearing of
vegetation to permit access to oiled sites, soil compaction by heavy machinery, erosion of shorelines and
siltation and de-stabilization of streams by removal of oiled woody debris. Habitat damage can reduce the
abundance and productivity of species using the habitat and allow the invasion of species tolerant of the
new conditions. Shoreline erosion, once started, may propagate with high water levels or storm action so
that damage is ongoing and increasing in severity.

Chemical dispersion of marine oil slicks reduces the amount of oil contacting shorelines and surface
species, such as aquatic birds. Dispersion increases the rate at which oil is removed from the surface, is
diluted and is available for microbial degradation. However, the dispersants may increase the impact of
oil spills by increasing the exposure of subsurface species. In the top 10 m of water, concentrations of
dispersed oil may be increased to levels toxic to fish and invertebrate embryos. While the intent of oil
dispersion is to dilute the oil to concentrations below toxicity threshold limits, even brief exposures can
cause delayed effects that are evident in the weeks, months and years following a spill. The net
environmental benefits of dispersion represent the trade-offs between protecting easily-identified surface
resources and much-less-obvious subsurface resources and coastal/near-shore environments. Reports that
dispersants and oil cause synergistic toxicity are incorrect. Dispersion simply increases the concentration
of oil to which organisms are exposed. Dispersants themselves are moderately toxic to aquatic species
when free in solution, but appear to be unavailable and non-toxic when mixed with oil. Thus, dispersant
toxicity will depend on how accurately it is applied to oil, whether it is completely mixed with oil, and the
concentration of ‘free’ dispersant in surface water if applications by spraying from aircraft miss the target.
The environmental impacts of subsurface dispersant applications at the 2010 Gulf of Mexico Deepwater
Horizon blowout are still being evaluated in ongoing studies. In Canada, there are concerns about the
potential impacts of chemically-dispersed oil on east coast fisheries. Evidence–based net environmental

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 128
benefit analysis is essential to support decisions about appropriate oil spill response options, including
dispersants.

Although it is clear from laboratory studies that oil is highly toxic to aquatic species, oil spills are often
followed by years of debate on actual impacts. Cause-effect relationships are often obscured by a lack of
baseline data on affected species and by a high variability in population characteristics and productivity.
Variations in oil exposure and effects on a given population are associated with movements into and out
of affected areas, avoidance of oil exposure, fisheries closures, delayed effects of toxicity and density-
dependent changes in reproduction and survival. Cause-effect is also obscured when chemical,
biochemical or molecular indicators of oil exposure and effects in live or dead specimens are not included
in monitoring programs. Without these indicators, it is difficult to eliminate alternative causes of disease
and mortality. Post-spill monitoring is often perfunctory and ecosystem recovery is defined by how
quickly organisms re-colonize spill sites and not by how quickly ecosystem function and productivity are
restored. Overall, there is a lack of solid epidemiological sampling designs and a potential bias towards
‘false negatives’, i.e., concluding that oil spills caused no or mild effects on aquatic species, when in fact
important effects occurred but were not detected. Impact on human use of natural resources are also not
considered in a systematic way and there is growing interest in assessing oil spill impacts on ‘ecosystem
services’, i.e., “the benefits provided by ecosystems to humans that contribute to making human life both
possible and worth living”.

Oil exploration and development in northern and Arctic regions has raised concerns about the potential
for unique sensitivities of Arctic species to oil spills. Based on a limited array of toxicity comparisons,
Arctic fish and invertebrate species do not appear any more or less sensitive than more temperate species
when tested at equivalent temperatures. Thus, data from temperate species could be used for a first
approximation of risks of oil spills to Arctic ecosystems. However, the behaviour and fate of oil in Arctic
ecosystems can be strongly affected by low temperatures, ice cover and the interactions of Arctic species
with ice. For example, marine mammals are forced to breathe at small gaps in ice sheets, where floating
oil may accumulate. Differences in oil spill impacts between Arctic and temperate regions may be driven
more by factors that affect the exposure of organisms to oil than by differences in sensitivity to oil.

The capacity to identify, measure and manage a wide array of potential effects of oil spill is often limited
by a paucity of appropriate tools and data. This chapter provides a series of specific research needs
throughout the text, and these needs are summarized as follows.

Research is needed:

 At ‘spills of opportunity’ and experimental oil spills to identify effects on aquatic species at the level
of populations, communities and ecosystems due to the acute and long-term toxicity of spilled oil;
 On how the behaviour and fate of fresh and weathered oil in different ecosystems interacts with the
habitat selection and unique life history traits of aquatic species to control species-specific oil
exposure and toxicity. These interactions are particularly important for diluted bitumen (dilbit)
relative to conventional oils, and northern and Arctic ecosystems relative to temperate ecosystems;
 To understand how different oil types (e.g., dilbit) and spill-control agents (e.g., chemical dispersants)
affect aquatic species under different environmental conditions, including interactions with
contaminants from municipal, industrial and agricultural effluents;
 On mechanically- and chemically-dispersed oil to determine how oil droplets affect the exposure and
toxicity of oil to different species and life stages of aquatic organisms. These studies include a need to
develop analytical methods to reliably discriminate droplet oil from dissolved oil. Studies on the
distribution and effects of dispersed oil under exposure conditions similar to those that may be
encountered during response operations (including subsurface injection of dispersants) are of
particular importance;

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 129
 On strategies and protocols to recover spilled oil that avoid or minimize habitat damage, methods to
restore damaged aquatic and riparian habitats and models to define the balance between the
environmental costs of natural remediation and oil spill cleanup;
 On methods for finding and measuring spilled oil in different compartments of freshwater and marine
ecosystems, for measuring exposure and toxicity of oil to aquatic species in situ, and for assessing the
structure and function of ecosystems that reflect the extent and time to recovery following oil spills;
 On the contamination of groundwater due to soil contamination and the entrainment of oil droplets
into bed sediments of rivers by surface water - groundwater exchanges;
 To develop standardized oil toxicity test methods for all classes of aquatic organisms to enable
reliable comparisons of toxicity among oils, species and different environmental conditions (e.g.,
salinity, temperature), including rapid changes in oil concentrations in water; and
 To refine models that predict acute and chronic toxicity to aquatic species, particularly to assess the
uncertainty caused by violating the assumptions underlying these models and to reduce the variance
of model predictions.

Introduction

A primary concern of oil spill responders is to understand and limit the effects of spilled oil on
ecosystems, natural resources and the economies and communities that depend on them. These concerns
are the main drivers of prospective or retrospective ecological risk assessments (ERAs) or environmental
impact assessments (EIAs) which are conducted to understand and mitigate the impacts of future
developments (e.g., oil pipelines) and to investigate the effect of oil spills or chronic oiling from urban or
industrial sources. Knowledge of the ecological impacts of spilled oil is derived from observational
studies following actual spills and experimental laboratory studies, primarily toxicity testing with
individual species. While laboratory studies of oil toxicity to species of invertebrates, fish and birds are
abundant and detailed, research on oil spill effects on marine mammals, populations of organisms and
ecosystem structure and function is often sparse or non-existent. The unevenness of understanding is due
to differences in the questions being asked and the logistical difficulties and costs of conducting larger-
scale research, including outdoor mesocosm experiments and full-scale field experiments.

The nature of concerns regarding oil spills is also changing. Past perceptions of oil spills were based on
high-profile shipping disasters, where large volumes of oil were spilled with widespread acute effects on
wildlife and shoreline oiling. Perspectives have changed due to the extensive research following the 1989
Exxon Valdez oil spill (EVOS) in Prince William Sound, AK. There is now a greater awareness of the
potential persistence of oil, the ongoing exposure of aquatic species due to their behaviour and life habits,
and the delayed and cumulative effects of exposure to oil and to other stressors. Concerns are growing for
the potential effects of spills in extreme environments as exploration and development expand into Arctic
ecosystems and into very deep offshore waters. Similar concerns relate to potential impacts during
extreme seasonal conditions (winter, flooding, droughts, etc.) and impacts in remote wilderness areas that
are not readily accessible for spill response. The expansion of mid-continent oil production in North
America has highlighted the increasing frequency of freshwater oil spills and the question of whether
knowledge of marine spills can be applied directly to managing spills to fresh waters. Recent spills of
ultra-light oil, heavy fuel oils (HFOs) and diluted bitumen raise questions about the fate, behaviour and
impacts of oil with extreme properties and whether the understanding of conventional oils can be applied
successfully to predicting impacts of these oils when spilled.

This chapter provides an overview of what is known about oil toxicity and ecological impacts of oil spills
to aquatic environments. The discussion of terrestrial effects is limited to the water - land interface, i.e., to
susceptible shorelines, including salt marshes or wetlands where spilled oil is blown ashore, and riparian
lands of rivers and lakes where spilled oil may be deposited by flooding. For shorelines and riparian
lands, there are also concerns about potential habitat destruction by the oil spill response itself. This

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 130
chapter is based on several comprehensive reviews (Engelhardt 1983; Leighton 1993; NRC 2005; Hodson
et al. 2011; Lewis and Pryor 2013; Beyer et al. 2015; Logan et al. 2015) to avoid repeating summaries
that have already been published. Information from these reviews is presented in a synthesized form but,
where appropriate, some original papers are explained in detail to illustrate a point. Where still valid,
research recommendations from these reports are repeated. Data from primary publications are
summarized in Appendix D, Table D.1, which was first assembled by Hodson et al. (2011), enlarged by
Logan et al. (2015), and enlarged and re-formatted for this review. The numbers of citations to original
papers reflects the upsurge of interest following the 2010 Deepwater Horizon (DWH) oil spill, and this
report is heavily weighted towards recent studies with fish, the most numerous in the literature.

4.1 Effects of Oil on Aquatic Organisms

The effects of oil on aquatic species start with exposure to oil or its components. Exposure may be
derived from:
Biomagnification – a food chain or food
 External coatings of oil, particularly in air-
web phenomenon whereby a substance or
breathing species that must surface frequently
element increases in concentration at
and cannot avoid contact with surface slicks of
successive trophic levels. Biomagnification
oil;
occurs when a substance is persistent and is
 Inhaled aerosols of particulate oil and volatile
accumulated from the diet faster than it is
hydrocarbons by air-breathing aquatic turtles,
lost due to excretion or metabolism.
birds and mammals that contact surface oil
(Engelhardt 1983);
 Oil ingested by aquatic birds and mammals that groom after being contaminated, that forage on
beaches or intertidal zones contaminated by oil, or that ingest oil associated with algae or with
organic or inorganic particulates (Engelhardt 1983; NRC 2005). Although invertebrates do not
metabolize or excrete petroleum hydrocarbons quickly, and can contribute to the dietary exposure
of predators, petroleum hydrocarbons do not typically biomagnify in food webs. Unlike persistent
chlorinated compounds, most hydrocarbons can be readily metabolized and excreted by fish,
birds and mammals. However, rates of uptake, metabolism and excretion will vary among species
and with environmental factors, such as temperature; and
 The dissolved components of oil that partition from water to lipids across respiratory membranes
of aquatic species that breathe via gills (invertebrates, fish), and across cell membranes of
microbes and algae.

The toxicity of oil to each species varies with exposure, and exposure is a function of oil type,
environmental conditions and the life history and physiology of each species. Details about exposure and
factors controlling the rates of exposure for different species will be discussed in the reviews of oil
toxicity that follow.

4.1.1 Immediate effects of an oil spill

The first and most obvious effect of an oil spill is the contact of oil with shorelines and organisms that
inhabit the water’s surface, including birds, marine mammals and reptiles. Birds are highly vulnerable
because they spend most of their time on water (diving birds) or shorelines (wading birds). Birds and fur-
bearing animals, such as seals, sea otters and sea lions, are immediately affected because oil sticks so
readily to feathers and fur. When waterfowl and fur-bearing animals come in contact with oil they usually
succumb by loss of buoyancy and hypothermia because oil destroys the insulating properties of plumage
and fur. Smooth-skinned mammals, such as dolphins and whales, are generally less affected by direct
contact with oil. Ingestion of oil, often from grooming to remove oil from fur or feathers, causes anemia,
pneumonia, intestinal irritation, kidney damage, altered blood chemistry, decreased growth and decreased
production and viability of eggs. Among reptiles, turtles are considered the most susceptible to surface

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 131
oiling, oiling of nests on beaches or direct ingestion of oiled prey. Possible effects of oil on turtles can
include egg and hatchling mortality, a reduction of hatchling size and weight and an increase in
respiration rate (RPI 1991).

Mortality due to hypothermia is not instantaneous. Lightly-oiled animals can sometimes recover by
grooming to remove small traces of oil, and many animals can be rescued and cleaned. Rates of recovery
will depend on: the extent of oiling; the capacity of different species to tolerate the stresses of capture,
cleaning, recovery in captivity and re-introduction to the wild; the dose of oil consumed by grooming
before rescue; and the size and technical capacity of the rescue effort (Clumpner 2015). Clumpner (2015)
identified a need for professional organizations to quickly rescue and treat affected animals.

The littoral zones of aquatic ecosystems extend from the high water mark of shorelines to relatively
shallow waters along shore, including the intertidal areas of marine ecosystems. The littoral zone is quite
sensitive to stranded oil because it serves as permanent and seasonal habitat for many aquatic species
uniquely adapted to periodic exposure to air (e.g., during tidal cycles), turbulence and a wide range of
temperatures. Oil may smother sessile species of invertebrates or plants that inhabit the littoral zone when
winds, tides or currents bring oil ashore. If the coating of oil is sufficiently thick, it prevents access to
sunlight, oxygen, soluble plant nutrients (e.g., N, P), planktonic organisms for filter feeders and removal
of waste by flushing with water. It will also impede the movements of organisms to feed, reproduce and
avoid predation. This is especially important in low energy environments, where layers of oil are not
removed by wave action. For salt marsh plants, oiling of the lower portion of plants and roots is more
damaging than coating of leaves and stems, especially if oiling occurs outside the growing season. More
damage is experienced if there is repeated contamination of sediments in areas where the oil may persist.
Shorebirds and wading birds can be affected by oil in the intertidal area. Land animals, such as raccoons,
that scavenge for food in intertidal areas and use them for shelter may ingest oil while eating exposed
prey and may become coated in oil while exploring exposed flats and grass beds. The challenge for
shoreline cleanup is to remove the oil without killing the organisms that are coated. In most cases, this is
impossible, and often the best that can be done is to remove the oil quickly and completely to allow re-
colonization of the affected area by healthy organisms from adjacent un-oiled habitats.

4.1.2 Plants

Lewis and Pryor (2013) provided a comprehensive review of publications on the toxicity of 41 crude oils
and 56 dispersants to 85 species of unicellular and multicellular algae, 28 wetland plants, 13 mangroves
and nine seagrasses. However, the authors concluded that the database was inadequate to support ERAs
of oil toxicity to algae and plants, primarily because the high diversity of methods and measurements
(e.g., 107 different response parameters) prevented generalizations. Effect concentrations varied by six
orders of magnitude (0.002 to 10,000 mg/L) and the most sensitive response was embryo fertilization of
Fucus. Few useful endpoints were derived for risk assessments because most studies were in response to
large marine spills and were experimentally diverse. For dispersants, most studies were of formulations
that are no longer used, although there have been a few studies of Corexit® EC9500A and Corexit®
EC9527A. For example, EC9527A was about 10-40 times less toxic than EC9500A to germination rates
of the brown macroalga, Phyllospora comosa.

Although oil toxicity to phytoplankton species has been reported frequently (e.g., Garr et al. 2014), it has
been difficult to determine if major oil spills cause irreversible damage. Effects on plankton are often
mitigated by the rapid dispersion and weathering of oil, recruitment of unexposed organisms from
adjacent areas due to currents and mixing, and rapid growth rates.

The primary interactions between oil and phytoplankton may be the physical adherence of oil onto their
surface (Lee et al. 1985) and the capacity of algae to absorb hydrocarbons and mediate oil transport
within the ecosystem as a result of filter feeding and/or sedimentation processes (Beyer et al. 2015).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 132
Shoreline and riparian plants are another matter. An experimental application of weathered medium-light
MESA (Medium South American crude oil) to a Nova Scotia salt marsh suppressed the growth of the
predominant plant species, Spartina alterniflora (Lee et al. 2003a). Although natural recovery occurred
within the same growth season, there was a secondary longer-term change in the plant community;
Salicornia, a more tolerant and opportunistic plant species, increased in relative abundance. Oil toxicity to
salt marsh plants also occurred following the DWH oil spill (Beyer et al. 2015). For some species, only
the emergent portion of the plant was affected, and vegetative cover was restored following cleanup.
However, where oil penetrated sediments, the roots died and marsh and coastal shorelines were subject to
accelerated erosion. This effect was halted only when plants recovered and re-invaded several years after
initial oiling. The role of chemical dispersants in facilitating the transport of oil into marsh sediments is
unknown (Lewis and Pryor 2013). Following the EVOS, aggressive cleaning of shorelines caused
significant impacts on abundance and reproduction of the marine macrophyte, Fucus sp., even though it
was relatively insensitive to oil (Van Tamelin and Stekoll 1996). The loss of this dominant plant caused a
major reduction in habitat and food for associated populations of coastal invertebrates and fish. Full
recovery of Fucus production required four years or more in the most affected areas.

The effects of oil on freshwater plants are much less studied (Lewis and Pryor 2013), although oil
tolerance has been observed in species of bulrushes. Transplants of the three-square bulrush sedge
(Scirpus pungens) collected from the shores of the St. Lawrence River survived, grew and produced new
shoots in sediments contaminated with MESA over a range of concentrations comparable to those
associated with oil spills (Longpré et al., 2000). Similarly, a spill of a heavy fuel (HFO 7102) into
Wabamun Lake, AB, had little direct effect on the abundance and productivity of softstem bulrush,
Schoenoplectus tabernaemontani (Thormann and Bayley 2008; Wernick et al. 2009). About 149,500 L
entered the lake after accumulating particles of soil and vegetation (Debruyn et al. 2007; Hollebone et al.
2011). Most remained close to shore, contaminating reed beds and sinking to the sediments as tar balls
(Hollebone et al. 2011). Bulrush biomass, cover, height and seed head production were unaffected despite
heavy oiling and residual contamination of sediments. Nevertheless, there was a major impact on
bulrushes caused by the cutting of oiled bulrushes at the sediment surface and the flushing and/or
vacuuming of contaminated sediments (Thormann and Bayley 2008). The amount of oiled plants was
reduced successfully, but so was the density and biomass of plant rhizomes. The severity of impacts
increased with water depth and included lower rates of seed germination, seedling growth, rhizome
survival and re-growth. As the predominant shoreline plant, the bulrush provided habitat for nesting
aquatic birds, including western grebes (Aechmophorus occidentalis), and reproduction of forage fish
species and northern pike (Esox lucius). Overall, the physical damage to bulrushes of cleanup operations
severely reduced the quality of habitat for key aquatic species for several years. Residual tar balls, some
of which leaked fresh oil, also persisted for up to 18 months (Hollebone et al. 2011).

Following the June 2012 spill of light sour crude to the Red Deer River, AB, Rood and Hillman (2013)
assessed the response of vegetation to floodplain and shoreline oiling and subsequent natural recovery at a
previously-established field site. Over the summer period, young balsam poplar trees lost leaves that had
been oiled, and growth was checked. However, as the oil weathered and dried on the trees, new shoots
emerged on contaminated stems and from roots underneath contaminated surface soils. The overall,
effects on trees were limited and due primarily to oil coatings that blocked carbon dioxide uptake and
water transpiration. This study demonstrated that leaving oiled vegetation undisturbed to recover on its
own prevented habitat damage due to cleanup and avoided colonization of damaged riparian lands by
invasive plants. However, the report was limited to one field season only, and no data were presented on
impacts on other types of vegetation or animals typical of the riparian zone.

In northern freshwater ecosystems, Hellebust et al. (1975) demonstrated that experimental spills of
Norman Wells crude oil to a lake in the MacKenzie River valley caused little effect on planktonic algae in
open water other than a blue-green algal bloom. However, the growth of attached filamentous algae was

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 133
inhibited in open waters, as was the growth and development of sedge species (emergent vegetation) in
wetlands at the edge of the lake and in oiled plots back from the edge of the lake. Effects were
characterized by the death of roots and emergent tissue of Carex species, and a lack of chlorophyll and
growth of Equisetum sp (horse tails) and moss (Scorpidium scorpiodes). Initial effects were caused by the
thick layer of oil that limited light penetration, but later effects were associated with the redistribution of
oil to the root zone of macrophytes and the accumulation of oil on substrates that supported periphyton.
Wetland macrophytes were highly sensitive to oil, particularly at the margins of open water, with
implications for overall productivity of wetlands and the destruction of habitat for aquatic or semi-
aquatic animal species.

Recommendation: Toxicity tests with plants should be conducted to define clear exposure-response
relationships, median effective (EC50) and threshold concentrations, and tissue concentrations
associated with effects. Field studies are needed to determine the comparative fate, behaviour and
effects of oil in wetlands with varied hydraulic regimes (bogs, fens, marshes, swamps), including
areas of permafrost. Habitat restoration methods for shoreline vegetation need to be developed to
ensure the rapid recovery of shoreline ecosystems and the optimum balance between oil cleanup
and habitat protection.

4.1.3 Invertebrates

The toxicity of crude and refined oils to invertebrates is derived directly from laboratory toxicity testing,
and indirectly from field studies of individual species or changes in structure of exposed invertebrate
communities (usually benthic invertebrates because they are relatively immobile and provide spatial
patterns of response). There is a reasonable database of acute median lethal concentrations (LC50) for
freshwater and marine species, but variations in sensitivity among species with different habitats, life
histories and developmental stages (i.e., embryos vs. adults) are not fully understood. For example, Lee et
al. (2000) found differences in sensitivity between the mystery snail (Viviparus georgianus) and the
mimic pondsnail (Pseudosuccinea columella) in a controlled oil spill experiment at a wetland site along
the St. Lawrence River (Ste. Croix, QC) that were attributed to feeding habits. V. georgianus, a
detritivore, assimilated contaminants directly from sediments, while P. columella, a herbivore, assimilated
contaminants indirectly, presumably from oiled vegetation. Although laboratory tests may provide an
objective and precise measure of toxicity, toxicity is not absolute and varies with species, life stage and
test conditions (Section 4.3), as well as the properties of the oil tested. For acute lethality, species
sensitivity distributions implied that marine invertebrates were no more or less sensitive to a variety of
test oils than fish species (Barron et al. 2013; details in section 4.2.1). As observed for other aquatic
species, dissolved petroleum hydrocarbons appear more bioavailable to invertebrates than particulate or
droplet oil (reviewed by Dupuis and Ucan-Marin 2015).

Effects on invertebrate species and communities were evident following the EVOS. The dominant taxa in
intertidal communities affected by the spilled oil and cleanup procedures included limpets, barnacles
(three species), mussels, snails (two species) and oligochaetes (worms) (Highsmith et al. 1996; Hooten
and Highsmith 1996). Effects were caused by oil toxicity, smothering and the loss of habitat when Fucus
beds were destroyed by cleanup activities. Cleanup procedures, such as washing, steaming or mechanical
removal, destroyed some animals, such as grazers (e.g., periwinkles), or their food supply (biofilms on
rocks). Two groups, barnacles and oligochaetes, increased in abundance because they rapidly colonized
space left by missing species and they consumed oil-degrading microbes. Potential impacts on other
organisms would include a reduction in food supply for consumers of intertidal plants and invertebrates,
such as shorebirds, diving ducks, otters and even bears. Recovery was incomplete two years following the
spill and contamination of mussel beds, including the mussels themselves, persisted more than three years
after the spill (Babcock et al. 1996).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 134
Following the DWH spill, laboratory experiments demonstrated the acute toxicity of mechanically- and
chemically-dispersed MC-252 (Macondo) oil to crabs (Rhithropanopeus harrisii, Callinectes sapidus)
(Anderson et al. 2014) and mysid shrimp (Americamysis bahia) (Hemmer et al. 2011). In contrast, field
studies during the spill found surprisingly few effects on invertebrates, with no reports of mass mortalities
on oiled shorelines or in coastal waters (reviewed by Beyer et al. 2015). Growth rates of shrimp were
reduced in coastal embayments affected by the spill (Rozas et al. 2014), likely due to the closure of
shrimp fisheries which increased their abundance (van der Ham and de Mutsart 2014) and depleted their
food supply (Beyer et al. 2015). Although caged filter-feeding mussels and oysters were exposed to oil-
contaminated particulates during the spill, there was no evidence of assimilation of oil-derived carbon in
samples collected after the spill, as indicated by stable isotope ratios in shell and soft tissues (Carmichael
et al. 2012b; Fry and Anderson 2014).

Oil dispersants enhanced the exposure, uptake and toxicity of oil to invertebrates, but the opposite was
true for fiddler crabs (Uca minax) perhaps because of complex relationships among oil droplet size,
methods of dispersion and species-specific pharmacokinetics (Chase et al. 2013). For example, in a
laboratory exposure of the copepod Calanus finnmarchicus to oil droplets from physically- and
chemically-dispersed Troll oil, filtration rates were inversely proportional to oil concentrations (Nordtug
et al. 2015) due to avoidance, clogging of copepod filtering structures or toxicity. Nevertheless, the
filtering rates were sufficiently high that zooplankton could conceivably enhance the rate of removal of
oil from water (Nordtug et al. 2015).

As with fish species, invertebrate embryo development is particularly sensitive to oil exposure. Exposure
of adult sea-ice amphipods (Gammarus wilkitzkii) to a continuous flow of water from an oiled bead
column caused little effect on adults but led to a high frequency of abnormalities in embryos carried in
their brood pouch (Camus and Olsen 2008). The fertilization success of oyster eggs (Crassostrea
virginica) was diminished by oil exposure, as were trocophore and D-stage development, with subsequent
increased rates of deformities, reduced overall survival and reduced activity of survivors (Laramore et al.
2014). These data suggest that trends and conclusions from studies with fish embryos (Section 4.1.4.2)
may also apply to invertebrate embryos.

The potential impacts of oil spills on terrestrial invertebrate species that interact with emergent
vegetation in salt marshes are not well studied (Beyer et al. 2015), a research need also relevant to
freshwater riparian lands affected by oil spills (e.g., the Kalamazoo River spill, Chapter 8).

Recommendation: Research is needed on interactions between the physiology, life history and
habitat characteristics of invertebrates and susceptibility to oil exposure and toxicity.

4.1.4 Fish

The toxicity of oil to fish is treated in greater detail than for other aquatic species, primarily because more
research has been done. Many of the interactions between the chemistry of oil and its toxicity to fish may
also apply to other species.

4.1.4.1 Acute toxicity

Fish are at risk of acute exposure to oil in the 24 to 48-hour period following a discrete spill to large
marine or freshwater ecosystems. However, fish kills are typically brief and localized because of the
rapid loss of the acutely lethal low molecular weight (LMW) components of oil due to dilution and
weathering. In contrast, acute mortality can be extensive when there is an ongoing point source of oil,
e.g., in a high-gradient river where oil is distributed rapidly but not highly diluted before weathering
occurs (e.g., pipeline break, Pine River, BC, reviewed in Hodson et al. 2011).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 135
The acute lethality of oil is associated primarily with LMW components, such as monoaromatics (e.g.,
BTEX), diaromatics (naphthalene and alkyl naphthalenes) and short-chain alkanes (< C12) (NRC 2005).
Metals, such as aluminum, copper, nickel, vanadium and mercury, occur in most crude oils, with
concentrations increasing from light to heavier crudes (reviewed by Dupuis and Ucan-Marin 2015).
However, these metals have not been associated with the acute toxicity of oil to aquatic biota, either
because concentrations are sub-toxic, or the metals are not sufficiently bioavailable in acute exposures.
Environmental contamination by metals, particularly mercury and vanadium, is evident near oil
production facilities (Lee 1983; Kelly et al. 2010). The sources are predominantly extraction processes,
and in the case of the Alberta oil sands, airborne and waterborne particulates released by the erosion of
surface soils during bitumen mining.

Sublethal effects of acute exposure are also evident and include biochemical signs of exposure to
polycyclic aromatic hydrocarbons (PAHs) (e.g., Ramachandran et al. 2004a), avoidance of oil solutions
and suppression of feeding behaviour (Lari et al. 2015) and delayed effects of embryo toxicity, such as
impaired swimming of juvenile fish (e.g. Mager et al. 2014; Section 4.7.1.2). However, most publications
on acute exposures concern only the lethality of petroleum hydrocarbons and modeling for risk and
impact assessments.

The acute lethality of oil is usually attributed to LMW hydrophobic petroleum hydrocarbons that partition
into lipid membranes and cause mortality by narcosis, a general term for an array of effects on lipid
membrane receptors and functions (Campagna et al. 2003). Narcotics include non-polar compounds (e.g.,
benzene, naphthalene) and polar compounds (e.g., heterocyclic hydrocarbons; Table 2.1), which differ
somewhat in their water solubility, bioavailability and toxicity. Larger compounds, such as alkyl PAHs
(alkyl PAHs) with three or more rings, contribute less to acute toxicity because they are taken up more
slowly than LMW compounds (McGrath and Di Toro 2009). Narcotics are sufficiently water-soluble to
reach maximum concentrations in water in contact with fresh, un-weathered oil within two hours (NRC
2005), but most are lost from water within two days by volatilization, biodegradation and dilution
(Chapter 2). As a consequence, lethal concentrations in water are not sustained unless there is an
ongoing discharge of oil (NRC 2005).

The acute toxicity of hydrocarbons to aquatic


biota can be estimated from octanol-water Kow (octanol-water partition coefficient): the
partition coefficients (Kow) (McCarty and ratio of the concentrations of a compound in n-
Mackay 1993). Aquatic organisms exposed to octanol (a water-insoluble phase that mimics tissue
solutions of hydrocarbons accumulate them in lipids and floats on water) to the concentrations in
proportion to Kow (Figure 4.1). Mortality occurs underlying water, at equilibrium.
when tissue concentrations reach a ‘critical
body residue’ or median lethal dose (CBR or BCF - The bioconcentration factor or ratio of
LD50) of approximately 5 mM in tissue lipid concentrations in tissue to concentrations in water.
(McCarty and Mackay 1993). There are linear
relationships between Kow and BCFs, or LD50s
across an array of hydrocarbons (Figure 4.1), so that LC50s for individual hydrocarbons can be estimated
mathematically from Kow. Curvilinearity reflects reduced rates of chemical uptake due to the increased
miscibility of compounds with water at lower values of Kow, and steric interference of diffusion of larger
compounds through membranes at higher values of Kow. Within the linear range of Kow models, mixtures
of hydrocarbons appear to act additively when expressed as toxic units (TU), allowing the estimation of
the toxicity of mixtures from the partial TU contributions of each component of the mixture.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 136
Figure 4.1. The general relationships among the octanol-water partition coefficients (Kow) of neutral organic
compounds and their bioaccumulation and acute toxicity (narcosis) to fish. The horizontal portion of the ‘body
residue’ regression represents the critical body residue or median lethal dose (LD50) of 5 mM common to most
narcotics. Adapted with permission from McCarty, L.S., and D. Mackay. 1993. Enhancing ecotoxicological
modeling and assessment. Body residues and modes of toxic action. Environmental Science &.Technology 27:1718-
1728. Copyright (1993) American Chemical Society.

These predictable relationships among properties Toxic Unit (TU): The ratio between the
of hydrocarbons form the basis for mathematical concentration of a compound in water and its
models to estimate the total dose received, the LC50. When water concentrations exceed the
dose-rate and the acute lethality of mixtures of LC50 (> 1.0 TU), test organisms will die
hydrocarbons from the concentrations measured in rapidly; when less than the LC50 (<1.0 TU),
water. These include the target lipid model (TLM) mortality occurs more slowly and fewer
(Di Toro et al. 2000; McGrath and Di Toro 2009); organisms die.
the Oil Toxicity and Exposure model - Two compounds act additively when 0.5 TU
(OILTOXEX) (French-McCay 2002); the of compound A plus 0.5 TU of compound B
PETROTOX model (Redman et al. 2014); and (or 0.3 of A and 0.7 of B) cause 50%
ecological models for the Barents Sea (Olsen et al. mortality of exposed organisms.
2013). These models provide a very convenient - Two compounds are termed synergists when
tool for estimating exposure of aquatic organisms toxicity appears more-than-additive.
to hydrocarbons, and the risk of acute lethality - Two compounds are termed antagonists
following oil spills. However, they are less when toxicity appears less-than-additive.
successful in predicting sublethal toxicity
associated with highly specific mechanisms of
action (MOAs), such as interactions of PAHs with specific cardiac receptors (Section 4.1.4.2).

A primary obstacle to modeling the acute lethality of actual spills is often the inadequate characterization
of the toxic components of oil or water samples (NRC 2005). The models are also based on a variety of
assumptions (text box) and cannot be validated before spills. Validation following a spill is challenging
due to the problem of finding dead organisms and characterizing their exposure to oil. Variance about
regressions relating log LC50 to log Kow is also large, with predicted values often one- to two-orders of

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 137
magnitude above or below measured values (McGrath and Di Toro 2009), and model errors are rarely
incorporated into quantitative uncertainty estimates in ecological risk assessments. The original LC50
values from which the models were first derived (e.g., Di Toro et al. 2000; French-McCay 2002) were
also generated prior to 2000 from a heterogeneous array of chlorinated and non-chlorinated compounds
and may not be representative of all petroleum hydrocarbons, especially for compounds with K ow values
greater than 5.3 (See section 4.3).

Assumptions inherent in models of oil mixture toxicity based on Kow and narcosis

 All chemicals in the TLM model share a single common mode of action (MoA) of narcosis;
 Hydrocarbon concentrations in water are constant with time;
 Tissue lipid concentrations are at equilibrium with water concentrations;
 Metabolism of hydrocarbons does not change tissue concentrations or rate of exposure;
 Metabolites of hydrocarbons do not differ in their toxicity from parent compounds;
 Concentrations of compounds in mixtures are adequately characterized;
 The Kow values of chemicals used to develop models encompass the range of Kow values for
petroleum hydrocarbons that contribute to acute lethality of oil;
 The models can predict the toxicity of hydrocarbons with log Kow > 6, beyond the linear range of
the statistical relationship (0-5.3) (McCarty and Mackay 2003);
 Models developed with data from un-substituted PAHs apply to alkyl PAHs;
 There are no synergistic or antagonistic interactions in mixtures, only additivity;
 There are no mixture interactions with other compounds in water, natural or anthropogenic; and
 Only dissolved hydrocarbons are present in the aqueous phase; non-aqueous phase oil (droplets) is
not present.

Recommendation: Models predicting the acute lethality of oil should be refined to narrow
confidence limits about model predictions by restricting toxicity data for model development to
petroleum hydrocarbons and to single test species.

4.1.4.2 Chronic and sublethal toxicity

Chronic and sublethal effects of oil on fish are associated with either chronic exposures (e.g., oil stranded
in sediments) or with the delayed or lingering effects of acute exposures. Effects range from the first
genetic and molecular responses of cells to impacts on rates of reproduction, growth, disease and survival.
Exposures to oil may be via respiratory uptake of hydrocarbons from water (most common), direct
contact with droplets (an emerging area of research), the diet or maternal transfer to eggs (not well
studied). Most studies of waterborne oil reference the concentrations of total PAHs (TPAHs) and/or total
petroleum hydrocarbons (TPHs) in test solutions that affect 50% of test organisms. The EC50 and LC50
values for chronic exposures range from 0.3 – 60 µg/L TPAH and 0.03 – 11 mg/L TPH (Figure 4.2,
Appendix D, Table D.1; Hodson et al. 2011; Logan et al. 2015). These concentrations are well within the
range of concentrations anticipated in surface waters near slicks of un-dispersed oil or in plumes of
chemically-dispersed oil and only brief exposures (one to 96 hours) to these concentrations can cause
embryo toxicity (McIntosh et al. 2010; Section 4.2.3). The variance among studies reflects differences in
species sensitivity, the mix of toxic components unique to each oil (Section 4.2.2), and the methods of
preparing, testing and characterizing solutions of oil (Section 4.3).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 138
Figure 4.2 Sublethal (EC50 values) and lethal (LC50 values) concentrations of total polycyclic aromatic hydrocarbons (TPAH) and total petroleum
hydrocarbons (TPH) reported for fish subject to acute (1-96 h) or chronic (>96 h) exposures to oil. The black rectangles represent the range of concentrations
reported to cause a given effect, and the solid circles indicate only one measurement was available. Data were drawn from studies reporting measured
concentrations (Appendix D, Table D.1: Kocan et al. 1996; Marty et al. 1997; Carls et al. 1999, 2008; Little et al. 2000; Brand et al. 2001; Couillard et al.
2005; Greer et al. 2012; Gulec and Holdway 2000; Pollino and Holdway 2002; Perkins et al. 2005; Shen et al. 2010; Olsvik et al. 2011; Wu et al. 2012;
Gardiner et al. 2013; Incardona et al. 2014; Mager et al. 2014; Martin et al. 2014; Dussauze et al. 2015; Madison et al. 2015; Sørhus et al. 2015).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 139
For juvenile or adult fish, the overall effect of oil exposure is to reduce fish production at spill sites
(Reviewed by Dupuis and Ucan-Marin 2015). Oil toxicity may include reduced rates of growth and
increased rates of infectious disease (Section 4.1.4.4), mutations and possibly cancer (Section 4.1.4.5) and
disruption of sexual maturation and reproduction.

Reproductive effects include lower concentrations of sex steroids, such as plasma estradiol, slower
gonadal growth and delayed time to sexual maturity (reviewed in Brown-Peterson et al. 2014). These
effects are more typical of chronic rather than delayed toxicity and may involve endocrine disruption by
hydrocarbons that mimic steroid hormones, or increased rates of hormone metabolism by enzymes that
oxygenate hydrocarbons. Establishing links between oil exposure and effects on reproduction is complex
because effects depend on the stage of maturation when fish are first exposed (Truscott et al. 1983).
Identifying and interpreting the effects of oil spills on reproductive performance of fish will be
complicated by a high variability in their sensitivity to endocrine disruption—timing is critical!

Research following the EVOS highlighted the high sensitivity of fish embryos to oil exposure, as judged
from the low concentrations of hydrocarbons causing toxicity (e.g., Carls et al. 1999; Appendix D, Table
D.1). Ecologically, embryo sensitivity is also imparted by their inability to avoid exposure to
hydrocarbons. Embryo-toxicity is often considered an effect of chronic exposure, i.e., an exposure that
constitutes a significant portion of an organisms’ life span or development stage. However, there are
instances of delayed effects, where a response typical of a chronic exposure occurs long after a brief or
acute exposure (Section 4.2.4). Delayed effects may not be obvious at the time of a spill, but may become
evident as changes in rates of recruitment, growth, reproduction, disease or mortality in the months or
years following a spill, depending on the life history of the species exposed. In this review, chronic
toxicity will refer to effects of exposures that exceed 96 hours or effects of acute exposures that are
evident 96 hours or more following exposure.

Embryo toxicity due to exposure to oil or to individual PAH is characterized by blue sac disease, a
syndrome that includes: pericardial and yolk sac edema; cardiotoxicity; craniofacial, ocular and spinal
deformities; degeneration of muscle somites, peripheral neurons and endothelial cells; fin erosion; failure
of swim bladders to inflate; altered behaviour and swimming ability; and induction of cytochrome P-450
(CYP1A) enzymes (e.g., Schein et al. 2009; deSoysa et al. 2012; Incardona et al. 2013; details in Section
4.1.4.4). Not all of these signs of toxicity are evident in embryos of each species exposed to oil, but all
species exhibit many of these signs.

The high sensitivity of embryos to oil toxicity is due


to the large number of genes that are up-regulated Cytochrome P450: The cytochrome P450
or activated only during embryonic development. It gene (cyp1a) codes for CYP1A proteins,
is critically important that these genes be activated enzymes that oxygenate PAHs as the first step
in the proper sequence, to the proper extent and at in their excretion. ‘CYP1A induction’ refers
the proper time for normal development. However, to the increased synthesis of CYP1A proteins
PAHs, the components of crude oil associated with when the cyp1a gene is up-regulated after
embryo toxicity (Section 4.1.4.3), can bind to specific PAHs or alkyl PAHs bind to the
protein receptors in cells to interfere with protein intracellular aryl hydrocarbon receptor protein
function or to activate a broad spectrum of genes. (Billiard et al. 2002).
Thus, oil exposure interferes with the carefully
orchestrated sequence of events essential for normal
development of tissues such as the heart, leading to impaired cardiac function (e.g., Incardona et al. 2004,
2009; Hicken et al. 2011).

The up- and down-regulation of a large number of genes represent the primary reactions to oil exposure,
indicating that hydrocarbons have interacted directly with cellular receptors to initiate gene transcription.
Activation of one gene can cause the secondary activation of many more, because they are often linked

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 140
directly or indirectly. One of the most commonly reported responses to exposure to crude oil and to
chlorinated hydrocarbons such as PCBs, is activation or up-regulation of the cyp1a gene, a member of a
large family of cytochrome P450 genes (e.g., Olsvik et al. 2011 2012; Sørhus et al. 2015). There are many
genes involved in embryo development that are ‘downstream’ of cyp1a and that are also activated with
exposure to PAHs. Activation of these genes at an inappropriate stage in embryonic development may be
one cause of PAH and dioxin-like toxicity (Billiard et al. 2008).

The CYP1A proteins are enzymes that oxygenate double bonds in PAHs as the first step in their
excretion. Thus, increases in CYP1A enzyme activity following gene activation cause a rapid increase in
the rate of excretion of PAHs. For example, the half-life of tissue retene (7-isopropyl-1-
methylphananthrene, a C-4 alkylphenanthrene) in rainbow trout is estimated as < 14 hours at 15 °C
(Fragoso et al. 1999). High rates of oxygenation of PAHs can prevent acute lethality by facilitating the
clearance of PAHs from tissues (Hicken et al. 2011). Not all PAHs will induce CYP1A enzymes (Basu et
al. 2001; Billiard et al. 2002), but all PAHs accumulated by fish exposed to oil are subject to oxygenation
by CYP1A enzymes because mixtures of PAHs in oil include those that cause induction. In vertebrate
species exposed to PAHs in oil, the up-regulation of cyp1a can increase CYP1A enzyme activity by up to
1,000-fold, in parallel with increasing toxicity (Brinkworth et al. 2003; Colavecchia et al. 2007; Madison
et al. 2015). Thus, CYP1A induction provides a reliable indicator of PAH exposure and potential effects
in vertebrates; invertebrate species do not express cyp1a genes and excrete PAHs more slowly than
vertebrates.

The oxygenation of PAHs can also release reactive oxygen species (ROS) and reactive metabolites of
PAHs into the cell (reviewed by Crowe et al. 2014). There are frequent reports that genes involved in
combating oxidative stress, suppressing mutations associated with the effects of ROS on DNA, and
responding to general cellular stress are up-regulated in oil-exposed fish embryos (e.g., Olsvik et al. 2011;
Madison et al. 2015). Exposure of cod larvae (Gadus morhua) to mechanically-dispersed oil caused the
expression of genes related to drug metabolism, endocrine system development and function and lipid
metabolism, with the highest response for cyp1a (Olsvik et al. 2011). Genes related to bone resorption
were up-regulated while those related to bone formation were down-regulated, consistent with
craniofacial and spinal deformities frequently reported in oil-exposed embryos.

Unlike acute lethality, there are multiple MOA that contribute to embryo toxicity, but they are not
mutually exclusive. Craniofacial deformities reflect the disruption of the normal transformation of neural
crest cells to specific cell and tissue types (de Soysa et al. 2012). Cardiac toxicity represents unique
interactions of PAHs with cardiac receptors, so that the heart does not develop to its proper form and
function (Incardona et al. 2004, 2006, 2013; Hicken et al. 2011; Scott et al. 2011). These interactions
were evident in vitro as impaired ion regulation, cell signaling and contractile rhythmicity in oil-exposed
cardio-myocytes of blue-fin (Thunnus thynnus) and yellow-fin (Thunnus albacores) tuna (Brette et al.
2014). Cellular effects corresponded to changes in cardiac function in whole embryos when oil-exposure
impaired atrial-ventricular signaling and coordination (Incardona et al. 2009). Edema and circulatory
failure of embryos may follow impaired cardiac function or be due to oxidative damage to lipid
membranes in endothelial cells of blood vessels following CYP1A metabolism of PAHs. CYP1A
induction was evident early in embryonic development and was prominent in gills, liver, kidney, eyes and
endothelial cells of blood vessels of fish embryos exposed to retene (Brinkworth et al. 2003), bituminous
sediments (Colavecchia et al. 2007) or oil (Incardona et al. 2013).

The predominant MOA underlying chronic toxicity varies with the PAH being tested. For example, the
embryo toxicity of pyrene depends on its interaction with the arylhydrocarbon receptor protein and
CYP1A induction (Incardona et al. 2006). In contrast, phenanthrene, retene and dibenzothiophene are
toxic even when CYP1A protein synthesis is blocked (Incardona et al. 2004, 2005; Scott et al. 2011). In
fathead minnow and white sucker embryos exposed to bituminous sediments, CYP1A enzymes were
induced in every tissue showing histopathology, and the extent of induction was correlated to rates of

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 141
mortality and pathology (Colavecchia et al. 2007). Although correlations imply a direct role of CYP1A
enzyme activity in toxic effects, they may simply reflect PAH concentrations in each tissue.

The metabolism of PAHs can also increase their toxicity and carcinogenicity (Section 4.1.4.5). Inhibition
of CYP1A induction demonstrated that metabolites of retene appear more toxic to trout embryos than the
un-metabolized parent compound (Hodson et al. 2007b; Scott and Hodson 2008; Scott et al. 2009).
Similarly, phenolic derivatives of 1-methylphenanthrene were up to four-fold more toxic to Japanese
medaka (Oryzia latipes) embryos than un-substituted 1-methylphenanthrene (Fallahtafti et al. 2012). The
pattern of –OH substitution affected toxicity and may reflect the propensity for biotransformation of
phenolic derivatives to p-quinones. Unlike acute lethality, chronic toxicity was unrelated to solubility or
Kow because all four OH-substituted congeners had the same modeled Kow. Similarly, co-exposure of trout
embryos to phenanthrene (does not induce CYP1A) and to an inducer of CYP1A enzymes caused greater
toxicity to trout embryos than exposure to phenanthrene alone (Hawkins et al. 2002), consistent with the
idea that metabolites were the toxic forms of PAHs. The MOA of alkyl PAHs likely includes oxidative
stress because the severity of embryo toxicity and signs of oxidative stress were reduced when rainbow
trout (Oncorhynchus mykiss) embryos were co-exposed to retene and Vitamin E, an antioxidant (Bauder
et al. 2005). The balance between the accelerated rates of excretion of PAHs and the production of excess
reactive intermediates will determine whether oxidative stress and pathology occur and will depend, in
turn, on the extent and duration of oil exposure and accumulation of TPAHs.

Research and monitoring of the effects on fish of the DWH spill focused largely on embryo toxicity in
laboratory or mesocosm experiments. The MC-252 oil caused similar cardiotoxicity as Alaska North
Slope (ANS) crude oil, and pathology and up-regulation of cyp1a genes by both oils were correlated to
concentrations of TPAH (Incardona et al. 2013). Oiled sediments from affected coastal embayments
caused the same toxic effects on zebrafish embryos as reference sediments spiked with MC-252 oil
(Raimondo et al. 2014) and were consistent with effects caused by chemically-dispersed oil, with a
threshold sediment concentration of TPAH between 2 to 27 mg/kg. Gulf killifish (Fundulus grandis)
embryos exposed to sediments from coastal areas contaminated by the DWH oil spill responded with the
typical signs of embryo toxicity (Dubansky et al. 2013) Similarly, killifish from areas contaminated by oil
showed the down-regulation of 1,070 genes and the up-regulation of 1,251 relative to un-oiled reference
areas (Garcia et al. 2012) and many gene responses indicated physiological and reproductive impairment
(Whitehead et al. 2012). The genomic responses were reproducible in laboratory exposures of killifish
(Pilcher et al. 2014) and spotted sea-trout (Cynoscion nebulosus) (Brewton et al. 2013) to MC-252 oil,
with clear activation of excretory pathways (e.g., cyp1a). Activation of cyp1a was associated with growth
depression in sea-trout (Brewton et al. 2013) and with gene expression in killifish indicative of changes in
gene transcription, cell cycle progression, RNA processing, DNA damage, oxidative stress and apoptosis
(Crowe et al. 2014; Pilcher et al. 2014). Impacts on the reproductive biology of spotted sea-trout sampled
from oil-contaminated coastal embayments of Louisiana and Mississippi were reported by Brown-
Peterson et al. (2014). Impacts included delayed maturation, smaller relative gonad sizes and a reduced
frequency of spawning, but no biological or chemical indicators of exposure to oil were reported.

In general, there is a good understanding of the waterborne concentrations of TPHs, TPAHs and specific
un-substituted and alkyl-substituted PAHs causing chronic toxicity, and sufficient data to support risk
assessments of oil spills.

Recommendation: Additional research is needed to reduce the uncertainties in translating


laboratory data for risk and impact assessments. In particular, relationships need to be established
among waterborne concentrations of hydrocarbons, tissue concentrations, genomic responses and
toxic effects. Furthermore, mechanisms of action and critical exposure periods for impacts on the
reproductive biology of sexually maturing fish need to determined, as does the role of PAH
metabolism in toxicity of oil to fish.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 142
4.1.4.3 Hydrocarbons causing chronic toxicity

To establish the connection between exposure and environmental impacts of spilled oil, it is essential to
know which constituents of oil cause toxicity and at what concentrations. The components of oil that
cause chronic toxicity to fish embryos are considered to be alkyl PAHs (Table 2.1), including
heterocyclics such as alkyl dibenzothiophenes (Rhodes et al. 2005). TPAHs comprise about 0.2 to 6.5%
by weight of oil, depending on its source and type, and alkyl PAHs comprise 80 - 95% of TPAHs in crude
and refined oils (Wang et al. 2003). Alkyl PAHs partition from oil to water more slowly than
monoaromatics, reaching peak concentrations within four to 12 hours (NRC 2005). Because of their
lower volatility, alkyl PAHs persist in water longer than monoaromatics and are dissipated primarily by
advection and diffusion. Over weeks or months, alkyl PAHs stranded on substrates will continue to
partition from oil into water (NRC 2005), as was evident from a comparison of the cardiotoxicity to
zebrafish (Danio rerio) embryos of two crude oils (Jung et al. 2013).

A number of indirect approaches have been used to identify the components of oil causing chronic
toxicity to fish embryos. Correlations between concentrations of alkyl PAHs in test solutions and the
chronic toxicity of ANS crude oil (Table 2.2, Appendix B) to fish embryo development were reported by
Carls et al. (1999). Using a toxic units approach, Barron et al. (2004) evaluated four mechanistic models
of the chronic toxicity of the constituents of oil test solutions to pink salmon (Oncorhynchus gorbuscha)
and Pacific herring (Clupea pallasi) embryos and concluded that toxicity was best described by summing
the toxic unit contributions of alkyl phenanthrene concentrations.

More direct evidence was provided by studies of the chronic embryo toxicity of individual PAH
congeners that produced the same syndrome of blue sac disease as did exposure to crude and refined oils.
The PAHs tested are commonly found in oil and included 3- to 5-ringed un-substituted PAH (Hawkins et
al. 2002; Incardona et al. 2004, 2006, 2011; Rhodes et al. 2005), alkyl phenanthrenes (Billiard et al. 1999;
Kiparissis et al. 2003; Turcotte et al. 2011), hydroxylated derivatives of alkyl phenanthrenes (Fallahtafti
et al. 2012), alkyl chrysenes and alkyl benzo[a]anthracenes (Lin et al. 2015), and dibenzothiophene and
dimethyl dibenzothiophene (Rhodes et al. 2005). Effects-driven-chemical-fractionation (EDCF) of crude
oil (Hodson et al. 2007a) and heavy fuel oil (Adams et al. 2014b; Bornstein et al. 2014) linked embryo
toxicity to fractions and sub-fractions of oil that contained 3- to 5-ringed alkyl PAHs. Alkyl PAHs were
progressively enriched in serial fractions of HFO in parallel with increased toxicity, and comprised up to
25% of total hydrocarbons in toxic fractions. Fractions dominated by LMW compounds, including BTEX
and alkyl naphthalenes (2-ringed PAH), or by high molecular weight (HMW) aliphatics (waxes), resins or
asphaltenes (Table 2.1) were non-toxic at the concentrations tested.

Overall, many of the chronic effects of oil on fish embryos can be attributed to 3- to 5-ringed alkyl PAHs.
The chronic toxicity of PAHs to fish has been associated with their binding to the AhR protein (Section
4.1.4.2). However, other components of oil may also be AhR agonists, as indicated by in vitro reporter
gene assays of hydrocarbons extracted from aquatic worms exposed to SARA fractions (Section 2.1.1) of
eight different crude and fuel oils (Vrabie et al. 2012). The AhR agonists were aromatic, hydrophobic,
resin-like compounds primarily in mid- to high-boiling point fractions, and some were resistant to
biotransformation in vitro. There were too many compounds in isolated fractions to identify specific
agonists, although these characteristics correspond to some un-substituted and alkylated PAHs and
heterocyclic aromatic compounds.

In all EDCF studies, most components of toxic fractions were unidentified and may have included polar
compounds not typically analyzed in detail in oil. Polar compounds in a biodegraded Norwegian Sea
crude oil and in 14 fractions of water accommodated fraction (WAF) solutions created by high
performance liquid chromatography (HPLC) separation were identified as agents affecting rainbow trout
liver cells in culture (Melbye et al. 2009). Responses included induction of CYP1A and vitellogenin (egg
yolk protein, an indicator of estrogen-like compounds), as well as cytotoxicity. The main contributors to

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 143
all three responses were the water-soluble components of the polar fraction that contained 70% of the
hydrocarbons, while fractions enriched with PAHs and alkyl PAHs were less potent. However, ketone
and quinone derivatives of a wide array of 3- to 6-ringed non-alkyl PAHs showed significant potency for
activating the cyp1a gene in mammalian cell assays (Misaki et al. 2007). In contrast, a survey of the
endocrine-disrupting potential of 11 crude and refined oils using in vitro mammalian cell assays
demonstrated that all contained compounds with some estrogenic or androgenic potency, but potencies
were 104 to 107-fold lower than those of the reference compounds estradiol and testosterone (Vrabie et al.
2010, 2011). Although polar compounds, such as phenols, are typically associated with endocrine effects,
no equivalent fractionation studies have been conducted to identify whether components of oil cause
endocrine disruption in vivo.

Similarly, the toxicological significance of the oxygenated hydrocarbons measured in MC-252 oil (Table
2.2) over an 18-month period of weathering following the DWH oil spill is unknown (Aeppli et al. 2012).
However, aromatic and polar compounds extracted from biodegraded ANS crude oil were chronically
toxic to embryos of inland silversides (Menidia beryllina) (Middaugh et al. 2002). Extracts of the saturate
fraction were non-toxic at the concentrations tested, but the fractions recombined represented the additive
effects of the aromatic and polar compounds. This experiment did not analyze oxygenated compounds or
the toxicity of untreated oil, but they demonstrate the presence of significant concentrations of residual
toxic compounds in biodegraded oil. It is essential that oil toxicity tests include as complete an analysis of
test solutions as possible, including polar fractions. Routine analyses by GC-MS identify about 40 - 50
different PAHs or families of alkyl PAHs, in addition to BTEX and saturates.

Recommendation: Research is needed to identify the fractions petroleum hydrocarbons associated


with different toxic effects.

Models have been developed for ERAs to predict the chronic toxicity of oil (e.g., TLM) (McGrath and Di
Toro 2009). However, chronic toxicity is not modeled directly. Instead, acute lethality (LC50) is
calculated by the model, and chronic toxicity is estimated from the average ratio of concentrations
causing acute and chronic toxicity (acute-chronic ratio or ACR) for algae, invertebrates and fish.
However, observed ACR values ranged 10-fold higher and lower than the average, likely because of large
differences in the physiology and biochemistry among widely differing plants and animals used as test
species. All of the assumptions underpinning the acute lethality model (Section 4.1.4.1) are inherent in the
chronic toxicity model, with additional critical assumptions about chronic toxicity that weaken its
application to risk assessment (text box). Not surprisingly, the TLM model does not predict the chronic
toxicity of petroleum hydrocarbons very accurately, and a further ‘correction factor’ of 2.0 must be used
to compensate for the average under-estimate of toxicity, which in itself is variable (McGrath and Di Toro
2009). While these models have been applied to the results of published toxicity tests with a variety of
species and a variety of chemical mixtures (McGrath and Di Toro 2009), the results were not very
satisfactory, primarily because the case studies examined were not designed for this purpose.

Assumptions inherent in models of acute lethality applied to chronic toxicity

 The MOAs for chronic toxicity and acute lethality are the same (narcosis), regardless of a wide
array of sublethal effects caused by oil (Section 4.1.4.2) and large differences in the genetics and
physiologies among algae, invertebrates and fish; and
 A single mean ACR derived from tests with algae, invertebrates and fish predicts the chronic
toxicity of all PAH to fish embryos.

Although current models carry a high risk that predicted toxic effects could be under- or over-estimated
by a wide margin, the advantages of modeling mixture toxicity for ERAs and EIAs are very compelling,

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 144
considering the cost and time needed to run chronic toxicity tests for every oil that might possibly be
spilled or every oil sample collected at a spill. Recently, Redman et al. (2014) used the PETROTOX
model to estimate the chronic toxicity of HFO and fractions of HFO created for an EDCF using the
reported concentrations of PAHs in each fraction. There was a good correspondence between toxicity
predicted from the PETROTOX model and toxicity measured directly (Adams et al. 2014b), although
more research is needed to identify sources of error in both the model and the EDCF.

Recommendation: Models of chronic toxicity must be developed from results of chronic


toxicity tests and not from acute toxicity tests via application factors.

4.1.4.4 Physiological and pathological responses

The molecular responses of fish exposed to oil in laboratory experiments (Section 4.1.4.2) suggest that
there should be equivalent responses in fish surviving an oil spill and related changes in physiological
functions that could affect the performance of those fish. Following the 2007 oil spill from the Hebei
Spirit off the coast of Korea, Jung et al. (2011, 2012) measured the activity of liver CYP1A enzymes and
the concentrations of PAH metabolites in bile of two benthic species, marbled flounder
(Pseudopleuronectes yokohamae) and flatfish (Paralichthys olivaceus), sampled in 2008 and 2009. These
biomarkers1 provided clear evidence that fish were accumulating and responding to PAHs from the oil
spill in a spatial pattern that corresponded to measured concentrations of PAHs in sediments. The
consistent responses relative to fish from a reference area during the two years following the spill
indicated the persistence of bioavailable hydrocarbons in sediments.

Balk et al. (2011) also observed similar CYP1A and bile metabolite responses of Atlantic cod and
Atlantic haddock (Melanogrammus aeglefinus) sampled from the North Sea close to oil drilling and
production platforms. Fish captured close to the platforms also showed enhanced activities of antioxidant
enzymes, changes to fatty acid composition, increased concentrations of arachidonic acids and reduced
concentrations of Vitamin E (antioxidant), indicating not only accumulation and metabolism of petroleum
hydrocarbons but also signs of oxidative stress. Elevated concentrations of DNA adducts in haddock
suggested a reaction of PAH metabolites with double bonds in DNA, mutagenicity and a potential for
carcinogenicity. Oxidative stress was also suggested by increased non-enzymatic antioxidant capacity in
serum of Gulf killifish (Fundulus grandis) exposed to MC-252 oil (Crowe et al. 2014).

Effects of PAHs and oil exposure on the development and function of the embryonic heart should impair
the capacity of survivors to swim, capture prey and avoid predation. Effects on embryonic circulation
persist into the juvenile stages of fish, a latent effect evident long after exposures to oil. Embryos of mahi-
mahi (Coryphaena hippurus) that survived 48 hour exposures to weathered MC-252 oil (1.2 µg/L TPAH)
showed a reduced critical swimming speed in swim tunnels when tested 24 days after exposure (Mager et
al. 2014). Fish exposed to oil (30 µg/L TPAH) for 24 hours as juveniles were 30-fold less sensitive than
fish exposed as embryos. For both groups, no changes occurred in metabolic rate or respiratory efficiency
(Mager et al. 2014). Delayed effects on capacity to swim were also observed in adult zebrafish 10 to 11
months after exposure as embryos to ANS oil (24-36 µg/L TPAH), a concentration that caused 10 - 15%
mortality of embryos during exposure (Hicken et al. 2011).

Exposure of fish to crude oil is also associated with impaired immunocompetence and an increased
susceptibility to disease, likely due to the effects of PAHs on cellular and humoral immune responses
(Reynaud and Deschaux 2006). However, relating the prevalence of disease to oil exposure in field
studies is complicated by the number of uncontrolled environmental stressors that act on fish
simultaneously or in different sequences and that can affect immune responses. For example, a variety of
hypotheses were examined to explain the crash in abundance of Pacific herring in Prince William Sound

1
Not to be confused with the term ;biomarkers’ used to describe internal hydrocarbon standards in oil analyses; Chapter 2

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 145
following the 1989 EVOS (Pearson et al. 1999), including: a direct effect of oil exposure on survival and
recruitment of larvae; an indirect effect of oil on the food web supporting herring; the population outgrew
its food supply; an epizootic of viral hemorrhagic septicemia (VHS) was triggered by environmental
conditions, an over-abundance of herring, impaired immune function, and the stress of oil exposure;
overharvesting; and natural chaotic population dynamics. Pearson et al. (1999) concluded that there was
no clear link between the decline in herring abundance and oil exposure and that an array of natural
factors contributed to the decline. This view was reinforced by Marty et al. (2003) who could not
establish a temporal link between epizootics and the 1989 oil spill. In contrast, Thorne and Thomas
(2007) related post-spill epizootics in herring with their surfacing behaviour and interaction with oil slicks
at the time of the spill, genetic abnormalities that corresponded to oil toxicity and trends in fishery
statistics and distribution of herring predators (sea lions). They concluded that the collapse of the herring
fishery was a five-year event triggered by the oil spill.

The immune response to pathogens is also complex and can lead to different interpretations of cause and
effect, depending on circumstance. For example, when juvenile Pacific herring were exposed to different
concentrations of oil in water and their immune systems were challenged by subsequent exposures to a
pathogenic bacterium that causes vibriosus, their response depended on duration of oil exposure
(Kennedy and Farrell 2008). After one day of exposure, resistance to infection was enhanced, while the
opposite occurred after 57 days of exposure. Resistance was associated with a short-term physiological
stress response while long-term sensitivity was associated with impaired ion regulation.

Fish diseases can also enhance sensitivity to oil exposure. The cumulative mortality of juvenile flounder
(Pseudopleuronectes americanus) increased with an eight-week exposure to sediments contaminated with
Hibernia crude oil (100 to 2,200 μg/g). At each sediment concentration, mortality was greater in fish
infected by the hemoprotozoan, Trypanosoma murmanensis, relative to un-parasitized controls (Khan
1991).

During the DWH oil spill, unusual external lesions on red snapper (Lutjanus campechanus) in the
northern Gulf of Mexico were associated with several pathogenic bacteria. Arias et al. (2013) proposed a
causal association between lesion prevalence and exposure to oil. They assumed that the affected fish
were immunocompromised by oil exposure, corresponding to previous experience with other fish species
(reviewed by Dupuis and Ucan-Morin 2015). However, no evidence was collected of actual exposure to
oil, so this proposed cause-effect relationship was not well supported.

Recommendation: Research is needed on the cause of epidemics in fish and the interactions among
pathogens, environmental stressors, exposure to oil and prevalence and severity of disease,
including chemical and biological markers of oil exposure and effects.

4.1.4.5 Gene mutations and cancer

Mutagenicity and carcinogenicity are often linked to metabolism of specific PAHs. It is well-established
that oxygenation of benzo[a]pyrene (BaP; Table 2.1), a pro-carcinogen, by CYP1A enzymes can generate
an array of non-carcinogenic and carcinogenic metabolites (e.g., diol-epoxides of BaP) (reviewed by
Shugart 1995). Carcinogenic metabolites cause gene mutations by covalent bonding to cellular DNA and
interfering with normal replication.

Non-alkylated, pro-carcinogenic PAHs such as BaP, are found at measurable concentrations in crude and
refined oils (Yang et al. 2011) but they comprise much less than 15% of total PAHs (Wang et al. 2003).
Although methyl phenanthrenes are demonstrated mutagens in cell culture assays (Lavoie et al. 1981,
1982), the mutagenicity of most alkyl PAHs is not well characterized, primarily because few of the
thousands of congeners of alkyl PAHs are commercially-available for testing. No papers were found that
report carcinogenicity of alkyl PAHs to fish.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 146
Recommendation: Research is needed to assess the mutagenicity and carcinogenicity of alkyl PAHs
to fish.

There has been a variety of studies reporting mutagenicity in fish and invertebrates following marine oil
spills (e.g., Bolognesi et al. 2006), but none so far have associated elevated rates of cancer in fish with
exposure to oil. Cancers are most evident in benthic species (sessile, in constant contact with bottom
sediments, feed on benthic organisms), such as English sole (Pleuronectes vetulus; marine) (Myers et al.
2003) and brown bullhead (Ameiurus nebulosus; freshwater) (Baumann et al. 1996). Benthic species
experience a longer and more intense exposure to potential carcinogens than those in the pelagic zone.
The absence of PAH-related cancers in fish following oil spills may reflect the slow onset of chemically-
induced cancers (months to years; Baumann et al. 1996), the even slower appearance of frank tumours
and the lack of long-term monitoring of fish populations at sites of major oil spills. Interpretation of data
on cancer prevalence in fish is also confounded by:

 The age of fish sampled at contaminated and reference sites (prevalence increases with age);
 Co-exposure to other carcinogens not derived from oil;
 Migration of fish to and from sites of exposure; and
 Cell proliferation in response to infections by viruses, bacteria and parasites (Baumann et al.
1996).

In contrast, at sites where high concentrations of pyrogenic PAHs predominate in sediments due to urban
and industrial development, there is clear evidence of cancer in locally-resident fish (Myers et al. 2003).
Cancer is characterized by premature mortality of older fish and corresponding changes in the age
structure of the population. However, overall abundance may not change because sexual maturation and
reproduction precede mortality (Collier et al. 1992; Bauman et al. 1996). The discovery of tumours in fish
exposed to pyrogenic PAHs may be related to the higher concentrations of carcinogenic PAHs, such as
BaP and nitrogen heterocycles, and to the long-term intensive studies typical of sites affected by urban
and industrial pollution.

Recommendation: Long-term monitoring of impacts following an oil spill should include


histological examinations of fish to assess the prevalence of cancer and other diseases in relation to
oil exposure.

4.1.4.6 Population and ecosystem consequences of oil toxicity

The effect of oil exposure on cardiac development, structure and function in fish embryos with subsequent
impairment of swimming stamina (reviewed in Section 4.1.4.4) indicates that population level effects on
fish will follow an oil spill due to a decreased ability of juveniles to capture prey and avoid predation.
The ecological relevance of embryo toxicity is the potential for reduced rates of growth, survival,
abundance and productivity. These effects would not be predicted from standard embryo toxicity tests
because they are delayed and evident only when the fish are free-swimming, months or years following
exposure. For example, Heintz et al. (2000) observed a 15% reduction compared to controls in the rate of
return of pink salmon released to the Pacific Ocean following sublethal exposures to oil as embryos. In
contrast, pink salmon exposed to sublethal concentrations of oil as juvenile fish showed no change in
returns as adults, compared to controls (Birtwell et al. 1999). The different outcomes of these two similar
studies are consistent with the observed high sensitivity to oil of the developing heart of embryos and the
much lower sensitivity of juveniles (Mager et al. 2014).

The behaviour and migration of salmon can also be affected by exposure of adults. Although Brannon et
al. (1986) found no change in the ability of healthy adult chinook salmon to return to their home stream
after a brief (one hour) exposure to Prudhoe Bay crude oil (PBCO), adult coho salmon (Oncorhyncus
kisutch) avoided a salmon ladder contaminated with oil (Weber et al. 1981). The numbers climbing the

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 147
ladder decreased in linear proportion to increasing oil concentrations. Avoidance is good news from the
perspective of survival, but bad news if salmon do not complete their spawning migration and reproduce.

Recommendation: Further research is needed to determine whether poor returns of adult pink
salmon that survived exposure to oil as embryos are related to changes in survival at sea, swimming
capacity or response to environmental cues that guide migration.

Despite the apparent high sensitivity of pink salmon embryos to oil exposure in the lab (Heintz et al.
1999), in situ (reviewed in Ballachey et al. 2014) and during migration following exposure to oil as
embryos (Heintz et al. 2000), the case for an effect of the EVOS on pink salmon populations in Prince
William Sound remains controversial. Population modeling (Geiger et al. 1996; Heintz, 2007) suggested
significant impacts of low-level exposures to oil-derived PAHs. However, the technical basis for
characterizing impacts and linking oil exposure to effects on recruitment and abundance was questioned
by Brannon et al. (2012), mainly because of the complexity of ecological interactions with the population
dynamics of salmon and the absence of clear evidence of a decline in natural production of salmon
(Brannon et al. 2006). Hatcheries and tributaries to Prince William Sound produced so many juvenile fish
annually that deficits in production in tributaries affected by the spill were not easily detected. However,
the opposite would occur if an oil spill coincided with the spawning of a rare or threatened species, such
as the eulachon (Thaleichthys pacificus), in the estuaries of west coast rivers (Hodson et al. 2011).

For the DWH oil spill, there were surprisingly few reports of impacts on wild fish at sea or along the
coast of the Gulf of Mexico despite numerous laboratory studies demonstrating effects of oil exposure on
fish embryos and adults. Recruitment of fish to coastal marshes of Alabama appeared unaffected, as
indicated by no change in marsh-associated resident or transient nekton (all actively-swimming aquatic
species) over a two-year period following the spill (Moody et al. 2013). However, there were no
indicators of potential exposures to oil, other than location. Similarly, the survival of fish embryos in
coastal seagrass habitat appeared unchanged in 2010 following the DWH oil spill, in comparison with a
previous five-year dataset (Fodrie and Heck 2011). Recruitment of red snapper on artificial reefs was
more strongly correlated to dissolved oxygen concentrations and the presence of predators (age-1 fish)
than to exposure to oil, and there was no evidence of year-class failure (Szedlmayer and Mudrak 2014).
Similarly, a comparison between satellite-based estimates of oil distribution and the modeled spawning
habitats of blue-fin tuna indicated that less than 10% of offshore spawning grounds, and 12% of the larval
distribution area, would overlap with the distribution of oil slicks (Muhling et al. 2012), suggesting little
impact on recruitment.

Fodrie et al. (2014) reviewed the lack of correspondence between observed effects on fish populations
and expectations of impacts from laboratory toxicity studies. They cited several major factors that obscure
population-level responses: high spatio-temporal variability in population demographics; movement of
fish into and out of affected areas (natural ‘re-stocking’); fisheries closures that removed a major source
of harvest mortality and that compensated for oil-related losses of fish; food web compensation (enhanced
productivity due to input of oil-derived carbon); and delayed effects over several generations. Other
factors could dampen population responses, such as behavioural avoidance of oil exposure, rapid dilution
of oil in the open ocean (particularly following chemical dispersion) and population compensation, such
as density-dependent changes in egg, juvenile and adult survival. As well, expectations of effects may be
heightened by potential high-sensitivity bias in laboratory tests, depending on experimental designs.
Without an obvious fish kill following a spill, as was the case with the Pine River in northern British
Columbia (see below) or Asher Creek, MO (reviewed by Dupuis and Ucan-Marin 2015), it is very
difficult to make an unambiguous case for sublethal, delayed or long-lasting effects. Nevertheless,
numerous laboratory studies (Appendix D, Table D.1) leave little doubt about the potential toxicity of
spilled oil. Difficulties in making the case for oil spill impacts in situ do not mean that population effects
do not occur; rather, highly variable population dynamics prevent firm conclusions. Overall, a potential
bias in studies of effects on fish populations appears to favour ‘false negatives’ (no effects perceived on

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 148
fish populations when there were effects) rather than ‘false positives’ (effects perceived that were non-
existent or due to other factors). These uncertainties need to be resolved and the capacity to define
population-level impacts improved.

Recommendation: (adapted in part from Fodrie et al. 2014): Research is needed to: 1) establish an
observation network to provide pre-spill baseline data on physical, chemical and biological
characteristics of aquatic ecosystems, including fish population, abundance and productivity,
particularly in areas where spills are likely; 2) determine individual-level measures of genomic,
physiological and demographic responses in baseline surveys as indicators of overall health, rates of
disease and exposure to oil and other contaminants (e.g., Stagg et al. 2000); 3) develop a better
understanding of the basic ecology and early life-history of fishes including movement, diet, habitat
use and longevity; and 4) establish the links between hydrocarbon exposure of individuals and their
molecular and cellular responses to signs of toxicity and pathology.

In freshwater ecosystems, EIAs and studies of the rate of recovery of oiled ecosystems are also limited by
a lack of data, appropriate assessment methods and the challenges of sunken oil. There are few field
studies of the effects of spilled oil on freshwater fish, other than counts of the dead and dying
immediately following a spill. The weakness of short-term census data is that counts of dead fish are often
inaccurate because of a low counting efficiency in wilderness areas that are hard to access. Following a
spill of sour crude into the Pine River, BC, in 2000, only a partial count of dead fish (1,600) was reported
because the river was difficult to access quickly. Dead fish either decayed or were consumed by
scavengers before they could be counted. As a consequence, estimates of the total kill varied widely, from
25,000 to 250,000 (Hodson et al. 2011).

Monitoring of time-to-recovery of oiled ecosystems relies too heavily on standing stocks of fish. Without
measuring fish population demographics, community structure, movements and productivity (e.g., rates
of reproduction and recruitment), actual recovery cannot be separated from ‘re-stocking’ by fish from un-
oiled regions of a watershed. Only small amounts of sunken residual oil in sediments could significantly
reduce embryo survival and recruitment of sediment-spawning fish species. Following the 2005 spill of
HFO 7102 to Wabamun Lake, AB (Section 4.1.2), most residual oil was trapped in shoreline reed beds
(Hollebone et al. 2011), but some was measured in offshore spawning shoals of lake whitefish
(Coregonus clupeaformis) (Short 2008). Debruyn et al. (2007) measured the survival and prevalence of
deformities typical of oil exposure of whitefish embryos held in egg containers installed on contaminated
and reference shoals in November 2005, four months after the spill. In April 2006, the prevalence of
severe skeletal deformities typical of oil exposure was significantly higher among hatched survivors at the
exposed site than at a reference site. The rates of pathology were correlated to the amount of TPAHs
accumulated by passive samplers installed adjacent to the egg containers.

Recommendation: Research is needed to: 1) develop methods to find and measure residual oil in
reed beds and bed sediments of lakes and rivers; 2) assess oil exposure and toxicity of residual oil to
embryos of fish species with different spawning media (e.g., sediments, vegetation, water column);
and 3) measure the density, survival and recruitment of naturally-spawned fish embryos in oil-
contaminated ecosystems.

Some of these research needs may be met by studies of actual ‘spills of opportunity’. The current
understanding of the ecological effects of oil in marine and coastal environments is derived largely from
studies of the 1989 EVOS. Innovative research on the distribution, effects, and persistence of oil in Prince
William Sound is cited throughout this chapter and clearly informs current research on the 2010 DWH
spill. For Canada, the legacy of the extensive research on these two spills is a much stronger capacity for
ERAs and EIAs and the opportunity to advance our knowledge by building on a solid technical base of
understanding.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 149
Much could also be learned if there were small areas of contaminated sites ‘set aside’ for natural
attenuation of spilled oil. A well-known example is a segment of contaminated shoreline untouched 30
years following the Arrow spill of Bunker C in Nova Scotia (Lee et al. 2003a). However, the greatest
potential for understanding ecological effects of oil pollution may be studies of oil spilled to experimental
marine and freshwater ecosystems. Small-scale experimental spills allow the fewest practical limitations
to field investigations and the greatest opportunity for research where gradients of natural environmental
stressors can be controlled. To date, the few experimental spills in Canada have been extremely
informative (e.g., Sergy and Blackhall 1987; Hodson et al. 2002; Venosa et al. 2002; Garcia Blanco et al.
2007), but their scopes were limited by the questions being asked. Additional experimental spills are
needed to keep pace with Canada’s rapidly expanding oil transportation network and the risks to a wide
array of ecosystem types.

4.1.5 Reptiles and amphibians

No species of sea turtles nest in Canada, but five of the world's seven species are sighted on an occasional
basis on both the east and west coast, including green (Chelonia mydas), Atlantic loggerhead (Caretta
caretta) (and a hybrid of the two), Atlantic ridley (Lepidochelys kempii), olive ridley (L. olivacea)
and leatherback (Dermochelys coriacea) turtles (www.thecanadianencyclopedia.ca/en/article/turtle/).
While several studies of turtles in the northern Gulf of Mexico were published before and after the DWH
oil spill, no links were established among the status of different species, indicators of health and exposure
to oil (Beyer et al. 2015). However, all studies identified a high risk of severe damage to populations
should there be contact with oil and emphasized the need for conservation efforts to sustain populations
that are under pressure from a variety of anthropogenic stressors.

Native species of turtles, other reptiles (43 species total) and amphibians (46 species) inhabit terrestrial
and freshwater habitats of southern Canada, with their northern distribution limited by the boreal forest.
No species inhabit the tundra zone (www.thecanadianencyclopedia.com/en/ ). Oil spills to aquatic
ecosystems and riparian lands will create significant risks of exposure of reptiles and amphibians to
petroleum hydrocarbons, as well as potential impacts on their habitats from oil spill cleanup. Turtles may
be coated with oil or inhale volatiles or aerosols of oil when they surface and the eggs of species that
spawn on shore may be exposed to stranded oil. Numerous studies have demonstrated that turtles,
particularly long-lived predatory species at a high trophic level, bioaccumulate persistent organic
pollutants and develop dose-dependent deformities (Rowe et al. 2009). However, the evidence for effects
of petroleum hydrocarbons, particularly PAHs, on turtles is less established. Bell et al. (2006) reported
deformities in snapping turtles (Chelydra serpentina) and painted turtles (Chrysemys picta) from a
wildlife refuge in Pennsylvania that was heavily contaminated by past industrial activity with metals,
PAHs and other organic compounds. The prevalence and severity of deformities far exceeded those at a
reference site, and were equivalent to or greater than those reported at other contaminated sites. Although
measured concentrations of PAHs in turtles at the contaminated site were far greater than at the reference
site, the cause of deformities was not conclusive because the study had insufficient resources to analyze
metals and other organic compounds. At this same refuge, the survival, home-range and water
temperature preference of four species of freshwater turtles contaminated by an oil spill were unaffected
relative to un-oiled turtles, despite their capture, rehabilitation by cleaning and release (Saba and Spotila
2003). There was a similar experience following the 2010 spill of dilbit to Talmadge Creek and the
Kalamazoo River. Large numbers of juvenile and adult turtles were rescued, cleaned and released
(USFWS 2015) and because some were captured and released up to five times, it was concluded that they
were unharmed by the experience.

Although the short-term effects of oiling on free-swimming turtles appear negligible, there were no
studies of the chronic effects of the dilbit spill or of the effects on the embryonic stages of turtles exposed
to dilbit on the exterior of egg shells. For Arabian Light crude oil, eggs of snapping turtles showed a very
low sensitivity to un-dispersed oil (WAF) or to chemically-dispersed oil (chemically-enhanced WAF or

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 150
CEWAF) (Rowe et al. 2009). After single doses of WAF or CEWAF were percolated through nests
containing viable eggs, there were no significant effects on hatching success, weight, length, lipid content,
DNA damage or overall survival post-hatch. The absence of effects corresponded to tissue concentrations
that were less than those observed in deformed turtles from a contaminated site (Bell et al. 2006). In
contrast, eggs of loggerhead turtles buried in nests on a Florida beach were killed by a spill of fuel oil,
although the cause may have been suffocation, not toxicity (reviewed in Rowe et al. 2009).

The unique adaptations and life histories of some species, such as beach nesting by marine turtles, may
create unique sensitivities to oil exposure and effects. A potential cause of mass die-off of marine iguanas
(Amblyrhynchus cristatus) following a diesel spill in the Galapagos Islands may have been due to effects
on algae, their primary diet, or on their unique gut symbionts (Wikelski et al. 2011). In fresh water,
amphibians (frogs, salamanders) that spawn in clear, shallow ponds may be subject to photo-enhanced
toxicity (Section 4.2.5). The toxicity of a diluent recovered from contaminated groundwater in California
to tadpoles of the southern leopard frog (Rana sphencephala), increased by three-fold in the presence of
UV-B at wavelengths and intensities typical of California (Little et al. 1998).

These limited studies reviewed are insufficient to draw firm conclusions about the relative risks of
toxicity to turtles and amphibians exposed to oil. The influence of variables, such as exposure regimes
(single doses vs. repeated doses or continuous exposures), species characteristics (e.g., the nature of the
egg shell and chorion and permeability to hydrocarbon uptake) and the relative toxicity of LMW and
HMW PAHs, remain unknown. Of particular importance in Canada are the interactions between oil
exposure and the natural stressors acting on turtles that are at the northern limit of their range. Extremes
of temperature and short seasons for growth and reproduction could aggravate even minor effects of oil
toxicity.

Recommendation: Research is needed on the toxicity of oil to embryonic and juvenile turtles and
amphibians to determine which species characteristics contribute most to sensitivity to oil and
which exposure routes are most harmful.

4.1.6 Birds

Aquatic birds are at risk of acute mortality following a spill if their feathers become contaminated. Bird
feathers have a function in flight, water repellency and thermoregulation, and all three functions are
significantly impaired by contact with oil. In severe cases, birds are unable to fly, at risk of drowning
because of loss of buoyancy and at risk of hypothermia and starvation due to heat loss, accelerated
metabolism to generate heat and a reduced ability to feed. Issues of thermoregulation are particularly
acute if birds are:

 Oiled during winter months (marine coastal zones, Great Lakes);


 In areas influenced by currents of cold Arctic water (e.g., Gulf of St. Lawrence);
 At higher latitudes (seasonal use of Arctic islands for breeding and nesting); or
 On migration flyways in spring and fall.

Oil spills in areas frequented by aquatic birds often kill hundreds of thousands of birds. For example, the
mean mortality (with 95% confidence limits) of over-wintering birds in the Gulf of Mexico during the
DWH oil spill was estimated as 600,000 birds (320,000 to 1,200,000) using a carcass sampling model,
and 800,000 birds (160,000 to 1,900,000 using an exposure probability model (Haney et al. 2014a, b).
The most affected species were predicted to be laughing gulls (Leucophaeus atricilla, 32% of the
population), royal terns (Thalasseus maximus, 15%), brown pelicans (Pelecanus occidentalis, 12%)
and northern gannet (Morus bassanus, 8%). A potential source of bias in the sampling model was the low
probability of recovering carcasses of oiled birds during coastal surveys. In a stranded bird survey, most
of the 7,229 dead birds collected were found near the coast adjacent to New Orleans and the stricken oil

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 151
platform (Tran et al. 2014), reflecting the source and distribution of oil during the spill. Shorebirds may
be similarly affected and in the Gulf of Mexico it was estimated that 28 species, comprising more than
1,000,000 birds, were at risk of exposure during the DWH spill (Henkel et al. 2012). For the dunlin
(Calidris alpina), a shorebird, 8.6% and 0.6% of live birds showed light to trace oiling in 2010/11 and
2011/12, respectively, indicating that as many as 100,000 birds could have been affected by oil (Henkel et
al. 2014). This number would be even higher without cleanup operations (Beyer et al. 2015).

Similar effects were evident in northern marine ecosystems affected by oil spills, particularly if the spill
coincided with breeding. In Prince William Sound, the estimated number of direct mortalities of seabirds
(with 95% confidence limits) following the EVOS was about 375,000 (300,000 to 645,000), based on the
numbers of oiled carcasses recovered and an intensive study of the proportion of carcasses that
disappeared due to scavengers before being enumerated (Piatt and Anderson 1996). The estimated half-
life of beached carcasses was ≈1 day and systematic bird surveys only began two weeks after the spill
(Ford et al. 1996). Early in the spill, 100% of recovered carcasses were oiled, but the percentage oiled
declined to zero within four months for a wide array of species. While acute mortality of birds, such as
the common murre (Uria aalge; comprised 70% of beached birds), could be attributed to oil exposure,
longer-term effects on abundance were much less certain due to ecosystem-wide changes in abundance
and species composition of prey (fish, invertebrates) in Prince William Sound (Piatt and Anderson 1996).

Local effects of oil spills may also have continent-scale impacts. Species such as the northern gannet,
common loons (Gavia immer) and a variety of shorebirds migrate to and from summer nesting sites in the
Arctic, inland lakes and both coasts of Canada. Mortality during incidents such as the DWH oil spill may
have continent-wide effects on population dynamics and abundance. Monitoring is needed to establish
baseline estimates and trends in abundance and productivity of nesting populations and to assess whether
oil spills in over-wintering areas or along major flyways have affected one or more species. For example,
after the DWH spill, new software tools, such as Google Fusion Tables and Google Maps, were applied to
an existing wildlife database to show whether the distribution of aquatic birds overlapped the area
affected by the spill (Tran et al. 2014).

Sublethal oiling of birds can also affect reproduction. Even a few microlitres of oil transferred from the
parent to the surface of eggs during incubation can be absorbed and cause deformities and embryo
mortality. The median lethal dose (LD50) of weathered MC-252 oil applied to the exterior of mallard
eggs (Anas platyrhynchos) was 0.5 µL or about 0.5 mg oil/egg (Finch et al. 2011). In contrast, dispersed
oil was four to six times less toxic than un-dispersed oil, suggesting that the dispersant mitigated oil
toxicity, perhaps by enabling a more rapid loss of oil from the egg’s surface. However, Corexit® 9500 by
itself at doses of 100 µL/egg caused acute mortality and lower doses reduced hatching success (Finch et
al. 2011).

In areas of Prince William Sound affected by the EVOS, reproductive rates appeared lower for birds that
feed in open water (e.g., kittiwake [Rissa tridactyla]) (Irons 1996) and on shorelines (e.g., black
oystercatcher [Haematopus bachmani]) (Sharp et al. 1996). Relative to un-oiled areas, winter survival
rates of female harlequin ducks (Histrionicus histrionicus) were reduced for up to 10 years following the
spill (reviewed in Ballachey et al. 2014). Effects on pigeon guillemots (Cepphus columba) were less
obvious due to the difficulties in finding well-scattered and concealed nests, but a general decline in
abundance seemed to be exacerbated by the spill (Oakley and Kuletz 1996).

Understanding the impacts of oil spills on aquatic birds is limited by the small number of comprehensive
surveys during oil spills and the difficulties in recovering oiled birds before they disappear. For example,
in busy shipping routes off Newfoundland, the disposal of waste oil and contaminated bilge water causes
a chronic and continuous loss of seabirds, such as thick-billed murres (Uria lomvia), as indicated by
winter surveys of 13 beaches on the Avalon Peninsula (Wiese and Ryan 2003). Although deaths related to
oiling have been increasing, the total number of birds killed could not be estimated from these data. Only

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 152
a small fraction of the Newfoundland shoreline adjacent to shipping routes could be surveyed, and the
proportion of affected birds that did not strand on shore was unknown. However, trends were validated by
significant negative correlations between the percentage of murres oiled and the annual harvest, implying
an impact of mortality on overall abundance.

For freshwater spills, data are not readily available for effects on aquatic birds. To cover damage
assessment needs, Irons (1996) recommended the immediate collection of data on oil exposure,
distribution, abundance and behaviour of birds at spill sites where baseline data is lacking. Dead fish and
birds were observed in shoreline reed beds following the 2005 spill of heavy fuel oil to Wabamun Lake,
AB (Birtwell 2008; Section 4.1.2), but estimates of total numbers have not been published. The ecological
significance of widespread bird mortality from oiling is also poorly understood. Experience with the
DWH spill indicates that widespread mortality of fish-eating birds may have had ‘top-down’ effects on
the marine ecosystem of the Gulf of Mexico, changing the abundance and age structure of fish
populations that supported the birds (Short 2015).

Recommendation: Research is needed to: 1) validate methods for estimating bird numbers and
their susceptibility to oil spills; and 2) establish long-term pre-spill baselines of population
abundance and an understanding of food web functioning and productivity in areas of Canada at
risk of oil spills, particularly where birds over-winter (e.g., the Great Lakes), along major
migratory flyways or in nesting colonies.

4.1.7 Mammals

Aquatic mammals are particularly susceptible to floating oil because they must surface at regular
intervals to breathe. Some species can sense the presence of oil, and captive bottlenose dolphins
(Tursiops truncates) have even been trained to detect and avoid oil (Engelhardt 1983), but field
observations indicate that most species do not consistently avoid spilled oil; within 24 hours of the EVOS,
killer whales were swimming through surface slicks (reviewed by Dupuis and Ucan-Marin 2015). Even if
mammals could detect oil, exposure may be unavoidable in Arctic ecosystems if ice can herd floating oil
into thick layers in breathing holes and leads between ice floes (Chapter 2). When mammals surface
through oil, their skin or fur will be contaminated, and they may inhale liquid oil or oil aerosols during
rapid, high-volume respiration, causing lung damage or narcosis and drowning. For filter-feeding baleen
whales, exposure includes consumption of oil-contaminated plankton and, in extreme cases, baleen may
be clogged by liquid oil (Engelhardt 1983).

Aquatic mammals that rely on fur for insulation face the same issues of acute mortality due to
thermoregulatory failure as do birds. Only a light oiling may lead to serious heat loss and the risk of
death by hypothermia or starvation. All species are not affected equally. For species that do not have fur
(e.g., cetaceans), the significance of direct contamination of skin is less clear and, in fact, oil may not
stick as readily to such species (Engelhardt 1983). Seals accumulate oil on fur, but hypothermia may be
offset somewhat by thick layers of blubber, although there is no information about newborns. Sea otters
(Enhydra lutris) do not have blubber for thermal insulation. If oiled, they must increase their time spent
foraging to compensate for an increased rate of metabolism due to stress and thermal regulation.
However, otters groom compulsively to maintain the water repellency and thermal insulation of their fur
and may starve if time spent grooming increases at the expense of time spent foraging (reviewed by
Sinclair 2012). Some otter populations in Prince William Sound experienced elevated mortality rates for
up to 22 years following the EVOS. Mortality was associated with intermittent exposure to residual oil in
intertidal sediments where otters forage for clams (reviewed in Ballachey et al. 2014).

Hydrocarbons consumed by aquatic mammals through grooming or contaminated diets can be


metabolized and readily excreted, but some is stored in blubber and other fat deposits. Engelhardt (1983)
speculated that stored hydrocarbons may be released into circulation when fat is consumed during periods

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 153
of physiological stress (e.g., low food availability, migration, lactation), and circulating hydrocarbons
may be bioavailable and toxic to the fetus or nursing newborns. Tissue pathology identified by necropsies
of animals recovered after oil spills included:

 Irritation of tissues around the eyes;


 Kidney pathology that may reflect the excretion of hydrocarbon metabolites;
 Hemolysis, anemia and erythropoietic dysfunction;
 Liver and kidney dysfunction, including uremia and dehydration;
 Lesions in the upper digestive and gastro-intestinal tracts; and
 Changes to the adrenal cortex suggesting significant responses to stress and increased kidney
function (Engelhardt 1983).

Mortality of a variety of sea mammals has been frequently associated with marine oil spills. Sea otters,
harbour seals (Phoca vitulina), Stellar sea lions (Eumatopius jubatus), killer whales (Orcinus orca) and
humpback whales (Megaptera novaeangliae) were most affected by the EVOS. Elevated rates of mortality
were evident in the months following the spill in oiled areas compared to un-oiled (Loughlin et al. 1996).
Higher concentrations of hydrocarbons in carcasses of otters, harbour seals and sea lions from oiled
areas strengthened the connection between oil exposure and effects.

As with fish, links between oil exposure and acute and chronic effects remain controversial because most
marine mammals are quite mobile and not easily studied. Mortality may not occur in the immediate
vicinity of an oil spill, particularly if oil causes slow starvation due to chronic respiratory damage,
disruption of thermal regulation or an inability to feed properly. For example, the estimated carcass
recovery rates of cetaceans after the DWH spill were as low as 2% (Williams et al. 2011), which limits
the statistical validity of any proposed cause-effect relationships. Without direct evidence of an exposure
to oil, it may be impossible to discriminate oil toxicity from multiple other causes of morbidity. During
the DWH oil spill, there was a prolonged ‘Unusual Mortality Event’ (UME) of bottlenose dolphin in the
Gulf of Mexico (Carmichael et al. 2012a) that began before the oil spill and persisted for months
afterwards. The UME overlapped in time and space with the spill and dolphins were observed swimming
and feeding in areas of oil-contaminated waters (Schwacke et al. 2014). Although the UME began before
the spill, its prolongation and severity were attributed to oil exposure, implying interacting and delayed
toxicity. Thinning of the adrenal cortex (hypoadrenocorticism, suggestive of chronic stress) combined
with bacterial pneumonia, the primary cause of death, corresponded to previous observations of poor
health in mammals exposed to other oils. Dolphins from oil-exposed areas were five times more likely to
have moderate-to-severe lung disease than dolphins from an un-oiled reference site. Venn-Watson et al.
(2015) strengthened the case for oil toxicity by examining dead and stranded dolphins collected from
oiled and reference sites. They confirmed the high prevalence of bacterial pneumonia and
hypoadrenocorticism, but eliminated algal toxins, morbillivirus infection and brucellosis as contributing
factors. Similarly, Litz et al. (2014) eliminated morbillivirus and brevitoxicosis as causes of the UME.
Although the proposed cause-effect relationship was biologically plausible and the effects corresponded
in timing and location to possible oil exposure, the case for oil toxicity was weakened by a lack of data on
concentrations of hydrocarbons or hydrocarbon metabolites in tissues, urine or bile. Similarly, there were
no measurements of molecular or biochemical markers of hydrocarbon exposure or toxicity, such as
elevated concentrations of CYP1A proteins or DNA adducts (Fossi and Marsili 1997). An analysis of
external wipes from 13 of 35 beached bottlenose dolphin carcasses collected between May 2010 and
November 2012 found residues typical of weathered Macondo oil (Stout 2015), indicating exposure at the
time of the spill and ongoing exposure over the next two years. Low concentrations of hydrocarbons in
digestive-tract materials were not consistent with dietary exposure. These data support the role of oil
exposure in the UME event, but the cause-effect case would be stronger if data on pathology and
contamination were linked for each stranded animal.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 154
During the DWH spill, skin biopsies of sperm whales (Physeter macrocephalus) sampled from the Gulf
of Mexico demonstrated nickel and chromium concentrations two to five times higher than background
(Wise et al. 2014). Concern was expressed for potential genotoxicity, but the forms of nickel and
chromium in MC-252 oil residues were not measured (the valence state determines carcinogenicity).
Overall, the links between tissue metal concentrations and exposure to the DHW spill were not clear,
given that the whales travelled widely throughout the Gulf, there are other sources of metals in the Gulf,
including the Mississippi River, and other measures of oil exposure (tissue concentrations of petroleum
hydrocarbons, biomarkers of oil exposure and effect) were not included.

Recommendation: There is a need for more in-depth research on the accumulation, disposition and
effects of petroleum hydrocarbons and metals in marine mammals, and methods to link biomarkers
of oil exposure and toxicity to specific sources of oil.

The DWH spill may have also affected marine mammal behaviour, as indicated by acoustic recordings of
sperm whales that suggested they had moved away from the spill (Ackleh et al. 2012). Information about
marine mammal behaviour may also be key to mitigating potential impacts on many species following
spills in Canadian waters. For example, the seasonal abundance, migrations and behaviour of humpback
whales in the Bay of Fundy, beluga, fin and blue whales in the Gulf of St. Lawrence, and killer and
humpback whales off the British Columbia coast are well established. This baseline information has
already influenced decisions on major shipping routes in the Bay of Fundy, in this case to avoid
collisions, and on the siting of a major oil terminal in the St. Lawrence River estuary to minimize the
potential exposure of beluga whales to oil spills.

Recommendation: Monitoring is needed in areas of existing and proposed oil-related activities to


ensure that the seasonal and spatial use of habitat by marine mammals is well described and that
risks of oil exposure are minimized.

4.1.8 Ecosystem services

The impacts of oil spills in aquatic ecosystems have been viewed traditionally in terms of oil effects on
individual species, populations or ecosystems. Ecosystem services provide a different perspective on oil
spill effects, focusing on values to human society that are affected, either negatively or positively, by
ecological impacts.

Ecosystem services are “the benefits provided by ecosystems to humans that contribute to making human
life both possible and worth living” (NRC 2013) and are derived from functioning ecosystems, i.e., the
interactions among plants, animals and microbes as primary and secondary producers, among predators,
prey and competitors, and between living species and environmental characteristics such as temperature
and habitat. The physical and chemical properties of an environment also provide services, such as
modifying climate by storing heat and carbon dioxide. Services can be classified into four categories
(NRC 2013):

 Provisioning services (e.g., material goods such as food, feed, fuel and fiber);
 Regulating services (e.g., climate regulation, flood control and water purification);
 Cultural services (e.g., recreational, spiritual and aesthetic services); and
 Supporting services (e.g., nutrient cycling, primary production and soil formation).

Ideally, each service can be valued economically, allowing the damage to aquatic ecosystems from an oil
spill to be expressed monetarily and in ways that are easily understood by non-ecologists. For example,
when access to an ocean beach is blocked by an oil spill for several months during the tourist season, the
loss of income from tourism can be accounted for in the overall cost of cleanup. An analysis of ecosystem

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 155
services can provide a basis for compensating the public for lost opportunities to use the beach and the
businesses or individuals who have suffered financial loss following a reduction in the tourist trade.

This concept may generate a much broader appreciation of ecological, social and economic impacts (i.e.,
on human well-being), the nature of damages included in a post-spill assessment, and the potential
options for remediation and restoration. Its implementation will require a broader and more systematic
framework for structuring programs of research and monitoring, for developing comprehensive models of
how ecosystem services are generated, delivered and affected, and for identifying the trade-offs and
balances in protecting some ecosystem services, potentially at the expense of others (Net Environmental
Benefits Analysis). More importantly, analyses of ecosystem services will define the data that must be
collected to support the approach, the extent and nature of remediation needed to restore services to their
pre-spill values and include the cultural and economic values of affected communities, particularly
indigenous communities, to create perspective on the question of “How clean is clean enough?”

A US National Research Council (NRC) Technical Panel provided a comprehensive review of the 2010
DWH oil spill to determine if the ecosystem services concept provided an additional and useful approach
for assessing resource damage of the oil and of the spill cleanup under the US Natural Resources Damage
Assessment process (NRC 2013). After examining four case studies in detail (wetlands, fisheries, marine
mammals and the deep Gulf of Mexico), they identified three primary obstacles to the comprehensive
application of the concept:

1. A lack of baseline measurement of goods and services produced by the Gulf of Mexico ecosystems
just prior to the harmful event. Only for fisheries of commercially-important species are there
adequate data for an ecosystem services analysis because the mandate of fisheries agencies is to
track abundance, harvest, price and overall economic activity;
2. The technical challenge of developing a comprehensive model to predict the ecological impacts
of an oil spill and associated economic impacts. Such models can be developed from existing
ecological and economic valuation models, but a major obstacle can be a lack of socioeconomic
understanding and data that describe complex human dependencies on natural systems; and
3. The challenges of realistic and accepted economic valuations of ecosystem services, particularly
for services that are not valuated in market activity (e.g., the value of deep sea sediments as a
repository for persistent contaminants from oil spills). If ecosystem services cannot be valued
easily, there is a risk that these services could be discounted or ignored in the decision-making
process (NRC 2013).

The relevance and value of the NRC (2013) report was derived from the legal requirements for Natural
Resources Damage Assessment under the 1990 US Oil Pollution Act (NRC 2013).

Recommendation: To benefit from an ecosystem services approach to assessing impacts, research is


needed to determine if existing regulatory frameworks for spill response, remediation and
monitoring provide an adequate basis for defining and assigning responsibility for its
implementation. Comprehensive ecological and socioeconomic models must be developed in
collaboration with communities at risk, particularly indigenous communities to evaluate the
significance of ecosystem services for marine and freshwater ecosystems.

4.2 Factors Affecting Oil Toxicity to Fish

4.2.1 Species tested

There are considerable differences among species in their sensitivity to oil, some of which relate to their
physiology and life history, some to the environmental characteristics of their habitats (e.g., temperature,
salinity) and some to differences among laboratories in test methods, experimental designs and responses

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 156
measured. Compilations of toxicity data across a wide range of species can illustrate the relative
sensitivity of an individual species (Appendix D, Table D.1), identify the more sensitive species and
define a benchmark for protecting the majority of species.

Barron et al. (2013) provided a useful


perspective on the relative acute lethality of an Species sensitivity distributions (SSDs) - compare
array of crude and fuels oils to aquatic species, the cumulative proportion of species (percentile)
as well as two dispersants tested without oil. affected by oil to a toxicity endpoint measured for
Species sensitivity distributions (SSDs) for 227 each species (e.g., the 96 hour LC50). SSDs display
measures of acute lethality (LC50s) for 67 fish the range of concentrations associated with a
and invertebrate species typical of warm and specific effect and can be used to estimate the
cold waters were expressed as measured concentrations that affect a given percentage of
concentrations of TPHs. Data were selected for tested species. The ‘HC5’, or hazardous
a mix of marine and freshwater species based on concentration harming the 5th percentile of species,
several criteria, including: is often used in ecological risk assessments to
establish toxicity benchmarks that should, in
 All life stages except embryos;
theory, protect 95% of all species from the effect
 Test duration (invertebrates: 48–96 represented by the endpoint.
hours; fish: 96 hours);
 Exposure regime—i.e., gradients of oil
concentrations administered as a continuous flow of (rare) or as static tests with daily solution
renewal (most common);
 Fresh oil only;
 No chemical dispersants; and
 Oil exposure characterized by measured concentrations of TPH.

HC5 values were derived from a curvilinear regression of percentile responses of 10 to 39 species against
log LC50 values (mg/L TPH). Invertebrate species were distributed evenly throughout the SSDs
indicating they were not consistently more or less sensitive than fish species, as expected for a non-
specific narcosis MOA typical of acute lethality. One acknowledged bias in this study was the separation
of species tested by oil type. As might be expected, northern oils (e.g., ANS crude) were tested more
frequently with cold water species, while southern oils (South Louisiana crude) were tested more
frequently with warm water species. However, for inland silversides, a warm water estuarine fish, relative
sensitivity in each SSD was not fixed but varied with the oil tested and the mix of species included
(Barron et al. 2013), suggesting no consistent bias (see also section 4.2.6).

For the chronic toxicity of oil to fish embryos, SSDs were prepared from the data summarized in
Appendix D, Table D.1, selecting only those studies reporting chronic lethal (LC50) and sublethal
(EC50) toxicity values expressed as measured concentrations of hydrocarbons in test solutions (Figure
4.3). Some SSDs are incomplete because there were too few data for chronic LC50s expressed as TPAH
concentrations, and too few data reporting chronic EC50s expressed as TPH concentrations. Nevertheless,
HC5 values calculated from the regressions of Figure 4.3, or estimated by eye, ranged over two orders of
magnitude, from 0.23 to 3.2 µg/L TPAH or 10 to 190 µg/L TPH (Table 4.1), and the SSDs indicated no
consistent differences between freshwater or marine species in sensitivity.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 157
Figure 4.3 Species sensitivity distributions for the sublethal (circles, EC50 values) and lethal (triangles, LC50
values) toxicities of oil, expressed as total polycyclic aromatic hydrocarbons (TPAH; Panel A) and total
petroleum hydrocarbons (TPH; Panel B) to marine (filled symbols) and freshwater species (open symbols). Data
were selected as the lowest EC50 or LC50 values for a given species, so that one species did not dominate the
analysis. Only datasets with 10 or more species were analyzed statistically with a non-linear regression (GraphPad
Prism Ver. 5; GraphPad Software Inc, LaJolla, CA). The horizontal line represents the 5th percentile of species and
intersects the regressions at an EC50 or LC50 equivalent to the ‘HC5’. Data were drawn from papers summarized
in Appendix D, Table D.1 (Kocan et al. 1996; Marty et al. 1997; Carls et al. 1999, 2008; Gulec and Holdway 2000;
Little et al. 2000; Brand et al. 2001; Pollino and Holdway 2002; Couillard et al. 2005; Perkins et al. 2005; Shen et
al. 2010; Olsvik et al. 2011; Greer et al. 2012;; Wu et al. 2012; Gardiner et al. 2013; Incardona et al. 2014; Mager
et al. 2014; Martin et al. 2014; Dussauze et al. 2015; Madison et al. 2015; Sørhus et al. 2015).

The primary limitations of establishing toxicity benchmarks from SSDs are the assumptions that the
species included represent the range in sensitivity of all species in all ecosystems under all environmental
conditions, and that the SSD truly reflects differences in toxicity among species and not the selection of
test species according to their tolerance of lab conditions or differences in test methods among
laboratories. Nevertheless, SSDs provide a reasonable first estimate of toxicity benchmarks, although a
comprehensive comparison among species and oils within one testing laboratory would increase the
confidence in their application to risk assessments.

Recommendation: SSDs should be improved by testing a wider array of test species, with toxicity
expressed as TPAH and TPH concentrations, or as concentrations of specific hydrocarbons
identified as the toxic components of oil.
Table 4.1 The concentrations of TPAHs and TPHs estimated to be hazardous to 5% of test species, as estimated
from species sensitivity distributions of chronic embryo toxicity (Figure 4.3)
HC5 calculated from HC5 estimated from
regression equations data distributions
TPAH – EC50 (µg/L) 0.23
TPAH – LC50 (µg/L) 3.2
TPH – EC50 (mg/L) 0.01
TPH – LC50 (mg/L) 0.19

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 158
4.2.2 Type of oil and oil characteristics

As indicated in Section 4.1, the acute toxicity of crude and refined oils has been attributed to LMW
hydrocarbons that cause lethality by narcosis, regardless of life stage, while MMW 3-5-ringed PAH are
thought to be the primary components that cause chronic toxicity to fish embryos. Based on this
knowledge, it might be expected that the relative risks of acute and chronic toxicity of different oils could
be assessed by simply ranking them in order of the concentrations of their toxic constituents. For
example, very light oils that are rich in monoaromatics and short-chain alkanes should be more acutely
lethal than heavy fuel oils with smaller proportions of these hydrocarbons. While this may generally be
true, the SSDs developed by Barron et al (2013) for acute lethality of crude and fuel oils demonstrated
that estimated HC5 values varied widely among oils, from 0.29 to 3.5 mg/L TPH. Light and heavy fuel
oils were more acutely toxic (0.29 and 0.56 mg/L TPH, respectively) than four crude oils (0.9 to 3.5 mg/L
TPH). The lightest oil (South Louisiana crude) was two- to three-times less toxic than the average for all
oils, based on measured TPH concentrations. No toxicity data for very light oils, such as Bakken crude,
were found.

By the same reasoning, oils that are enriched in alkyl PAHs (e.g., heavy fuel oils) should cause greater
chronic toxicity to fish embryos than oils with low concentrations of alkyl PAHs. Virtually all crude and
refined oils tested to date cause a similar array of toxic responses in fish embryos (Appendix D, Table
D.1), including light and heavy crudes (Wu et al. 2012), refined products, such as diesel (Schein et al.
2009) and HFO (Incardona et al. 2012a; Martin et al. 2014), and bitumen (Colavecchia et al. 2004, 2006)
and products derived from bitumen, such as Orimulsion (Boudreau et al. 2009), and diluted bitumen
(Madison et al. 2015). However, toxicity is highly variable and may be confounded by the relative
proportions of different alkyl PAH in each oil. For example, Martin et al. (2014) measured the toxicity of
a crude and heavy fuel oil to trout embryos expressed as measured concentrations of TPAHs in test
solutions (Table 4.2). If toxicity was determined solely by TPAHs, there should be little difference
between the oils in TPAH concentrations causing embryo toxicity. However, HFO was 4 to 5-times more
toxic than crude oil, perhaps due to higher proportions of the more embryotoxic 3- to 5-ringed alkyl
PAHs, differences in bioavailability of the toxic constituents or the presence of toxic compounds not
typically analyzed in oil, such as polar compounds (e.g., phenolic derivatives of alkyl PAHs, alkyl
dibenzothiophenes; reviewed in Section 4.1.4.3).

It is often difficult to express relationships between PAH concentrations in oil and toxicity quantitatively
because the physical characteristics of oil (e.g., density, viscosity) modify toxicity by controlling droplet
formation and the rate of partitioning of hydrocarbons from oil to water. The proportion of HMW waxes,
resins and asphaltenes will control the physical characteristics of oil (Chapter 2), which in turn affect the
dispersability of oil in water. Thus, relative to light oils, heavy oils are more difficult to disperse
mechanically in water, are less likely to form fine droplets and may appear less toxic than light oils
because their toxic constituents partition more slowly to water.

Weathering of oil changes the relative concentrations of different toxic components of oil, so that toxicity
may change over time depending on what compounds are lost by volatilization and oil-water partitioning
(e.g., Carls et al. 1999). Microbial transformations of hydrocarbons (review, Neff et al. 2013) may also
create hydroxylated derivatives that are more toxic than the parent compound (Hodson et al. 2007b;
Fallahtafti et al. 2012). Thus, the measured toxicity of fresh oil may over- or underestimate impacts of
spilled oil as it weathers (Neff et al 2000). This bias may also occur in lab studies due to the risk of
thermal degradation of hydrocarbons if oil is heated to simulate weathering.

Recommendation: Microbial metabolites in weathered oil and their contribution to toxicity need to
be determined.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 159
Table 4.2 A comparison of the concentrations of TPAHs in two test oils to the concentrations in test solutions
causing chronic toxicity (25 day EC50 values) to trout embryos (Martin et al. 2014).
MESA crude oil HFO 6303
TPAHs in undiluted oil (µg/g) 13,415 42,586
Median effective concentration
(25 day EC50) for sublethal
23 5
embryo toxicity to rainbow trout
(µg TPAH/L)

Recent spills of diluted bitumen (dilbit; e.g., Kalamazoo River, Chapter 8) have raised concerns about
unique threats to aquatic environments because the composition and proportions of hydrocarbons in dilbit
differ considerably from conventional crude oils. As outlined in Chapter 2, dilbit is prepared from
bitumen, a highly-weathered oil depleted in LMW compounds. When bitumen is extracted from the oil
sands, many of the polar and water-soluble components are removed by water washing, including
naphthenic acids (NAs), naturally-occurring, linear and cyclic carboxylic acids found in the acidic
fraction of crude oil and raw bitumen. The NAs account for much of the acute toxicity of oil sands
process water. Because extracted bitumen is too viscous to flow through pipeline, diluents such as natural
gas condensates are added to reduce viscosity, in effect replacing many of the LMW compounds lost
during bitumen formation. The amount of diluent added changes to accommodate seasonal temperature
effects on viscosity and the chemical composition changes with its commercial source. As a consequence,
there may be differences in the chemistry and toxicity of diluent between summer and winter blends, and
among batches.

No publications were found describing the acute lethality of fresh or weathered dilbit to fish. Bitumen
itself is chronically toxic to fish exposed to bituminous sediments, as demonstrated by blue sac disease in
embryos of white sucker (Catostomus commersoni) and fathead minnow (Pimephales promelas) and
endocrine disruption of slimy sculpin (Cottus cognatus) (reviewed in Dew et al. 2015).
Alkylnaphthalenes, the PAH of lowest chronic toxicity, are mostly lost during the formation of bitumen,
so that Access Western Blend (AWB) dilbit contains a lower proportion than conventional crude oils
(Yang et al. 2003). Nevertheless, the concentrations of 3-5-ringed alkyl PAHs in AWB are equivalent to
those of conventional oils, suggesting that chronic toxicity should not differ markedly. There is only one
published report on the chronic embryo toxicity of AWB dilbit (Madison et al. 2015). The EC50 for
Japanese medaka fell in the 37th percentile of an SSD for an array of crude and refined oils (Figure 4.3).
This rank suggests that dilbit may be slightly more embryo-toxic than average, but until there are
equivalent data for other species, any conclusions about the perceived toxicity of dilbit may be
confounded by species differences in sensitivity.

Spills of dilbit also do not suggest any unusual toxicity. Unpublished reports about a 2007 spill of
approximately 100,000 L of dilbit to Burrard Inlet indicated that the primary acute effect was oiling of
aquatic birds, most of which were rescued (Dew et al. 2015). Other effects were due to spill cleanup, i.e.,
removal of oiled Fucus sp followed by mortalities of species, such as barnacles and limpets that use
Fucus beds as a habitat. Hydrocarbon contamination was measurable in some species of crab but not
others, and contamination of sediments and mussels (Mytilus edulis) was still evident four years post-
spill. Other long-term ecological effects were not investigated. The impacts of this spill were relatively
mild, likely because it occurred in a harbour well-equipped to handle spills, the response was rapid and
most dilbit was recovered. Weather and tide conditions also limited the spread of the spilled dilbit and the
spill did not coincide with salmon or bird migrations. However, the sediment contamination indicated that
some dilbit sank, which may account for much of the unrecovered portion.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 160
In fresh water, the 2010 spill of dilbit from a ruptured pipeline contaminated Talmadge Creek and the
Kalamazoo River, MI, caused significant environmental impacts (reviewed in detail Chapter 8). Dilbit
that flowed overland to the creek floated initially, but 10 - 20% sank due to the rapid evaporation of
volatile components and the accumulation of soil particles and suspended particulates in the creek and
river (reviewed by Dew at al. 2015). As with the Burrard Inlet spill, there were direct effects on animals
in the river, including large numbers of oiled turtles, birds and mammals, most of which were treated and
survived. Although there were no immediate fish kills, fish abundance and diversity decreased in oiled
areas and there were clear signs of oil exposure and effects on fish, as indicated by physiological markers
of exposure to petroleum hydrocarbons and poor health. Sediments contaminated with dilbit were toxic to
benthic invertebrates, but toxicity varied somewhat with sediment characteristics (reviewed by Dew et al.
2015). However, the biggest impact was the damage to aquatic and riparian habitats by the spill cleanup
(reviewed in Sections 4.2.7).

Overall, the direct toxic effects of dilbit in both freshwater and marine ecosystems do not seem
dramatically different from those of other oils. However, dilbit does differ from other oils in its behaviour
when spilled. The very rapid loss of volatiles during weathering increases viscosity and reduces the
susceptibility to mechanical or chemical dispersion, which may limit the partitioning of PAHs into water
(Section 4.2.3) and the extent of exposure of aquatic biota.

Recommendation: Research is needed to determine how the behaviour of fresh and weathered
dilbit in various aquatic habitats affects exposure and toxicity to aquatic species.

More generally, the risk of toxicity of an array of oils is not absolute, but depends on viscosity and how it
controls droplet formation and the partitioning of the toxic components of oil into water. Differences in
the resistance to droplet formation can be reduced experimentally by chemical dispersion (Section 4.2.3)
to enable comparisons of toxicity among oils with less bias from differences in viscosity. Thus, to
compare toxicity among oils for ERAs, toxicity test data should be generated by methods that maximize
the surface area of oil exposed to water and the bioavailability of hydrocarbons to test species (reviewed
in section 4.3).

4.2.3 Chemically- vs. mechanically-dispersed oil

Chemical dispersants are used to reduce the amount of oil on the water’s surface and the risks of oil to
surface species, such as seabirds, marine mammals and turtles, and to shoreline marshes, estuaries and
beaches. Dispersion transfers oil into the water column where it can be diluted, dissolved in water, and
degraded by microbes. The downside of dispersion is the risk of toxicity from the dispersants themselves
and increased exposure of aquatic species to the toxic constituents of oil (NRC 2005). Thus, chemical
dispersion may increase the impacts of oil on subsurface aquatic resources by making the components
more bioavailable. For example, PAHs from four different oils were six to 1,100-fold more bioavailable
to juvenile trout from dispersed oil (CEWAF) than from un-dispersed (WAF) (Ramachandran et al.
2004a), depending on the oil tested. Similarly, the toxicity of CEWAF of crude, diesel and heavy fuel oils
to trout embryos was greater than the toxicity of WAF when exposures were expressed as the amount of
oil added to test systems (Schein et al. 2009; Adams et al. 2014a; Martin et al. 2014). Following
dispersion, the ecological risks of spilled oil will increase because the volume of water contaminated with
toxic concentrations of oil may increase by up to 1,100-fold during the period of plume dilution.

When oil is added to water, the rate of partitioning of hydrocarbons from oil to water is largely a function
of droplet size, i.e. the surface-to-volume ratio (Redman 2015). With physical dispersion, oil droplet size
and stability in water varies with oil characteristics such as viscosity. Heavier and highly-weathered oils
are harder to disperse and appear less toxic to aquatic species. Chemical dispersants reduce the median
size of oil droplets and stabilizes droplets in water and can even enhance the dispersion of heavier oils.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 161
Thus, the apparent increase in toxicity following chemical dispersion can be attributed primarily to an
increased bioavailability of the toxic constituents of oil.

When toxicity is expressed as oil loadings, large differences in toxicity between solutions of WAF and
CEWAF have been interpreted as synergistic interactions between the dispersant and oil toxicity (e.g.,
Rico-Martinez et al. 2013). This is a common mistake when the recommended practice of measuring
hydrocarbon concentrations in test solutions is not followed (NRC 2005; Coelho et al. 2013; Adams et al.
2014a). When concentrations of TPHs or TPAHs are measured in test solutions, differences in toxicity
between CEWAF and WAF, among oils, and among CEWAFs prepared with different dispersants,
largely disappear (e.g., NRC 2005; Wu et al. 2012; Incardona et al. 2013; Adams et al. 2014; Dussauze et
al. 2015; Redman and Parkerton 2015). Toxicity is determined by the measured amount of oil in solution,
not by oil loading.

Modern dispersants are less acutely lethal than mechanically-dispersed crude oil (WAF), and far less
toxic than chemically-dispersed oil (CEWAF) (NRC 2005; Hemmer et al. 2011). Nevertheless, adding a
toxic agent to disperse oil that is also toxic might seem counterintuitive.

Under ideal conditions, chemical dispersants are not readily bioavailable to aquatic species. For
example, Corexit® 9500 dissolved in water is toxic to fish embryos, but non-toxic when first mixed with
mineral oil to create chemically-dispersed mineral oil (Adams et al. 2014a). When properly mixed, most
dispersant appeared to be sequestered by the mineral oil and not bioavailable at toxic concentrations.
However, for applications under field conditions, often by airplane, there is no guarantee that all of the
dispersant will land on oil slicks or be properly mixed with oil. Where the dispersant misses the slick or
fails to mix with oil, the potential for dispersant toxicity to aquatic species in subsurface waters is
increased.

In laboratory tests of chronic embryo toxicity, the presence and effect of free dispersant are indicated by
early mortality of eggs, before signs of blue sac disease appear. When mortality occurs later and coincides
with the onset of blue sac disease, it is likely caused by PAHs and not by dispersant (Kuhl et al. 2013;
Adams et al. 2014a). The array of gene transcriptional responses in fish embryos is also somewhat
different between dispersant and dispersed oil (Olsvik et al. 2012), suggesting a direct interaction of
dispersant with transcription. When dispersants are used at oil spills, the unique molecular responses of
fish to dispersants coupled with chemical analyses of un-sequestered dispersant could provide tools to
diagnose specific dispersant effects in the field.

In older fish, dispersants may be endocrine disruptors. Nonylphenol ethoxylates, a common constituent,
can degrade to nonylphenol, which interacts with the estrogen receptor of vertebrates. However, only two
of eight commercially-available dispersants caused weak estrogen-receptor activity in a mammalian cell
line, and no androgen receptor activity, although all dispersants were cytotoxic at high concentrations
(Judson et al. 2010).

Recommendation: Research is needed to determine whether dispersants and chemically-dispersed


oil are endocrine disruptors in sexually-maturing fish in vivo.

The intended benefit of chemical dispersion is a more rapid dilution of oil-contaminated water throughout
the water column, assuming that mixing energy is sufficient. Models of the fate of WAF and CEWAF of
spilled oil along 200 nautical miles of Norwegian coast over a 30-day period showed that the overlap in
distribution between toxic concentrations of oil (0.1 to 1.0 μg/L TPAH) and the eggs and larvae of
northeastern Arctic cod (Boreogadus saida) was reduced somewhat by chemical dispersion (Vikebø et al.
2015). While the results suggest that dispersants could be used to reduce the potential for embryo toxicity,
the model was highly site- and species-specific because it included many assumptions about oil and
egg/embryo distribution. Most importantly, the model did not consider the relationships among

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 162
concentration, exposure time, embryo developmental stage and oil toxicity. Although high concentrations
of oil in water may be transitory following dispersion (i.e., less than 24 - 48 hours) (NRC 2005), only
brief exposures (one to 2.4 hours) of Atlantic herring (Clupea harengus) eggs to realistic concentrations
of CEWAF were sufficient to increase rates of deformities and mortalities at hatch and swim-up, 14 days
later (McIntosh et al. 2010; Greer et al. 2012). Thus, transient exposures of embryos to high
concentrations of oil at a critically-sensitive stage of embryonic development may be highly toxic, even
though the modeling scenario indicated that 30-day average concentrations were sub-toxic.

Recommendation: Research is needed to define and model the interactions among concentration,
exposure time and toxicity for different life stages of fish.

Chemical dispersion of oil also has implications for coral reef species which may be particularly sensitive
because of their shallow habitats and complex species interactions. At spill sites, the exposure of coral
reefs to oil will increase if chemically-dispersed oil is mixed to depths of one to 10 m. One- to five-day
exposures of embryos and larvae of different coral species to CEWAF caused greater toxicity than
exposures to WAF, as indicated by reduced rates of fertilization and survival (reviewed by NRC 2005).
There were also long-term effects of brief exposures, expressed as elevated rates of mortality and reduced
growth and coverage one to two years following a spill. Complete recovery was evident within 10 years.
In lab studies, Corexit® 9500 dispersant alone was about 15 times more toxic to corals than to anemones.
Sublethal exposures to dispersants or to dispersed oil reduced the feeding behaviour of corals and brief
exposures (8 – 24 hours) to CEWAF caused delayed mortality and reduced growth, even though survivors
rapidly depurated accumulated oil (Shafir et al. 2007; Mitchelmore and Baker 2010). Clearly, dispersants
should not be used in the vicinity of coral reefs, a recommendation that corresponds to current practice in
shallow waters.

In Canada, corals on both the Pacific and Atlantic coasts inhabit deeper and colder waters (e.g.,
http://ibis.geog.ubc.ca/biodiversity/efauna/BritishColumbianCorals.html) so that dispersion of surface oil
is less likely to cause their exposure to petroleum hydrocarbons. However, the novel use of dispersants to
manage the discharge of oil from the wellhead during the DWH oil spill demonstrated that benthic
organisms in waters as deep as 1,300 m are at risk of exposure to chemically-dispersed oil.

The NRC (2005) review also questioned the assumption that oil dispersion would protect marine birds by
removing oil from the surface. The protection of birds from surface oil depends on the efficiency of oil
dispersion, but diving birds could be exposed to subsurface plumes of dispersed oil, and presumably this
may extend to marine mammals.

Recommendation: Research is needed to: 1) assess the toxicity of dispersed oil to deepwater corals,
ground fish and invertebrate species that have high economic importance (e.g., lobster, crab,
scallops); 2) model the distribution of deepwater plumes of dispersed oil in relation to areas of
known fisheries productivity, such as the fishing banks of Canada’s east coast or unique habitats
like the sponge reefs off Canada’s west coast; and 3) assess the potential for effects of chemically-
dispersed oil on marine bird and mammals.

Decisions on the use of dispersants following an oil spill use a ‘Net Environmental Benefit Analysis’
(NEBA) to identify the trade-offs between impacts on surface and subsurface resources (NRC 2005).
Although the NEBA concept is logical, the effectiveness of its application is dependent the level of
information available for comparisons of risk between surface versus subsurface resources and offshore
versus shoreline impacts.

Recommendation: To support evidence-based decisions on net environmental benefits, baseline


data are needed on abundance of species that frequent areas at risk of oil spills. Where data are
inadequate, rapid methods are needed for surveying subsurface aquatic resources at the time of a

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 163
spill. The behaviour, fate and effects of dispersed oil should be investigated under field conditions
typical of Canadian waters at risk of oil spills and the toxicity of new chemical agents used for
sorption, solidifying, burning, herding, specialty cleaning and bioremediation of oil (Chapter 6)
(Tamis et al. 2012).

4.2.4 Sediment contamination

The NRC (2005) reviewed a series of studies on experimental chemical dispersion of oil in near-shore
areas of tropical, temperate and Arctic ecosystems. While chemical dispersion reduced the amount of
residual oil on shorelines, the interactions of dispersed oil with sediments were less clear, particularly
with fine-grained sediments that were not the focus of most studies. A recent and very comprehensive
review of the interactions among spilled oil, oil dispersants, the rate of formation of oil-mineral
aggregates and oil precipitation to sediments demonstrated that dispersant use accelerates oil precipitation
to sediments (Gong et al. 2014).

Bottom-dwelling (benthic) organisms will be exposed to dilbit or to oil that has settled into the sediment
or is present chronically due to natural seeps. Extensive lethality has been observed when high quantities
of oil reach the bottom following spills or well blowouts (Teal and Howarth 1984). Sensitive species give
way to opportunistic species in such cases. Oil that has mixed with sediments can persist for very long
periods of time, e.g., 30 years after the Arrow spill in Nova Scotia (Lee et al. 2003b) and 25 years after
the EVOS in Alaska (reviewed by Ballachey et al. 2014). Persistence is due to the slow biodegradation
that takes place under the anoxic conditions characteristic of benthic environments. Organisms in constant
contact with contaminated sediments (infauna and epifauna) are at a higher risk of adverse impacts,
including impaired feeding, growth, development and recruitment, that may alter population dynamics
and community structure. In coastal waters, macroalgae (e.g, kelp) may experience decreased
reproduction, bleaching and mortality upon exposure to oil, and effects on macroalgae reduce habitats and
abundance of other benthic species. As discussed in Section 4.1.4.6, benthic fish species are particularly
susceptible to an increased prevalence of cancer following exposure to carcinogenic and mutagenic PAHs
typical of pyrogenic PAHs, but there have been no long-term studies of the prevalence of cancer in
benthic fish chronically-exposed to spilled oil.

Benthic species may also be subject to hypoxia if there is a high biochemical oxygen demand (BOD) of
sediments due to organic enrichment from an oil spill and oil toxicity to benthic species could be
underestimated if hypoxia is not taken into account. For example, oil from the DWH spill was more toxic
to adult and embryonic sheepshead minnows (Cyprinodon variegatus) when tested under hypoxic
conditions (2.5 mg/L dissolved oxygen) relative to normoxic conditions (> 5.9 mg/L oxygen; > 80%
saturation) (Hedgpeth and Griffitt 2015).

Recommendation: Benchmarks are needed for petrogenic PAHs in marine and freshwater
sediments to guide sediment cleanup following an oil spill.

4.2.5 Photo-enhanced toxicity

The aromatic components of oil (e.g., PAH) may be susceptible to photooxidation due to the interaction
of UV light with double bonds (Chapter 2). The photo-modified products of these reactions include
phenols or epoxides, which are more water-soluble than the parent compounds. Thus, photooxidation of
PAHs in surface waters to more water-soluble derivatives may contribute to their more rapid dilution to
non-toxic concentrations. However, photo-modification may also create hydroxylated derivatives that are
more toxic than the parent compound (e.g., Fallahtafti et al. 2012).

A more important process is photo-enhanced toxicity, i.e., toxicity due to photooxidation in vivo, which
exposes cells to highly ROS, such as singlet oxygen and hydroxy radicals. Although ROS react quickly

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 164
and do not persist long enough in water to cause toxicity, the generation of ROS in cells allows their
immediate reaction with double bonds in lipids (lipid peroxidation), proteins (inactivation of enzymes)
and nucleic acids (mutations). These reactions are reproducible in vitro, and toxicity in vivo can be
predicted from the physical and chemical characteristics of PAHs (Greenberg et al. 1993; Ren et al.
1994). For example, photo-enhanced embryo toxicity is greater for pyrene than for retene (Räsänen, et al.
2011). All species have oxidative defense and DNA repair systems to protect cells from damage by ROS
generated by normal oxidative metabolism, but the rate of ROS production in tissues during photo-
enhanced toxicity far exceeds the capacity to neutralize them. The result is a rapid destruction of cells,
tissue necrosis and mortality of the exposed organism.

The risk of photo-enhanced toxicity will be greatest in the clear waters of the open ocean, in coastal and
freshwater ecosystems where turbidity is low and UV light penetrates to the sediments, and at northern
latitudes with seasonally long days. Species or life stages most at risk are transparent or semi-
transparent (Barron and Ka’aihue 2001), including many zooplankton species (sub-Arctic copepods)
(Duesterloh et al. 2002), crab larvae (Alloy et al. 2015) and the embryos of fish that are un-pigmented at
hatch (e.g., vendace [Coregonus albula]) (Vehniainen et al. 2003). There is little risk to pigmented
embryos (e.g., pike; Esox Lucius) (Hakkinen et al. 2004; Räsänen et al. 2011) or to juvenile and adult fish
(Barron et al. 2005).

For marine species, such as Pacific herring, co-exposure to UV light, including normal sunlight, increased
the toxicity of WAF and CEWAF of crude oil by up to 48-fold (Barron et al. 2003; NRC 2005), as was
the case for freshwater zebrafish embryos exposed to crude oil (ANS, MC-252) and to HFO (Hatlen et al.
2010; Incardona et al. 2012b). WAF alone increased the prevalence of chronic, sublethal cardiotoxicity in
exposed embryos, whereas WAF combined with natural sunlight caused more rapid mortality due to
severe tissue necrosis. Cardiotoxicity was correlated to concentrations of tricyclic PAHs in test oils, but
photo-enhanced toxicity may involve other compounds, such as heterocycles (e.g., carbazoles). UV light
caused a greater increase in toxicity of HFO compared to ANS (Hatlen et al. 2010), likely because of
higher concentrations of TPAHs involved in photo-transformations.

While there are many studies of photo-enhanced toxicity to fish and aquatic invertebrates, the significance
remains controversial because few field studies have confirmed these effects at actual oil spills. Sellin
Jeffries et al. (2013) developed a risk model relating PAH concentrations in surface waters following the
1989 EVOS to the depth of UV-A light penetration and the occurrence of embryos of Pacific herring in
surface waters. They concluded that less than 1% of the herring embryo biomass would have been present
in surface waters containing sufficient PAHs to cause injury by photo-enhanced toxicity. In contrast, field
experiments with caged Pacific herring embryos and passive water samplers suggested that the toxicity of
HFO spilled by the grounding of the Cosco-Busan in San Francisco Bay in November 2007 was enhanced
by natural UV light (Incardona et al. 2011). The passive samplers demonstrated the presence of other
contaminants derived from urbanized watersheds at both oiled and reference sites. Although
hydrocarbons associated with photo-enhanced toxicity were only present at oiled sites, the results do not
preclude interactive toxicity among petroleum hydrocarbons, anthropogenic compounds and natural
stressors, such as temperatures that fluctuate with tidal cycles.

Recommendation: Additional studies are needed to verify that photo-enhanced toxicity occurs at oil
spills, and the extent to which it may interact with other anthropogenic contaminants and natural
stressors.

4.2.6 Arctic vs. temperate spills

The interest in oil resources by Arctic coastal nations and the expansion of oil exploration and
development along the Mackenzie River Valley and in the Beaufort Sea, raise concerns about the impacts
of oil spills in sub-Arctic and Arctic ecosystems, and the toxicity of oil at low temperatures. Recent

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 165
reviews of oil spills under Arctic conditions conclude that impacts will depend on the interactions among
the unique physical characteristics of the environment (e.g., the presence of sea ice), the chemistry, fate
and behaviour of spilled oil at low temperatures (Chapters 2, 3 and 6), and the biology of primary,
secondary and tertiary production by species uniquely adapted to Arctic conditions (Lee et al. 2011;
AOSRT 2014).

The perceived risk of oil toxicity to Arctic species is derived primarily from laboratory toxicity studies,
most focused on fresh oil. Most field studies of spilled oil in the Arctic, concern Arctic waters of Norway
(AOSRT 2014), Chapter 8 and Lee et al. (2011) review the Baffin Island Oil Spill experiment (BIOS) in
detail. The AOSRT (2014) review was concerned primarily with marine conditions, and effects on
freshwater species were not addressed. However, as indicated in Section 4.2.1, there appears to be little
systematic difference between freshwater and marine species in sensitivity to oil exposure, and
presumably this extends to Arctic species, although no studies were found that tested this assumption.

The implications for Arctic mammals are covered briefly in Section 4.1.7 and relate mostly to the
increased potential for exposure to oil when it is confined to breathing holes in ice. For Arctic
invertebrates and fish, effects will depend on the extent of their exposure to oil entrained or dissolved in
water or trapped under ice where species congregate. Under-ice algae that grow in light passing through
surface ice, support communities of organisms, including periphyton grazers, predatory fish and marine
mammals. These under-ice communities would be uniquely sensitive to a subsurface discharge of oil
because buoyant oil would accumulate in high concentrations under the ice, where there may be a risk of
photo-enhanced toxicity (Lee et al. 2011; Section 4.2.5). As AOSRT (2014) concluded, the exposure of
marine Arctic species to oil could correspond to two scenarios: spiked exposures (brief and rapidly
declining exposure; see Section 4.3.1) typical of open ocean spills; and prolonged or chronic exposures
associated with oil trapped under ice.

Other unique conditions in the coastal shelf of the Beaufort Sea that could enhance the impacts of oil
spills have been identified (Carmack and Macdonald, 2002; Lee et al. 2011; ADNR 2014). For example, a
thermal bar formed each summer along the shore west of the MacKenzie River creates a zone of low
salinity water where river and salt water mix. This zone is highly productive and attracts salinity-tolerant
fish species from coastal tributaries. Oil contamination in this zone could have significant impacts on
abundance and productivity of both freshwater and marine fish species, and other species such as beluga
whales that seasonally occupy critical habitat in the near-shore.

Whether Arctic species are uniquely sensitive to oil due to adaptations to life between -1 and 2 °C
remains an open question. Comparisons of toxicity among temperatures will be complicated by the
effects of temperature on the kinetics of partitioning of hydrocarbons from oil to water, the
pharmacokinetics of bioaccumulation, and the activity of enzymes involved in metabolism and excretion
of hydrocarbons. In assessing the risks to fish, consideration must be given to their unique adaptations to
Arctic conditions, such as high tissue concentrations of antifreeze proteins. For example, the unusually
high sensitivity of Arctic cod to oil exposure may be related to a lack of kidney glomeruli to limit the loss
of antifreeze proteins. Excretion of PAH metabolites via the bile instead of the kidneys may prolong their
exposure to PAHs by reducing excretion rates. Because comparative data on hydrocarbon metabolism
and toxicity are sparse, tools developed for monitoring oil exposure and effects in temperate species
should be interpreted with caution when assessing the potential impacts of oil on Arctic fish species.

Preliminary studies suggest little difference in oil toxicity between Arctic and temperate species (AOSRT
2014). A comparison of published data on the toxicity of two components of crude oil and 13 crude oils
to an array of marine species across five phyla (annelids, arthropods, chordates, echinoderms and
molluscs) concluded that there was no systematic difference in acute lethality (<8 day exposures) between
temperate and Arctic species (de Hoop et al. 2011). Similarly, biomarker responses of Arctic cod exposed
to the water soluble fraction of crude oil (e.g. induction of cytochrome P450 and antioxidant enzymes;

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 166
DNA damage) were identical in direction and fell within the range of those observed for temperate
species exposed to similar concentrations of oil (Nahrgang et al. 2010). de Hoop et al. (2011) concluded
that data from temperate species could be used for a first approximation of risks of oil spills to Arctic
ecosystems. However, chemically-dispersed oil was less toxic at 2 °C to Arctic species of fish and
invertebrates than un-dispersed (Gardiner et al. 2013), in contrast to greater toxicity of CEWAF to
temperate species at 10 - 28 °C (Section 4.2.3). The greater toxicity of WAF than CEWAF at low
temperatures may be explained by a more prolonged retention of acutely toxic volatile compounds at cold
temperatures compared to warm (Perkins et al. 2005). Understanding the effect of temperature on the
relative toxicity of chemically-dispersed oil will require detailed chemical analyses of hydrocarbons in
test solutions and tissues of test organisms.

Overall, the differences in oil spill effects between Arctic and temperate regions appear to be driven more
by factors that affect exposure of organisms to oil than by differences in toxicity of oil. Recent reviews of
oil toxicity in Arctic ecosystems by Olsen et al. (2013) and AOSRT (2014) identified research needs
similar to those of temperate ecosystems.

Recommendation: Research on the effects of oil should take into consideration unique aspects of
Arctic ecosystems by considering: 1) interactions among low temperature, physiological
adaptations of Arctic species to low temperature and the kinetics of hydrocarbon solubility, uptake
and excretion; 2) the habitat, life histories and reproductive capacity of regional species at risk; 3)
chronic or sublethal endpoints related to growth and embryonic development; 4) heavy fuel oils
widely used in Arctic shipping; 5) the efficacy and effects of oil dispersion under Arctic conditions;
6) long-term studies of ecosystem recovery; and 7) inclusion of Arctic communities in the design
and conduct of research.

4.2.7 Marine vs. freshwater spills

The history of oil spills is dominated by major shipping accidents at sea or near marine coastal
ecosystems. Much of what is known about the potential effects of oil spills is derived from detailed
research following incidents, such as the EVOS in Prince William Sound, AK, in 1989, and the more
recent DWH spill in the Gulf of Mexico in 2010. However, as reviewed in Chapter 1, the frequency of
marine shipping accidents has been declining due to improvements in tanker safety programs and
construction standards. In Canada, most recent spills are inland, reflecting the locations where oil is
produced, transported and used.

As the result of an aging pipeline infrastructure, major oil spills have occurred in Russia, such as the
Komi spill that resulted in the release of 100,000 - 350,000 tonnes of crude oil and produced water into a
river near the town of Usinsk (Owens et al. 1999). As media attention and scientific reporting on these
spills have been limited, the public’s attention on oil spills in North America has been focused on marine
oil spills and the large cleanup efforts involved. It is important to examine the level of environmental
impacts, costs and technical challenges that may be encountered following spills to fresh water.

A comparison of the 2010 DWH oil spill from a well blow-out in the Gulf of Mexico and the 2010 spill of
Western Canadian Select dilbit (Table 2.2) to the Kalamazoo River, MI, from a pipeline break highlights
some important differences between marine and freshwater spills (Hodson and Williams 2012; Chapter
8.7). The most obvious difference was one of scale. The DWH spill discharged about 780,000 m3 of crude
oil over a 12-week period, contaminating an area ranging from 6,500 to 180,000 km2. By contrast the
Kalamazoo River spill lasted only17 hours, the volume lost was 235-fold smaller (3,320 m3), and only 60
km of river was oiled. At first glance, the potential for ecological impacts would appear much smaller for
the Kalamazoo River. However, in both cases, there were significant impacts on aquatic and shoreline
ecosystems, significant technical challenges in dispersing or recovering oil, and the cleanup continued for
at least four years due to residual oil present as slicks, tar balls and contaminated sediments.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 167
The challenges of cleaning up a freshwater spill were reflected in a substantial difference in the estimated
costs of cleanup ($50 US/L spilled for the DWH vs. $300/L spilled for the Kalamazoo River) (Hodson
and Williams 2012). The higher ‘per litre’ costs to clean up the Kalamazoo River spill seems
counterintuitive given the difference in size of the affected ecosystems. In marine systems, tidal cycles
nourish broad shoreline and intertidal ecosystems that are rich in species adapted to daily flooding,
exposure to air and input of nutrients. Because of tides, the intertidal area subject to oil contamination,
and the number of species affected, is quite large. In contrast, water level fluctuations in freshwater
systems are much smaller and less frequent, the shoreline ecosystem is narrower and the shoreline species
diversity and density are lower (Sergy and Owens 2011).

Relative to the Gulf of Mexico, the Kalamazoo River is of low volume with a limited dilution capacity
and a limited area for oil to spread and disperse. Dilbit remained in thick layers with less opportunity for
weathering by photolysis and biodegradation and a greater opportunity for exposure of aquatic biota to
toxic hydrocarbons. Although mixing energy from wind or waves was relatively low in the Kalamazoo
River, water depths ranged from only a few centimeters to a few meters, increasing the interactions
between surface water and sediments. The high ratio of shoreline length-to-water surface area meant that
a greater proportion of shoreline was contaminated and seasonal flooding carried oil into forested riparian
lands.

Although the DWH spill was much larger than the Kalamazoo River spill, the affected Gulf of Mexico
open water ecosystem was proportionately larger (Cleveland 2013), ensuring a very high dilution capacity
and opportunity for oil to spread, disperse and degrade. The open sea provided abundant mixing energy
due to wind and waves and an extended time before oil reached shore (four to five weeks) (Cleveland
2013), where impacts were similar in nature to those of the Kalamazoo River. Although the great depth of
the Gulf (1,500 m at the site of the well blowout) limited interactions between surface oil and sediments, a
unique feature of the DWH spill was the physical and chemical dispersion of oil at the wellhead.
Approximately, 75% of discharged oil did not reach the surface, forming a subsurface plume at 1,000 -
1,400 m depth. This plume appears to have contaminated sediments of the Gulf of Mexico over a wide
area (Valentine et al. 2014), but the extent and degree of contamination is still a matter of debate and
research. In the Kalamazoo River, contamination of shallow (< 10 m) sediments was evident as frequent
sheens that resurfaced from the sediments to the relatively calm water surface and considerable effort was
expended in trying to recover sunken oil.

In offshore marine waters, oil usually floats because its specific gravity is less than that of salt water
(Chapter 2) and low turbidity reduces the likelihood of oil-mineral aggregates that may sink and
contaminate sediments. As floating oil spreads, surface slicks thin and weather rapidly, and are subject to
more rapid photolysis and biodegradation (Chapter 2). For the DWH spill, winds and currents kept much
of the surface oil offshore, enabling recovery by skimming, burning of surface oil and surface chemical
dispersion for enhanced dilution and biodegradation. Estimates of the fate of the spilled oil are highly
uncertain, but it is likely that only a relatively small proportion (but still a large volume) of the oil reached
shore, although weathered oil continues to wash ashore as tar balls or oily materials. For the Kalamazoo
River spill, the specific gravity of un-weathered dilbit was close to that of fresh water, increasing the risk
of sinking as it weathered and lost its lighter components. The risk of dilbit sinking was increased by the
formation of oil-mineral aggregates (Figure 2.2) due to contact with soil before entering the Kalamazoo
River and with high concentrations of suspended solids in the water caused by flooding.

Effects of oil on aquatic receptors vary with their relative sensitivity and extent of exposure. No studies
have been published that specifically compare the relative sensitivity of marine and freshwater species,
although publications on toxicity (Figure 4.2, Appendix D, Table D.1) do not suggest any major
differences. However, the bioavailability of PAHs to salinity-tolerant fish species exposed to WAF and
CEWAF decreased with increasing salinity (Ramachandran et al. 2006), suggesting a greater sensitivity
of fish to spilled oil in fresh or low-salinity waters. In marine ecosystems, there is a potential to harm

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 168
more species than in fresh water due to a higher biodiversity. In contrast the potential impacts of spills on
single species may be more severe in freshwater habitats. Avoidance is possible (e.g., in clean tributaries),
but habitats are small and physically constrained and there is less oil dilution. Overall, oil spill effects will
be site- and species-specific, depending on how the biology and ecology of each species interact with the
potential for oil exposure. In both marine and freshwater ecosystems, impacts will vary with seasonal
aggregations of species for spawning, synchronized hatching of eggs, schooling behaviour and habitat
selection.

Oil spill cleanup options (Chapter 7), such as booming and skimming, burning of surface oil and chemical
dispersion, are well established for marine ecosystems and are designed for open water applications. For
offshore spills, delays may be significant before oil is blown onto sensitive shorelines, allowing the
opportunity to plan and prepare a response to protect those shorelines. For the DWH spill, the deepwater
plume of dispersed oil presented new challenges in understanding potential impacts and how to mitigate
them. Although the plume was ‘good news’ in that the oil was apparently ‘gone’, subsequent studies
identified a ‘bathtub ring’ of contamination at depths where the plume encountered sediments (i.e., at
1,500 m) (Valentine et al. 2014). The significance of this contamination and the extent of damage to
benthic ecosystems is still a subject of active research.

At open ocean spills, cleanup efforts are viewed as mostly positive because oil is physically removed from
the surface (skimming, burning, chemical dispersion) to protect surface species and sensitive shorelines.
The large-scale of the open ocean also reduces the risk and overall impacts of smoke from burning oil
and the transient increase in dispersed oil concentrations in subsurface waters. In freshwater ecosystems,
burning and chemical dispersion of oil are not accepted practices because they present toxicity risks to
human populations and to small aquatic ecosystems. While booming and skimming work in some
circumstances, access of equipment to oil spills may be limited by topography (e.g., canyons), distance
(remote wilderness with no road access) or season (spring thawing of ice roads). Strong currents in high-
gradient rivers also reduce the effectiveness of booms and skimmers, particularly during spring floods
when ice and debris are present. The sinking of dilbit in the Kalamazoo River presented challenges for
cleanup technologies and a heightened standard of ‘How clean is clean enough?’ due to its proximity to
rural and urban development. New technologies were developed to capture oil droplets suspended in the
water column (‘snares’) and to refloat oil from contaminated sediments (‘poling’, ‘tillers’, water jets,
etc.), but their overall effectiveness in terms of cost and percentage of oil recovered are not well
established (Chapter 7).

Spill cleanup often causes significant habitat damage, in addition to the physical and chemical effects of
oil. The effects of mechanical cleanup of wetlands and marshes along shorelines differ between marine
and freshwater ecosystems. Root systems in marine marshes are dense and can bear greater weight than
more porous root systems of freshwater marshes (Sergy and Owens 2011), so greater care is needed in
freshwater ecosystems to minimize habitat damage. Shorelines and riparian lands are damaged by the
construction of access roads and stable pads for large equipment, by removal of oiled vegetation and un-
oiled vegetation to access shorelines, the excavation of contaminated soils and the destabilization and
erosion of shorelines, marshlands and river channels following excavation. For the Kalamazoo River
watershed, riparian lands along both sides of Talmadge Creek (site of the spill) were deforested for access
roads and some islands that had been over-washed by oil during flooding were removed completely.
Habitat destruction on this scale requires extensive remediation and the impacts on ecosystem structure,
function and services (e.g., access by locals for recreation) may be evident for decades. Other examples
are reviewed in Chapter 8.

Overall, the scale of oil spills in freshwater ecosystems may be smaller than in marine systems, but the
severity of impacts relative to the amount of oil spilled may be greater due to the smaller scope for
dilution and degradation of oil, the close proximity of shorelines and sediments, the confinement of
aquatic species to the spill sites and high densities of single species during migration or spawning. The

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 169
fate of much of the oil spilled to freshwater ecosystems is often unknown, which raises concerns for
potential interactions of surface oil with sediments. Entrainment of droplet oil into bed sediments by
surface water (hyporheic flows) creates a risk of groundwater contamination and toxicity to early life
stages of aquatic species that live in or spawn in sediments. The technical difficulties of cleanup are
aggravated by limitations on the use of common tools, such as chemical dispersants and burning, the
inaccessibility of some spill sites and the potential for greater damage to sensitive ecosystems than is
caused by the oil itself. The pressures on companies that spilled oil, on agencies responsible for
regulation and on responders cleaning up the oil to justify their actions, to clean up the spill quickly and
to prevent future spills may be that much greater when spills are near urban development.

Recommendation: Research is needed on the behaviour and effects of oil in freshwater ecosystems
to take into account the diverse scales, processes, water movements and biological communities in
such systems.

4.2.8 Interactive and cumulative effects

The effects of oil spills on aquatic ecosystems are often viewed in isolation of other stressors, particularly
if they occur in the open ocean. However, with shipment of oil through Canada’s major ports, waterways
and watersheds, and spills from pipelines in areas of urban and industrial development, there is a
potential for oil spills to waters already affected by contaminants from other activities or legacy
contaminated sites. A few examples include:

 Mining effluents: metals, metalloids (e.g. selenium), mine flotation reagents, acid, suspended
solids;
 Pulp mill effluents: organic contaminants, chemical and biochemical oxygen demand (COD and
BOD);
 Municipal effluents: pharmaceuticals, personal care products, BOD, nutrients;
 Agriculture: pesticides, nutrients, suspended solids, BOD, pathogens (livestock); and
 Wood preservative plants: creosote, copper, chromium, arsenic.

In addition to the potential for interactive toxicity, organic wastes can alter oil fate and behaviour. High
concentrations of dissolved organic material (e.g., pulp mill and municipal effluents) may act as chemical
dispersants, promoting the formation of oil droplets and the solubilization of petroleum hydrocarbons. As
with chemical dispersants, dispersion by organic wastes may remove surface slicks more rapidly than in
pristine waters, diluting oil to non-toxic concentrations and promoting biodegradation. However, in
confined and shallow fresh waters, they would also increase the exposure of aquatic species to toxic
concentrations of oil without the benefit of open water dilution. Similarly, increased turbidity due to soil
erosion (agriculture, mining, forestry) and the discharge of organic particulates in effluents may promote
the formation of oil-mineral aggregates and increase contamination of sediments.

Other stressors could include normal events in the life cycle aquatic species. A massive spring die-off of
bluegill sunfish (Lepomis macrochirus) in Lake Winona, MN, followed a winter spill of Bunker C oil
(Fremling 1981). Mortality was delayed until spring spawning, likely due to the combined stresses from
oil exposure, loss of condition over the winter, rising temperatures and spawning.

Natural oil seeps or contaminated sediments that occur in oil-rich areas such as the McMurray oil sands
formation in the Athabasca River watershed, and methane seeps in the Arctic Ocean may also influence
the fate and effects of spilled oil. Natural sources of hydrocarbons create a potential for background or
baseline toxicity (e.g., Colavecchia et al. 2004, 2006) and foster microbial communities adapted to
biodegradation of oil. The balance between an increase in the ‘toxic load’ of hydrocarbons and the rate of
their removal by microbial degradation under different environmental conditions will be site-specific.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 170
In summary, the cumulative and interactive effects of oil and other contaminants are not well
defined and are little studied.

Recommendation: Research is needed to determine interactions between other


anthropogenic contaminants and the fate, behaviour and effects of oil in coastal and
freshwater ecosystems, and to develop methods that discriminate between the effects of oil
and other stressors.

4.3 Limitations of Current Toxicity Test Methods with Oil

The quality of ERAs for proposed petroleum developments and EIAs following oil spills depend entirely
on the quality of data used to predict impacts and to incorporate into models. Toxicity data are derived
primarily from laboratory tests that describe the quantitative relationships between oil concentrations
and the responses of exposed organisms. Ideally, toxicity data should be comparable among laboratories.
However, as discussed in Section 4.1.4.1.1, the accuracy and relevance of toxicity benchmarks derived
from toxicity data for a diverse array of species developed by a variety of laboratories depend on a
number of largely untested assumptions, including that:

 Concentrations of petroleum hydrocarbons are known and constant during a toxicity test;
 The relationship between the amount of oil added to a test solution and the concentration of
petroleum hydrocarbons in solution is monotonic;
 All petroleum hydrocarbons are freely dissolved in aqueous test solutions;
 The constituents of oil as a complex mixture remain in constant proportion to each other across
the range of test concentrations and over time; and
 The exposure scenarios in laboratory tests match the scenarios of actual oil spills.

For most published studies of oil toxicity, these assumptions are rarely tested, and when the assumptions
have been tested, they are rarely met due to the low water solubility of hydrocarbons. Redman and
Parkerton (2015) reviewed the implications for the outcome of toxicity tests with crude oil, and provided
recommendations for the optimal test method, cited throughout the following text. This section provides
strong support for improving and standardizing methods for oil toxicity testing and chemical analyses of
test solutions. Standardized tests provide a technically sound basis for comparisons of toxicity among
species, environmental condition, and different sources of oil, but they do not necessarily provide data
that match scenarios of actual spills.

4.3.1 Constant vs. time-varying exposure concentrations

Because the components of oil are hydrophobic, with log Kow values ranging from < 3 to > 6.5, the loss of
hydrocarbons from test solutions is continuous, and test concentrations are not constant. The more water-
soluble LMW components are lost primarily by volatilization and possibly by biodegradation. The less
soluble components are lost by adsorption to the walls of exposure vessels, to the test organisms and to
surface films. As a consequence, in static toxicity tests without daily renewal of test solutions, there is a
continuous decline in concentrations of oil over the course of the experiment. Within 24 to 48 hours,
virtually no hydrocarbons are bioavailable to test species (e.g., alkyl PAH) (Kiparissis et al. 2003). These
exposures are equivalent to ‘spiked exposures’ (NRC 2005), the continuous dilution method
recommended by CROSERF (Chemical Response to Oil Spills Ecological Effects Research Forum; see
below).

In tests of solutions of hydrocarbons that are renewed daily (static daily-renewal protocol), the loss of
hydrocarbons from solution creates a 24-hour saw-tooth pattern of exposure (Kiparissis et al. 2003).
These test conditions are the most commonly reported for oil toxicity tests. However, in interpreting these
tests to develop benchmarks, it is unknown if test organisms respond to the nominal concentration

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 171
(amount added to the test solution), the initial concentration (time = 0 hour), the average concentration
over the duration of the test, or some integral of concentration and time (e.g., Toxicity Index (TI) =
concentration x time) (NRC 2005). The issue is further complicated because the rate of loss of
hydrophobic compounds from test solutions increases with hydrophobicity (Kow), creating a mixture in
which the relative proportions of each constituent are changing constantly. These test conditions create a
great deal of uncertainty in risk assessments of oil toxicity, and in comparing toxicity among different
oils, different species and different environmental conditions.

These problems can be resolved, in part, with continuous-flow exposure systems. These include: oil-
desorption columns (partitioning of hydrocarbons from oil-coated gravel to flowing interstitial water)
(Carls et al. 1999; Martin et al. 2014); re-circulating tanks with pumps to recycle and mix surface oil into
the water of exposure tanks (e.g., Dussauze et al. 2015); and continuous addition of fresh oil to flowing
water, e.g., using jets to mechanically generate reproducible suspensions of oil droplets in exposure
solutions (Nordtug et al. 2011a). Relative to static or daily-renewal protocols, oil-desorption columns
slow the rate of decline in oil concentrations. However, they cannot sustain constant concentrations
because oil-water partitioning depletes the finite amount of hydrocarbons in the column. The continuous
re-circulation of surface oil in toxicity test tanks sustains relatively constant concentrations of
hydrocarbons in water. However, concentrations decline somewhat (Dussauze et al. 2015) due to
weathering or absorption to the apparatus. The most stable exposure system is the continuous generation
of fresh suspensions of oil droplets by oil-water injection (Nordtug et al. 2011a). In tests with haddock
embryos, this system provided a relatively constant exposure regime in terms of TPHs and TPAHs, with
less than a 25% decline in concentrations over an 18-day test (Sørhus et al. 2015). The primary drawback
is the need to dispose of a large amount of oily wastewater. Although methods for generating a
continuous flow of oil-contaminated water are available, they are not yet widely used or characterized
under different salinities, temperatures and flow conditions.

Recommendation: Research is needed to test and validate continuous-flow methods that improve
the stability of exposure concentrations throughout oil toxicity tests.

An alternative way to interpret and apply oil toxicity data developed with time-varying exposures is to
incorporate pharmacokinetic equations describing rates of hydrocarbon accumulation, metabolism and
depuration into models that relate hydrophobicity to tissue doses causing toxicity (Redman and Parkerton
2015) (Section 4.1.4.1). Mathews et al. (2008) tested this approach with the target lipid model (TLM)
developed by diToro et al. (2000). Although the results were promising, data for pharmacokinetics of
individual hydrocarbons were scarce. Although the potential role of hydrocarbon metabolism by
cytochrome P450 enzymes (sections 4.1.4.4 to 4.1.4.6) was acknowledged, metabolism and the toxicity of
metabolites were not included in the model. The TLM modified for pharmacokinetics included the same
untested assumptions about uptake, MoA and toxicity of hydrocarbons as the original (Sections 4.1.4.1
and 4.1.4.2), and the datasets used to develop and validate the model were not designed for that purpose.
The modified model also did not address the effects of environmental (e.g., temperature, salinity) and
biological factors (e.g., species) on rates of PAH uptake, metabolism and depuration.

Recommendation: The uptake and depuration rate constants for a wide array of individual
PAHs and alkyl PAHs should be measured for a variety of species and life stages of fish,
and at different temperatures and salinities.

4.3.2 Exposure time

Acute toxicity tests typically involve 24 to 96 hour exposures and chronic tests may last days, weeks or
months, depending on the response being measured and the life history of the test organism. These tests
were originally developed to regulate specific chemicals and the discharge of industrial effluents and may
not represent exposure conditions at actual oil spills. NRC (2005) criticized 48 to 96 hour tests because

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 172
they would be overly conservative, predicting greater toxicity than would be observed in an oil spill
where the acutely lethal LMW components would volatilize or be biodegraded in less than 24 hours.
However, this critique was written before the DWH oil spill for which there was a continuous discharge
of oil for four months, broadening spill scenarios to include longer exposure times.

Nevertheless, most oil spills are of short duration and information is needed on the long-term effects of
short-term exposures, i.e., delayed or lingering toxicity. Studies of delayed effects are few in number,
although recent studies demonstrated that brief exposures (1 to 2.4 hours) of Atlantic herring embryos to
realistic concentrations of CEWAF of crude oil following egg fertilization caused equivalent effects on
embryo deformities and survival as continuous exposures throughout embryo development (Section
4.2.3). Exposures of fish embryos to low concentrations of oil impaired swimming capacity when the fish
had developed as juveniles, well after the exposure had ceased and despite a lack of obvious toxicity
during the embryo exposure (Section 4.1.4.5). Longer-term delayed effects on pink salmon migration
were also evident years after embryos survived exposure to WAF of ANS crude (Section 4.1.1.4).

If exposures of fish embryos for less than 48 hours are sufficient to impair recruitment, growth, survival
and reproduction of juveniles, the impacts of brief oil exposures may be far greater than anticipated.

Recommendation: Research is needed to develop brief and simplified toxicity testing protocols
calibrated to realistic exposure scenarios and that assess delayed effects on fish embryos of acute
exposures to oil.

4.3.3 Measurements of oil concentrations in toxicity tests

Varying proportions and concentrations of hydrocarbons in test solutions during toxicity tests generate
considerable uncertainty in the estimated toxicity of oil. Uncertainty is even greater if hydrocarbon
concentrations in test solutions are not analyzed. Toxicity expressed in terms of oil loadings greatly
underestimates actual toxicity. Without a detailed chemical analysis of test solutions, the effects of
chemical dispersion on oil toxicity can also be misinterpreted (Section 4.2.3).

Bias in estimates of toxicity of mechanically- and chemically-dispersed oil can be reduced by frequent
analyses of test solutions to define the actual gradient of exposure concentrations, to assess whether
exposures are constant throughout the test, and to establish means and variances for calculating
endpoints (e.g., EC50s). Frequent analyses of test solutions by GC-MS provide a detailed understanding
of test concentrations of oil and of the hydrocarbons associated with toxicity. Unfortunately, they can also
be very costly, which may explain why many papers do not include measured concentrations of oil.
Alternatives include fluorescence spectroscopy that can be calibrated against analyses of TPHs and
TPAHs (Adams et al. 2014b; Martin et al. 2014). Fluorimetry is well-established as a survey tool in field
studies of oil spills (e.g., Kim et al. 2010; Lee et al. 2012). At the concentrations of oil typical of chronic
toxicity, there is a linear correlation between measures of TPAHs by GC-MS and concentrations of TPHs
determined by fluorescence, enabling the expression of endpoints in terms of TPAHs (Martin et al. 2014),
provided that quality assurance and quality control methods are followed.

As a result of the widespread use of dispersant during the DWH spill, there is also a need for inexpensive
methods to characterize the concentrations of the ingredients of dispersants, in toxicity test solutions and
treated surface waters (Place et al. 2010). Dispersant analyses would clarify the relative contributions of
dispersant and oil to the observed toxicity of dispersed oil when dispersants are not completely
sequestered by oil droplets. The active ingredient of Corexit® 9500 (dioctyl sodium sulfosuccinate) has
been measured successfully by liquid chromatography-mass spectrometry (Kujawinski et al. 2011), but
the cost precludes its frequent application for most toxicology research.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 173
Recommendation: Inexpensive analytical methods should be developed for petroleum
hydrocarbons and for chemical dispersants that can detect the toxic components at concentrations
at or below the threshold of toxicity for most species (Section 4.3).

4.3.4 Confounding of risk assessments by oil droplets

The presence of microscopic oil droplets (<100 μm) in test solutions introduces another major source of
uncertainty in estimates of oil toxicity. Uncertainties are due to significant differences among
laboratories in protocols for preparing solutions of oil for toxicity tests, in design of exposure systems
and in the analytical characterization of test solutions (Redman and Parkerton 2015). Chemical analyses
of test solutions typically combine both particulate and dissolved oil because water samples are extracted
with solvent. Each measurement represents the sum of hydrocarbons in droplet and dissolved phases, but
results are often interpreted as if all hydrocarbons were freely dissolved and bioavailable. This bias
applies equally to analyses of water sampled during oil spills. Thus, risk assessments comparing
measured toxicity to measured concentrations of hydrocarbons in surface waters will be confounded by
significant overestimates of oil concentrations in water and underestimates of oil toxicity. The clear
implication for risk and impact assessments is that reliance on published toxicity data, without
consideration of the inherent biases in toxicity publications, creates a high risk of predicting false
positives (effects when none would occur) or false negatives (no effects when they would occur). These
biases reduce the extent to which data can be compared or combined across studies and limit the
database for assessments to a very small number of studies.

For many fish species, embryo toxicity appears to be caused by the dissolved components of oil, not
particulate or droplet oil (e.g., Carls et al. 2008; Olsvik et al. 2011). If true, the role of droplets is simply
to transfer free-phase oil to the subsurface where partitioning from droplets to water contributes
dissolved hydrocarbons (Redman 2015). However, droplets may also be a direct route of oil exposure for
some species but not others. Oil droplets have been observed to adhere to the gills of juvenile trout
exposed to chemically-dispersed oil in fresh water (Ramachandran et al. 2004b) and to the eggs of
Atlantic haddock in salt water, but not to eggs of Atlantic cod (Sørhus et al. 2015). Oil was more
embryotoxic to haddock than to cod (Nordtug et al. 2011b; Sørhus et al. 2015), suggesting little
contribution of droplets to toxicity. These results were consistent with experiments in which filtered (no
droplets) and unfiltered (droplets) of mechanically-dispersed oil caused little difference in the level of
gene expression of cod larvae (Olsvik et al. 2011). Relative to WAF, CEWAF of North Sea oil enhanced
the accumulation and toxicity of oil droplets to filter-feeding invertebrates (Calanus finnmarchicus)
(Nordtug et al. 2015). Other filter feeders (e.g., mussels) accumulate both dissolved and particulate oil but
the relative contribution of each phase is not well-defined and may vary among species.

Recommendation: Research is needed to determine the extent to which oil droplets contribute to oil
exposure, why droplets interact directly with some species but not others, and the mechanisms and
environmental factors that promote droplet interactions with biological membranes.

There are no commonly-accepted methods for separating droplets from dissolved oil in samples collected,
although centrifugation and filtration have been tried. Centrifugation is a challenge due to the small
differences in specific gravity between water and oil. Contact of unpreserved samples with centrifuge
tubes also risks the loss of dissolved hydrocarbons by absorption to the tube. Similarly, concentrations of
dissolved hydrocarbons in water samples may be increased or decreased during filtration because droplets
that accumulate on filters create an increasing volume of liquid phase oil for oil-water partitioning. To
avoid these problems, hydrocarbons in water samples could be extracted and concentrated with passive
samplers instead of solvents (Redman 2015). With solid phase micro extraction, dissolved hydrocarbons
are absorbed onto a solid matrix followed by GC-MS analysis of the solid phase (Parkerton et al. 2000; de
Perre et al. 2014). Solid phase micro extraction reduces the cost, time and potential error of oil analyses,

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 174
but assumes that the sorbents accumulate hydrocarbons only from the dissolved fraction in water and not
from droplet oil.

Droplet bias can be detected by measuring HMW markers of oil, such as phytane, which relatively water
insoluble and should be non-detectable if solutions contain only dissolved hydrocarbons. The extent of
droplet contamination can be estimated from the ratios of TPAH to HMW alkane concentrations in whole
oil and in oil solutions, or by comparing measured concentrations of PAH to their solubility limits
calculated from Kow (Redman et al. 2012, 2014). At the moment, few clear strategies exist in fluorescence
spectrometry for discriminating dissolved from particulate oil in field samples, although measurements of
fluorescence quenching have been used to estimate the interferences by dissolved organic matter in
laboratory-prepared mixtures (de Perre et al. 2014).

Recommendation: Research is needed to optimize the use of solid phase micro extractions for
measuring dissolved hydrocarbons, to separate dissolved from particulate oil in fluorescence
analyses, and to harmonize sampling and analytical methods between field and laboratory studies.

Landrum et al. (2012) reviewed the challenges of testing the toxicity of complex mixtures of organic
chemicals and arrived at similar conclusions, with several useful recommendations to improve the utility
of experimental data in ERAs (Text box). While it may not be practical to apply all of these
recommendations to oil toxicity tests, they should be considered when interpreting and reporting results
of toxicity tests.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 175
Guidance for evaluating the aquatic toxicity of complex mixtures of organic chemicals (Landrum et al. 2012)

Establishing a dose-response relationship:

 Separate all dose-response analyses by endpoint;


 Differentiate constant exposures from pulsed or intermittent exposures;
 Identify potential biotransformation products that may be ultimately responsible for the observed effect(s);
 Measure both external concentration and internal (body residue) dose for mixture components, including
biotransformation products, to determine those at which different toxicity endpoints are observed and to help
identify potential biotransformation products that may be responsible for the observed effect(s);
 If possible, measure body residues for all chemicals in the mixture suspected of contributing to mixture toxicity;
this provides more certainty than measures of external concentrations for actual target site concentrations,
particularly when there are multiple exposure routes;
 If all mixture components are not or cannot be analyzed, transparently document the associated uncertainties so
that surety of cause is not assumed;
 Ensure that measurement endpoints are appropriate to the mechanisms of toxic action of the mixture and its
components so that comparisons are not ‘apples to oranges’. Consider and account for toxicity modifying factors
that can confound apparent conclusions, including environmental conditions of the test, life stages and behaviour
of the test organisms, possible external or internal biotransformation of mixture components, and temporal dilution
or variation in concentration of mixture components in the exposure medium (water, diet) or in tissues of test
organisms;
 Do not assume that correlation proves causation - it does not;
 Consider all data and do not ignore inconsistencies that may well provide critical information to assist in
determining causation;
 Determine causation relative not just to the mixture but to its constituent components, including biotransformation
products - such products can sometimes be more toxic than the parent compounds;
 Conduct confirmation studies, such as bio-effect directed fractionation, or tests with individual or surrogate
mixtures of mixture components, and/or perform modeling based on known mechanisms, such as molar or TU
addition; and
 Discuss any and all uncertainties explicitly and transparently - do not overstate the cause of the toxicity when only
inference has been established by focusing on a portion and/or portions of a mixture.

4.3.5 Realism in toxicity tests

The concern expressed in this review for maintaining constant concentrations of oil during toxicity tests
runs counter to concerns that tests reflect the ‘reality’ of oil spills, when concentrations of oil vary
enormously, both spatially and temporally. The ‘spiked’ CROSERF protocol for preparing and testing
solutions of WAF and CEWAF models the rapid loss of hydrocarbons from the water column at actual
spills by standardizing one decay rate of oil concentrations in test solutions (Singer et al. 2000). The
method was described as more realistic than tests in which concentrations remain relatively constant.
However, given the site- and spill-specific fate and behaviour of oil, standardizing tests to one time-
varying scenario creates the same lack of realism as a constant exposure scenario. Interpreting exposure
concentrations that vary temporally is difficult and presents significant problems for predictive models of
toxicity used in risk assessments that assume constant exposure concentrations (Sections 4.1.4; 4.3.1).
Predictive models treat toxicity as a fixed reference point that can be modified by known environmental
conditions to estimate effects under natural conditions.

Given that results of toxicity tests are strongly influenced by protocols chosen for preparing and testing
oil solutions (Redman and Parkerton 2015), laboratory tests can never be realistic. However, methods
can be standardized to provide data comparable among labs, species and oils. Thus, NRC (2005)
suggested that the CROSERF method be refined to:

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 176
 Standardize mixing energies for WAF and CEWAF;
 Prepare gradients of test concentrations by diluting stock solutions instead of preparing unique
stock solutions for each concentration;
 Prolong the exposure to measurable concentrations of hydrocarbons;
 Replace nominal exposure concentrations with measured concentrations of TPHs and/or TPAHs,
and report the time during the exposure when concentrations are measured;
 Include treatments to test the potential for photo-enhanced toxicity; and
 Substitute open test chambers for sealed test chambers to allow volatiles to evaporate during the
test.

Many of these recommendations have been followed in recent studies. For ERAs, effects under realistic
conditions could be estimated by applying toxicity data to the unique site-specific circumstances of each
potential or actual oil spill.

Recommendation: Research is needed to determine the relationship between toxicity and exposure
time, and the nature of delayed toxicity following acute exposures (Section 4.3.2).

4.4 Research Needs and Recommendations

The following is a synthesis of research needs identified in previous sections. Many involve the
development and standardization of protocols for lab toxicity testing, post-spill field assessments and
monitoring of impacts.

4.4.1 Research Needs

Research is needed:

 At ‘spills of opportunity’ and experimental oil spills to identify effects on aquatic species at the level
of populations, communities and ecosystems due to the acute and long-term toxicity of spilled oil;
 On how the behaviour and fate of fresh and weathered oil in different ecosystems interacts with the
habitat selection and unique life history traits of aquatic species to control species-specific oil
exposure and toxicity. These interactions are particularly important for diluted bitumen (dilbit)
relative to conventional oils, and northern and Arctic ecosystems relative to temperate ecosystems;
 To understand how different oil types (e.g., dilbit) and spill-control agents (e.g., chemical dispersants)
affect aquatic species under different environmental conditions, including interactions with
contaminants from municipal, industrial and agricultural effluents;
 On mechanically- and chemically-dispersed oil to determine how oil droplets affect the exposure and
toxicity of oil to different species and life stages of aquatic organisms. These studies include a need to
develop analytical methods to reliably discriminate droplet oil from dissolved oil. Studies on the
distribution and effects of dispersed oil under exposure conditions similar to those that may be
encountered during response operations (including subsurface injection of dispersants) are of
particular importance;
 On strategies and protocols to recover spilled oil that avoid or minimize habitat damage, methods to
restore damaged aquatic and riparian habitats and models to define the balance between the
environmental costs of natural remediation and oil spill cleanup;
 On methods for finding and measuring spilled oil in different compartments of freshwater and marine
ecosystems, for measuring exposure and toxicity of oil to aquatic species in situ, and for assessing the
structure and function of ecosystems that reflect the extent and time to recovery following oil spills;
 On the contamination of groundwater due to soil contamination and the entrainment of oil droplets
into bed sediments of rivers by surface water - groundwater exchanges;

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 177
 To develop standardized oil toxicity test methods for all classes of aquatic organisms to enable
reliable comparisons of toxicity among oils, species and different environmental conditions (e.g.,
salinity, temperature), including rapid changes in oil concentrations in water; and
 To refine models that predict acute and chronic toxicity to aquatic species, particularly to assess the
uncertainty caused by violating the assumptions underlying these models and to reduce the variance
of model predictions.

4.4.2 Operational/preparedness Needs

Baseline monitoring is needed in areas where oil is shipped or produced to resolve uncertainties about
‘cause and effect’ should a spill occur. Baseline monitoring should include:

 The physical and chemical characteristics of aquatic ecosystems; and


 The ecology, population dynamics, productivity, population demographics, abundance and
seasonal and spatial use of habitat by locally-resident and transient aquatic species.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 178
CHAPTER 5: MODELING OIL SPILLS IN WATER

Abstract

Oil contains thousands of compounds and its properties can change drastically as natural weathering
processes proceed immediately following a spill. For this reason, earlier modeling of oil behaviour and
transport was phenomenological with a heavy reliance on experimental observations. Since the early
1980s, advances in oil spill modeling have been largely focused on oil dispersion (formation of oil
droplets), formation of oil particle aggregates (especially after the Exxon Valdez oil spill in 1989),
emulsification, evaporation and general transport in open waters, as well as other types of ecosystems. As
a consequence of the Deepwater Horizon spill and emerging concerns over the environmental risks
associated with offshore oil and gas activities in frontier regions, more advanced numerical models have
recently emerged to better predict various processes, especially in situations where no direct
measurements could be made, such as in deep water or in the Arctic. Major modeling advances have
emerged in the areas of dispersion, biodegradation, dissolution and oil-in-ice behaviour. To improve our
understanding on the influence of environmental factors on the fate and behaviour of spilled oil, modeling
efforts continue to be needed in the area of oil-in-ice, the prediction of oil dispersion by waves, oil droplet
formation from blowouts, formation of oil particle aggregates and the biodegradation of oil droplets under
various environmental conditions, including temperature, salinity, nutrients and light (to elicit the
significance of photooxidation processes).

Introduction

In modeling, it is often said that a large spill of oil is actually multiple spills. This is because the oil
properties change with time and distance from the source and the tools adopted near the source could be
ineffective 20 km from the source. Also the tools used on day one could become obsolete on day 10
because of oil weathering or transport to a different type of environment (e.g., from open water to
beaches). In addition, oil spills could cover large areas and it is cost prohibitive to model all processes
occurring at the smallest scales. Yet, the impact is localized, dependent on the concentration at a
particular receptor. Earlier modeling efforts of oil spills relied on using simple formulations for oil
weathering that are calibrated to experimental studies. The resulting formulations have been used in
subsequent oil spill models. However, no rigorous field validations were available or forthcoming due to
lack of experimental replication or accurate characterization of the environmental conditions (such as
wave energy, etc.). Improvements on these formulations have taken place in recent years in various
categories, including emulsion, dispersion and biodegradation of oil. However, more work is needed,
including the formation of oil particle aggregates (OPAs), a fraction that was previously not accounted for
in mass balance calculations, which also influences the transport (e.g., sedimentation rates), fate (e.g., oil
biodegradation rates) and effects (e.g., bioavailability) of the residual oil.

The change of oil properties with time is known as ‘weathering’ and it affects the subsequent transport
and behaviour of oil. Chapter 2 discusses these effects in detail. For this reason, this chapter addresses
first the modeling of the fate of oil due to weathering and then the modeling of the transport of oil in
various compartments of the environment: lakes, rivers, shorelines, wetlands, etc. To facilitate
understanding of the information in this chapter, most modeling equations discussed were not included in
the write-up; rather, the equations were explained by discussing the terms in simple English.

5.1 Modeling of Oil Weathering

When oil is spilled into the environment, its physical and chemical properties change over time in a
process known as weathering, which includes the following processes: spreading, evaporation,
dissolution, dispersion, photooxidation, emulsification, OPAs and biodegradation (Figure 5.1). All of
these are discussed in detail in Chapters 2 and 3. Analyzing and quantifying the degree of oil weathering

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 179
(i.e., oil fate) is a necessary step to determining oil transport and persistence in the environment. For this
reason, this chapter starts by addressing oil weathering.

Figure 5.1 Illustration of the fate process of oil spills in water. Image from Water Resources and Modeling Group
of ERM Inc. (http://gemss.com/cosim.html).

5.1.1 Oil evaporation

The modeling of oil evaporation is based on equations originally developed for water evaporation (Sutton
1934; Brustaert 1982), where the exchange of water between the water body and the atmosphere was
modeled using a mass transfer formulation in which the exchange depends on the water vapour pressure
in the atmosphere and on a mass-transfer coefficient that represents the hydrodynamics at the water-
atmosphere interface (Brustaert 1982). Sutton (1934) developed a mathematical expression for water
evaporation having the terms mass transfer coefficient, concentration of the evaporating liquid, wind
speed, size of the water body and a ratio of kinematic viscosity (defined in Chapter 6) of the air to the
molecular diffusivity in air. Sutton’s formula shows that as the wind speed increases, water evaporation
increases, which confirmed empirical observations.

Most of the modeling of oil evaporation relies on using the vapour pressure of various oil components
based on distillation data and average boiling points. Blokker (1964) presented one of the earlier models
under this category. Yang and Wang (1977) developed a model that accounted for the role of the slick
area in affecting evaporation. However, one of the most widely used models for oil evaporation was
developed by Stiver and Mackay (1984). That model included terms for the volume fraction evaporated,
absolute temperature, a ratio called the ‘evaporative exposure’, and constants obtained based on
experimental data. Similar models include those developed by Tkalin (1986) and Hamoda et al. (1989).
The latter correlated the mass transfer coefficient to the API gravity of oil and the salinity of the water
and found that oil evaporation increased slowly proportional to salinity. Bobra (1992) compared the
Stiver and Mackay (1984) model to extensive experimental data of oil evaporation and reported that the
model performed well for the first eight hours, but overestimated long-term evaporation in most cases.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 180
In a series of later studies, Fingas (2011, 2013) challenged basing oil evaporation models on water
evaporation, arguing that the evaporation of oil is not limited by the air boundary layer, but by the
diffusion of oil components to the interface between the oil and the atmosphere. The author also noted
that the vapour pressure of water is much lower than most of the hydrocarbon compounds of interest for
evaporation. Thus, the author concluded that the limiting factor for oil evaporation was in the liquid oil
and not in the air above it. The studies included evaporation experiments of diesel, ASMB (Alberta Sweet
Mixed Blend) oil and water in wind tunnels at different wind speeds and observed that the wind speed
increased the evaporation of water. However, the increase of oil evaporation with wind speed occurred
only when comparing to the no-wind condition. All non-zero wind speeds gave essentially the same oil
evaporation rate. Based on these results, Fingas (2013) developed two different empirical equations that
predict evaporation of oil based on the properties of the oil, one for diesel and one for heavier oils. Both
equations indicate that evaporation decreases with time under given environmental conditions.

A study by King et al. (2013) in a flume tank exposed to the atmosphere found that the density of the
Access Western Blend increased to more than that of fresh water while the density of the Cold Lake
Blend remained lower than that of fresh water. They used a hyperbolic law for estimating the increase in
mass density with time of the two products.

5.1.2 Oil dissolution

In general, dissolution is not a major mechanism for oil spilled onto the water surface, as most of the
components that dissolve have high vapour pressures and thus evaporate quickly. Thus, evaporation and
dissolution for surface oil are competitive
Oil evaporation dominants the mass transfer processes with evaporation the dominant one
rates for surface oils, while for underwater (French‐McCay 2004). However, dissolution
releases and in the aquifer, oil dissolution could become important for oil away from the
becomes important. For underwater releases, atmosphere, such as from underwater releases,
droplet size distribution is the most important namely the Deepwater Horizon (DWH) spill
parameter affecting the oil dissolution rate. (Camilli et al. 2012) or in the subsurface, such as
in the aquifer of Bemidji oil spill site (Essaid et al.
1995; Cozzarelli et al. 2010; Ng et al. 2014). The
compounds that dissolve rapidly and result in
measurable concentrations in the water column are the monoaromatics, also known as BTEX (benzene,
toluene, ethylbenzene and xylenes). Naphthalene, a polycyclic aromatic hydrocarbon (PAH), also has a
high solubility in water, and so are short chain alkanes, such as methane, ethane, propane and butane
(Ryerson et al. 2012).

Mackay and Leinonen (1977) and Mackay et al. (1980) conducted theoretical investigations of mass
transfer at the oil-water interface and concluded that the dominant mechanism of transfer of soluble
hydrocarbons into the water column during an oil spill is by dispersed oil droplets rather than surface
slicks. Thus, droplet size distribution is the dominant factor in determination of the oil dissolution rate in
subsurface oil release. Small droplets (e.g. ~ 100 μm in diameter) ascend slowly through water and could
remain in the water column for days or months. This buoyancy contributes significantly to oil dissolution.
Large droplets (e.g., greater than a few hundred µm in diameter) rise to the water surface rapidly, so
dissolution from those droplets is relatively inconsequential (French‐McCay 2004). Mackay et al. (1980)
showed that dissolution from surface oil is negligible.

The dissolution of oil differs from that of a pure compound due to the competition of high solubility
compounds reaching the oil-water interface and subsequently the water column. In addition, polar
compounds in the oil, such as the resins and asphaltenes, tend to migrate to the oil-water interface.
However, they tend to remain at the interface due to their low solubility in water. Therefore, these

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 181
compounds tend to lower the dissolution rate of the remaining oil components in the droplet into the
water column.

5.1.3 Photooxidation

Photooxidation results from exposure of spilled oil to sunlight, which causes some components in the oil
to oxidize, potentially facilitating their chemical and biological degradation. Yang et al. (2015) studied
the photochemical behaviour of crude oil in a simulated freshwater environment and found that lower
molecular weight alkanes were photooxidized more rapidly than the higher molecular weight compounds.
They also found that PAHs can be photooxidized more readily than alkanes, which is consistent with
prior studies (Garrett et al. 1998; Maki et al. 2001; Prince et al. 2003; D’auria et al. 2009; Radović et al.
2014). Prince et al. (2003) also report that the majority of aromatic fractions were converted to resins and
asphaltenes. However, a more recent study (Aeppli et al. 2012) indicated that photooxidation produces
oxygenated hydrocarbons that might persist in the environment and can produce carbonyl compounds
(compounds with a C=O group such as carboxylic acids, aldehydes, amides, etc.) and alcohols, which are
soluble in water and thus can be carried by water flow (Chapelle 2001).

5.1.4 Oil emulsification

When oil persists on the water surface for hours and days, it tends to incorporate water droplets within it,
forming water-in-oil emulsions. This process is the opposite of oil dispersion where oil droplets are
incorporated into the water column. The water-in-oil mixtures are grouped into four classes: stable, meso-
stable, unstable and entrained water (Fingas and Fieldhouse 2004). The characterization is based on the
persistence of the water-in-oil mixture and also visual observation. Stable emulsions, called mousse
(Chapter 3), tend to be semi-solid material with a reddish-brown (even orange) colour and can persist for
weeks. Meso-stable emulsions tend to be liquid and brownish to black in colour and tend to persist for
several days to a week under laboratory conditions. However, they can persist much longer in the
environment. The remaining two classes are black in colour, and the water content of these water-in-oil
emulsions varies. Stable emulsions contain around 80% water. Meso-stable emulsions initially contain
around 70% water, and the amount of water decreases to around 30% after a week. The unstable and
entrained-water classes contain usually around 10% water by volume (Fingas and Fieldhouse 2004), and
are common in low viscosity oils (viscosity less than 20 cp1) or high viscosity oils (viscosity higher than
1,000 cp, such as weathered diluted bitumen).

Emulsification appears to increase as the resins and asphaltenes in the oil increase. These latter
compounds act as surfactants and thus situate at the oil-water interface (Bobra 1992). Both asphaltenes
and resins comprise polar compounds. Observations (Fingas et al. 1996, 2000, 2002) revealed that the
weight fraction of asphaltenes was typically higher than 7% in a stable emulsion, and that the weight
fraction of asphaltenes and resins was higher than 3% in a meso-stable emulsion. The presence of waxes
(high molecular weight alkanes) was found to also promote emulsification but to a lesser extent than
asphaltenes and resins. Thus, the stability of water-in-oil emulsions seems to depend primarily on the
asphaltene content. However, it is thought that resin molecules, which are smaller than asphaltene
molecules, tend to initially locate at the oil-water interface, causing the formation of the emulsion, but
that asphaltenes subsequently migrate toward the interface and render the emulsion stable. In that sense,
asphaltenes seem to be a stronger surfactant than the resins. Also, in fairly general terms, emulsification is
promoted by the viscosity of the oil (i.e., oils that have moderate to high viscosity tend to emulsify more
than lighter oils). Thus, emulsion formation increases with weathering of the oil, especially due to
evaporation and dissolution. Both processes cause the viscosity of the oil to increase. Therefore,
emulsification models depend on the outcome of the evaporation and dissolution models.

1
Centipoise (cp) is a measurement unit of dynamic viscosity in the cgs (cm-g-sec) system of units.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 182
Traditional modeling of emulsification relied on predicting three parameters: the water content in the
emulsion, the class of emulsion (discussed above) and the resulting viscosity of the emulsion (Fingas and
Fieldhouse 2004; Xie et al. 2007). The water content is important because in a stable emulsion, it is 80%
by weight, which means that an emulsion contains approximately five times the initial volume of the oil.
This poses major challenges for removing and treating the emulsion in the environment. Mackay et al.
(1980) developed an equation to estimate the water content of a water-in-oil emulsion containing two rate
constants for water incorporation and a third rate constant that depends on the coalescing tendency of the
emulsion. Mackay and Zagorski (1982) suggested a stability index, which is calculated from the
asphaltene and wax content in the oil, other oil components and temperature. Thus, for a content of
asphaltene > 7% the stability index is > 1.22, and for the content of asphaltenes and resins > 3% the
stability index declines to 0.67. Values of the stability index between 0.67 and 1.22 represent a meso-
stable emulsion, while values less than 0.67 represent an unstable emulsion. Zeidan et al. (1997) proposed
a chemical composition index to determine the class of emulsion. The model contains terms for weight
fractions of asphaltenes, resins and waxes. The dynamic viscosity of the emulsion can be estimated by the
method of Mooney (1951). Xie et al. (2007) also considered the demulsification of the emulsion and
represented it using an exponential decay curve for meso-stable and unstable emulsions. They noted that a
minimum amount of energy was commonly needed to initiate the emulsion, and they also developed
estimation methods based on the sea state.

In a departure from the literature on emulsification modeling, Fingas and Fieldhouse (2004) argued that
the constituent oil fractions are not sufficient to fully predict the emulsification class. For example,
diluted bitumen contains more than 20% by weight of either asphaltenes or resins, yet it does not form an
emulsion because of its high dynamic viscosity (more than 10,000 cp). They considered an extensive
database of emulsions and developed empirical relations to predict the classes of emulsions, and the
corresponding water content and viscosity.

Emulsification was found to reduce oil evaporation and dissolution. Experiments by Ross and Buist
(1995) revealed that the intensity degree of oil evaporation decreases with increasing water content and
slick thickness, both due to emulsification. Xie et al. (2007) investigated these findings and developed an
empirical model for oil evaporation from an emulsion. The model contained terms for the weight fraction
of water, the oil evaporation from pure oil and the evaporation from the emulsion.

5.1.5 Oil dispersion

For surface oil spills, wave action breaks the oil slick into droplets (i.e., oil dispersion) propelling them
into the water column. An empirical formula was developed by Delvigne and Sweeney (1988) and
Delvigne (1994) for dispersion. The important terms in the formula were the particle size, the particle size
interval, the entrainment rate per unit surface area of oil particles with particle sizes in a specific interval
around the particle size term, an empirical constant dependent on the oil type and weathering state, the
energy dissipated during breaking waves per unit surface, the fraction of surface area covered by oil, and
the fraction of sea surface hit by breaking waves per unit time. The energy dissipated term and the sea
surface fraction of breaking waves per unit time were further modeled using two additional semi-
empirical formulas containing terms for root-mean of wave heights, wind speed and threshold wind speed
for generation of breaking waves.

In another model, Mackay and Leinonen (1977) assumed that the dispersion of oil from the slick was the
sum of two rates: a breaking wave rate and a non-breaking wave rate. This model included terms for the
volume fractions of oil droplets permanently dispersed by non-breaking and breaking waves, the slick
thickness and the fraction per second of slick area forced into the water column by the breaking waves.

Probably the most appealing aspect of the Delvigne and Sweeney (1988) model is that it estimates the
amount of oil entrained based on the wave energy. Another appealing aspect of this model is the ease-of-

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 183
use and low computational cost. However, no validation of the model has been made in field situations
due to the large scale of natural systems and variability in the sea.

Zhao et al. (2014a,b) developed the model VDROP, a comprehensive model that estimates the droplet
size distribution (DSD) based on the properties of the oil and the hydrodynamics of the system. While
most models of its type consider the interfacial tension to be the only force resisting breakup, the VDROP
model allows for situations where the viscosity of the oil resists the breakup of the droplets, which occurs
when the oil is highly viscous and/or when dispersants are used, causing the interfacial tension to decline.
Zhao et al. (2014a) considered that, after each breaking wave event (assumed every two minutes),
droplets smaller than 100 µm were driven into the water column away from the surface. The energy
dissipation rate in the model can be roughly estimated based on the sea state (i.e., amount of breaking
waves).

5.1.6 Oil particle aggregation

As oil approaches coastal regions it tends to interact with the higher load of suspended particles in the
water column to form OPAs, which move differently from oil droplets or particles alone (Figure 5.2).
The suspended particles include mixtures of silt,
clay, detrital organic matter and algae or plankton,
Alternative terms to OPA have been used, such whose concentrations near shorelines are much
as OMA (oil mineral aggregates) and OSA (oil higher than encountered by oil offshore, especially
suspended particulate matter aggregates). The if the oil came from a surface spill. There are two
term OPA was coined by the Boufadel group major types of OPA: 1) oil droplets coated by
while addressing the diluted bitumen spill in the small particles resulting in the so-called Pickering
Kalamazoo River in Michigan 2010. The group emulsion (Le Floch et al. 2002; Frelichowska et al.
argues that existing terms are restrictive as the 2010); and 2) large particles adsorbing oil onto
particulates do not have to be minerals, and the them and trapping oil between them (Stoffyn-Egli
suspension state is not an intrinsic property, and Lee 2003). The first type is the most common
rather depends on the hydrodynamics. and results in a larger volume of trapped oil, and
sometimes can consist of many oil droplets
binding together forming a large OPA (Le Floch et
al. 2002; Omotoso et al. 2002). In this type, the oil droplets of interest are typically larger than 50 µm,
whereas the particles are typically smaller than a few µm.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 184
Figure 5.2 Formation and movement of various types of OPAs in marine systems. The term OPA comprises OSA
and OMA (as developed by the Kenneth Lee group). This image is reprinted from Gong, Y., Zhao, X., Cai, Z.,
O’Reilly, S.E., Hao, X. and D. Zhao. 2014. A review of oil, dispersed oil and sediment interactions in the aquatic
environment: Influence on the fate, transport and remediation of oil spills. Marine Pollution Bulletin 79: 16-33.
Copyright Elsevier (2014).

Of interest in modeling the OPAs is the critical particle concentration, which is the concentration that
results in complete coverage of the oil droplet, observed to be around 30 to 50% of the oil concentration
expressed in the same units (Ajijolaiya et al. 2006). For example, if the oil concentration in the water
column is approximately 100 mg/L, then a sediment concentration of 30 to 50 mg/L would ensure
complete coverage of the oil droplets.

The formation of aggregates depends on the viscosity and adhesive properties of the oil droplets, the
surface area and chemical and physical properties of the particles, and the salinity of the water and other
factors such as the presence of dispersants (Lee 2002; Khelifa et al. 2005). High viscosity oils (dynamic
viscosity > 500 cp) or those with high adhesion (> 50 g/m2) (Jokuty et al. 1995) tend to bind to sediments
rapidly. The efficiency of trapping increases with the hydrophobicity of sediments until it reaches a
maximum and then decreases (Sun et al. 2010). This is because as the particles become more
hydrophobic, they tend to attach more closely to oil. However, if they become too hydrophobic, they tend
to clump together and thus their chance of adhering to oil decreases. In addition, the increase in salinity
enhances OPA formation, and enhancement is high when the salinity increases to > 1,000 mg/kg (Sun et
al. 2010).

Common modeling approaches for the OPAs rely on estimating the trapping efficiency by empirical
expression fitted to experimental data. This is unlike the approach that has been used for modeling oil
dispersion in chemical engineering (Tsouris and Tavlarides 1994) or in oil spill mitigation (Zhao et al.
2014). A justification for the simplification in modeling OPAs is probably due to the fact that, unlike oil
droplet models where only oil droplets interact with each other, an OPA model would require the
interaction of droplets, particles and OPAs, resulting in an excessive number of equations to capture all
interactions. Nevertheless, existing OPA models cannot predict based on the fundamental properties of
the oil or particles, and they constitute a challenge when predicting the behaviour of oil in systems with

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 185
high sediment (or particle) content, such as streams (Fitzpatrick et al. 2015) and shorelines (Owens and
Lee 2003).

5.1.7 Oil biodegradation in open water

Since the biodegradation of oil depends on the oil-water interfacial area (Atlas and Hazen 2011), the
biodegradation of slicks is considered negligible in comparison to that of oil droplets. Therefore, for oil
spills on the water surface, dispersion is a key factor affecting biodegradation. The approach adopted in
NOAA’s Automated Data Inquiry for Oil Spills (ADIOS2) model consists of treating the oil as various
classes (C4-C12 alkanes; 2-3 ring aromatics) to allow for the reduction of the size of oil droplets following
biodegradation (Viveros et al. 2015). This is a simple approach as the ADIOS2 model is intended for
short-term prediction, mainly during the response to an oil spill. A recent work by Yassine et al. (2013)
modeled oil biodegradation using Monod kinetics and a quasi-steady state approximation for the
dissolution of low solubility hydrocarbons in the water column. They reported good agreement with data.
Vilcáez et al. (2013) assumed that the microorganisms covered oil droplets completely. The simulation
results revealed that small oil droplets biodegraded faster due to their larger surface area per unit mass.
Their model suggested that oil droplets biodegraded faster than dissolved oil components due to their
assumption of complete microbial coverage of oil droplets. However, they provided no data to confirm
their assumption.

5.1.8 Oil biodegradation within beaches

When oil reaches the shoreline, it percolates into the subsurface when the tide falls, and after a few days a
residual amount of oil remains entrapped within the sediments by capillary forces, regardless of the action
of the subsequent tides and waves.

Numerical models have been used to predict biodegradation of petroleum hydrocarbons since 1980.
However, most of them aimed at modeling the biodegradation of dissolved hydrocarbons (Borden and
Bedient 1986; Essaid et al. 1995; Clement 1997). In the model of Borden and Bedient (1986), the
biodegradation of dissolved hydrocarbons was assumed to be limited by the availability of oxygen. One-
and two- dimensional models were used to investigate the effect of microbial kinetics and oxygen transfer
on hydrocarbon biodegradation. Essaid et al. (2003) used the U.S Geological Survey (USGS) solute
transport and biodegradation code BIOMOC coupled with the USGS universal inverse modeling code
UCODE to quantify the BTEX dissolution and biodegradation at an aquifer site located in Bemidji, MN.
Essaid et al. (1995) used a two-dimensional model to simulate the sequential aerobic and anaerobic
biodegradation of volatile and non-volatile fractions of dissolved organic carbon. Clement (1997)
developed a modular computer code called RT3D to describe reactive multispecies transport in three-
dimensional groundwater systems with consideration of multiple electron acceptors for hydrocarbon
biodegradation. These studies also accounted for the hydrocarbons in an immobile oil phase by adding oil
sorption under equilibrium (Borden and Bedient 1986; Essaid et al. 1995) or non-equilibrium conditions
(Clement 1997) into their models. The model SEAM3D (Widdowson et al. 2002) was developed to
account for sequential electron acceptors in aquifers contaminated with dissolved hydrocarbons. In this
model, the hydrocarbons used molecular oxygen as an electron acceptor, followed by nitrate, manganese,
iron, sulfate (sulfate reducing), and finally carbon dioxide (methanogenic). SEAM3D was used to
evaluate the potential biodegradation of oil trapped in the sediments of a beach in Grand Isle, LA (OSAT
2011), and it was found that the dissolution of oil was a limiting factor for the migration of oil off site.
However, all of the aforementioned models applied to dissolved hydrocarbon or relied on physical
dissolution from the immobile/oil phase into the mobile phase for biodegradation to occur. Thus, the
biodegradation of the low solubility hydrocarbons, which occurs by microorganisms colonizing the oil-
water interface, cannot be effectively accounted for.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 186
Nicol et al. (1994) presented a framework for one-dimensional modeling of biodegradation of residual
low solubility petroleum hydrocarbons in fully saturated porous media. The biodegradation of residual oil
was limited by electron acceptor (dissolved oxygen) and nutrients (nitrogen and phosphorus), which were
transported through the porous medium by advection and dispersion. El-Kadi (2001) developed a model
for hydrocarbon biodegradation in tidal aquifers considering the impact of the water content and
temperature on microbial activities. Both dissolved and trapped hydrocarbons were considered. The
model assumed that trapped hydrocarbons were immobile in the porous medium and were subject to the
same biodegradation mechanisms as dissolved hydrocarbons. For modeling trapped/residual hydrocarbon
biodegradation, both studies (Nicol et al. 1994; El-Kadi 2001) assumed that the hydrocarbons dissolved in
the aqueous phase by biosurfactants produced by the oil degrading bacteria are directly metabolized by
the bacteria.

The model BIOB (Geng et al. 2013, 2014; Torlapati and Boufadel 2014) was developed to simulate the
biodegradation of low solubility hydrocarbons in porous media. In the model, the rate of hydrocarbon
biodegradation was considered proportional to biomass growth (Rittmann and McCarty 2001;
Tchobanoglous et al. 2002). In order to consider the washout of hydrocarbons in a beach environment, a
hopane-normalized term was added to the equation of hydrocarbon removal following the work of
Venosa et al. (1996). The biomass growth rate is assumed to be limited by the availability of
hydrocarbons and nitrogen, expressed as a dual-Monod formulation (e.g., Molz et al. 1986). In field
environments, the size of the microbial population in a porous medium is constrained by factors, such as
pore space, production of inhibitory metabolites and sloughing of microbial biomass (Weise and
Rheinheimer 1977; Kazunga and Aitken 2000; Rønn et al. 2002). To avoid unreasonably high microbial
populations, Geng et al. (2014) used the formulation of Kindred and Celia (1989) and Schirmer et al.
(2000) introduced a term to the model to account for the decrease in biomass accumulation when the
microbial concentration approaches its maximum value. The BIOB model was calibrated to laboratory
column data reported in Boufadel et al. (1999) and Venosa et al. (2010) and to field data reported in
Venosa et al. (1996). The model was able to reproduce the experimental data of oil, oxygen consumption,
carbon dioxide production and biomass as represented by the Most Probable Number. The BIOB model
was also capable of capturing the impact of nutrient concentration of microbial growth and hydrocarbon
biodegradation.

5.2 Oil Transport

In this section, the transport of oil is addressed in systems of increasing complexity. First the spreading of
oil on the water surface is considered. This is followed by a discussion of the transport of oil on lakes,
where wind and waves are the major conveyance mechanisms. Then, the transport of oil in streams and
rivers is examined, which adds the complexity of the interaction of the main channel with stagnant zones
and with the hyporheic zone (the subsurface surrounding the stream). Finally, the transport of oil in
estuaries is reviewed where vertical salinity gradients could confine the dispersed and dissolved oil to the
upper portion of an estuary.

5.2.1 Spreading

Fay (1971) and Hoult (1972) developed physical arguments for the spreading of oil spilled on a calm sea.
They explained that initially the oil spreads due to gravity because oil rests higher than the water surface,
and afterwards it spreads due to the interfacial tension between oil and water. The water inertia resists the
spreading of the oil in the initial phase of the spill, and the water viscosity becomes the main resisting
force after the initial phase. However, when wind and/or waves are present, the modeling of surface
spreading would need to account for the elongation of the oil along the wind direction (Lehr et al. 1984).
Thus, instead of a circular spill based on the Fay’s analysis, one would obtain an ellipse with the long axis
aligned with the wind direction.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 187
5.2.2 Transport in lakes and sea

The movement of oil on lakes is driven to a large extent by the action of winds because the current
velocity is small. Wind affects the movement of oil on water in three ways:

 Wind can transport oil floating on the water surface without affecting the water beneath it. The speed
of oil movement carried by the wind could be up to 6% of the wind speed and is typically assumed to
be 3 to 4% in some models (ASCE 1996; Barker 2011; Boufadel et al. 2014);
 Wind over a fetch of say 5 km can generate waves, which can alter the trajectories of oil. As the wave
propagates, the water velocity beneath it changes: the water velocity is forward (in the direction of
wave propagation) under the wave crest and backward under the trough. Thus, floating objects, such
as an oil slick on the water surface or oil droplets dispersed in the water column, move forward when
the crest passes and they move backward when the trough passes. However, the forward movement is
greater than the backward movement resulting in a net forward transport known as the Stokes drift
(Dean and Dalrymple 1984). The maximum value of the Stokes drift is at the water surface, and in
general it decreases exponentially with depth. Waves also break up the oil slick into oil droplets,
causing dispersion, and propels them into the water column. The subsequent transport of oil droplets
would depend on the size of oil droplets, as noted by Elliot et al. (1986) and later by Boufadel et al.
(2006, 2007). These studies explained the comet shape of oil spills, wherein a large mass of oil
propagates ahead of a tail. The large oil droplets (which typically hold most of the oil mass) have a
higher buoyancy and thus remain near the water surface, and get advected further forward by the
Stokes drift than smaller oil droplets that tend to be deeper due to diffusion and wave motion; and
 Wind can generate Langmuir cells (Langmuir 1938; Leibovich and Ulrich 1972), which are
circulation cells within the water column perpendicular to the wind propagation direction. They
would transport the oil along windrows and cause it to down-well into the water column. A salient
example of this was the 1993 Braer oil spill off the coast of Shetland, Scotland, where the oil
disappeared off the water surface to reappear hours later, and the movement was attributed to
Langmuir cells (Farmer and Li 1994).

5.2.3 Transport in streams and rivers

The transport of the bulk of spilled oil in a stream or river can be easily predicted based on the stream
discharge and stream morphology, which provide the hydrodynamic properties needed by the model, such
as stream velocity and the turbulent diffusion coefficients (Fischer et al. 1979; Rutherford 1994).
However, a small but non-negligible amount of oil remains entrapped in dead or stagnant zones near the
banks of the river and in the hyporheic zone, which is the subsurface surrounding the stream (Jones and
Mulholland 1999). These zones behave as transient storage zones retaining water and oil from the main
channel and releasing them at a slow rate, thus extending the residence time of an oil spill in the stream.
The transport of oil in rivers could also be affected by the regional groundwater table, and whether the
reach is gaining or losing water (Jones and Mulholland 1999). Field studies revealed that groundwater in
rivers gaining groundwater could confine plumes into the main channel, which would increase their
downstream transport due to the larger velocities near the thalweg2 of streams (Ryan and Boufadel 2006).

Hyporheic flows (i.e., the flow between the main channel and the hyporheic zone 3) are very important
ecologically, as the hyporheic zone provides a spawning ground for fish. Fish embryos may be adversely
affected by hydrocarbons, as hyporheic outflows from oil-contaminated rivers may carry dissolved and
particulate oil into the hyporheic zone. Hyporheic flows also carry oxygen into sediments and remove
metabolic waste products, such as ammonia, due to the metabolism of communities of bacteria, fungi and

2
The thalweg is a line along the stream length connecting the lowest points of the streambed.
3
The hyporheic zone is a region beneath and alongside a stream bed, where mixing occurs between the shallow groundwater
and surface water.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 188
invertebrates that inhabit the hyporheic zone. If down-welling in shallow water transports oil into the
hyporheic zone, particulates may be trapped as free-phase oil in the interstitial spaces between sediment
particles and act as a source of dissolved hydrocarbons from oil-water partitioning (Hodson et al. 2011).
In a field study of a tidal stream abutting Prince William Sound, Carls et al. (2003) reported hyporheic
flows that would carry dissolved hydrocarbons to pink salmon (Oncorhynchus gorbuscha) spawning in
the stream.

Hyporheic flows are generated by two non-exclusive mechanisms: 1) the irregular bed form generates a
pressure gradient along the hyporheic zone, causing stream water to enter the subsurface upstream of a
dune and to exit it at the lee side (Elliott and Brooks 1997; Tonina and Buffington 2007, 2009; Buffington
and Tonina 2009); and 2) the subsurface heterogeneity in a streambed (Ryan and Boufadel 2006, 2007a,b)
can engender hyporheic flows even though the bathymetry is flat (Ryan and Boufadel 2007a,b).
Hyporheic flows are also affected by the groundwater table (Harvey and Fuller 1998) as they increase
under baseflow conditions (i.e., low water table) and decrease under high flood conditions (i.e., high
water table). Thus, an oil spill during baseflow would introduce oil more into the subsurface than an oil
spill during high flows.

In many situations, a better understanding of sediment transport in rivers is needed to understand the
transport of oil and OPAs. Sediment transport in rivers occurs through three complementary mechanisms:

 Wash load consists of very fine particles that are not present in abundant concentrations on the bed.
Examples include suspended organic matter. Therefore, knowledge of bed material composition does
not allow one to predict wash load transport very well;
 Suspended load is the load that is suspended in the water column but still interacts with the bed. The
interaction usually occurs at ripples causing the shear stress to increase and sediments to be
suspended from the lee side of the ripples; and
 Bed load has continuous contact with the bed and has a direct relation to the bottom stress in rivers.
Both suspended and bed loads are commonly predicted using sediment transport models. Thus, to
predict the formation of OPAs, sediment transport models might not provide a complete description
of the transport, as they focus on the suspended load and bed load, while the formation of OPAs
depends also on the wash load.

The types of OPAs greatly affect their transport. It is worth noting that the transport of OPAs is different
from the transport of sediments for two key reasons:

 The OPAs do not in general have the same shape as oil droplets, and their density could vary from
that of oil to a density much greater than that of water (due to sediments). Thus, they move differently
from both the oil droplets and the sediment particles; and
 OPAs can form based on interaction with the wash load of rivers and not with the suspended and bed
loads. Therefore, traditional sediment transport, which predicts only suspended and bed loads, might
not be able to capture the dynamics of OPA transport when the OPA is generated based on interaction
of oil with the wash loads.

5.2.4 Transport in estuaries

Flow in estuaries is characterized by reversals in flow direction due to tides and by the presence of a
freshwater-saltwater interface. Large estuaries could have waves and could be strongly affected by wind
and by the Coriolis force, which would tend to push the water to one of the banks of the estuary (Fischer
et al. 1979). Estuaries tend to have a salt water toe overlain by fresh water exiting to sea. Therefore, for
oil spilled inland in the estuary, dispersed or dissolved oil moving seaward would upwell to the water
surface when encountering the saltwater toe (Short 2015). Thus, previously dispersed oil could re-
coalesce again. The interaction of fresh and saline waters in estuaries is important not only for physical

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 189
transport of spilled oil, but also because mixing of these waters often results in the combination and
settling of clay particles due to the suppression of the double layer when fresh water approaches high
salinity waters (Clark 1996).

5.2.5 Transport of oil within beaches

Transport of oil within beaches is not commonly modeled, as the free-phase oil washes out fast, and one
ends up with residual amounts of oil trapped by capillarity and by organic content in sediments (Brost et
al. 2000). However, sand beaches contain little organic material. Neglecting the organic content in
sediments, Etkin et al. (2008) developed a holding capacity of beaches based on the beach sediment type
and the tide range using the subsurface flow model MARUN (Boufadel et al. 1999).

The persistence of oil in a beach depends on the type of beach sediments and the beach groundwater
table. While addressing the Exxon Valdez oil spill (EVOS), Li and Boufadel (2010) found that oil
persisted at locations where the groundwater table drops into a lower sediment layer of low permeability,
causing the entrapment of oil. Beach hydraulics was also found to affect the oxygen concentration in the
oiled regions of the beach resulting in low oxygen concentrations (Boufadel et al. 2010) suspected to
cause the persistence of the Exxon Valdez oil spilled in Prince William Sound. The mechanism is
illustrated in Figure 5.3. The beach receives dissolved oxygen from the inland groundwater table and
from the seawater during the high tide. As the tide ebbs, the oxygen propagates into the beach and gets
consumed by abiotic and biotic reactions. For the EVOS, it was found that the oxygen concentration at
the oiled region is around 1.0 mg/L resulting in anoxic conditions, which prevented aerobic oil
biodegradation.

Figure 5.3 Beach showing the propagation of oxygen (blue arrows) through the beaches of Prince William
Sound, AK. Reprinted with permission from Boufadel, M.C., Sharifi, Y., Van Aken, B., Wrenn, B.A. and K. Lee.
2010. Nutrient and oxygen concentrations within the sediments of an Alaskan beach polluted with the Exxon Valdez
oil spill. Environmental Science & Technology 44: 7418-7424. Copyright (2010) American Chemical Society.

The role of nutrients in oil biodegradation has been noted since the 1970s (Atlas and Bartha 1972). But it
gained a large visibility following the EVOS (Prince 1993; Bragg et al. 1994), where more than 50 tons of
neat (i.e., solid) nutrients were dissolved and applied onto the beach surfaces. Wrenn et al. (1997)
conducted tracer studies on a Delaware beach and found that the bulk of the nutrient plume persists in the
beach for about 24 hours and thus re-application would be needed. This is understandable noting the
hydraulics within the beach, as illustrated by Figure 5.3.

Tracer studies in Prince William Sound beaches were conducted by applying tracers on the beach surface
at low tide (Li and Boufadel 2011) and by releasing them deep into the beach (Boufadel and Bobo 2011;
Boufadel et al. 2011). In all cases, it was found that the applied solution moves slightly downward and
landward during rising tides, and seaward during falling tides, confirming the numerical studies on these
beaches (Guo et al. 2010; Xia et al. 2010).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 190
More recently, Geng et al. (2015) noted that sand beaches in the Gulf of Mexico could have different
limiting factors depending on the location within the beach: nutrients could be limiting in the shallow
zone of the upper intertidal zone, while oxygen could be limiting in the deep area of the mid-intertidal
zone.

Bioremediation has been considered one of the major non-intrusive technique for mitigating the impact of
oil on shoreline ecology (Pritchard et al. 1992). The addition of nutrients is known as biostimulation
(discussed in greater detail in Chapter 6), while the addition of microorganisms is known as
bioaugmentation (Zhu et al. 2001). Prince et al. (1994) reported high degrees of success when nutrients
were added to oil polluted beaches to remediate the EVOS. Venosa et al. (1996) conducted a
bioremediation study on an experimental Delaware beach intentionally treated with oil in quintuplicate
plots, and found that addition of nutrients to concentrations larger than 2.0 mg/L in the interstitial pore
water enhanced oil biodegradation several-fold. However, they noted that the addition of a microbial
culture grown on site in drums exerted no further enhancement beyond simple nutrient addition. Later, Li
et al. (2007) developed a nutrient delivery technique that applies to most beaches, which consists of
applying the nutrient solution onto the beach surface following the falling high tide and to rely on the
beach water table to carry the applied solution to the oil-contaminated zone.

5.2.6 Floodplains

An oil spill that occurs during a flood would result in oil deposition onto the floodplain, whose hydraulics
is more or less disconnected from that of the river under normal conditions. This causes the oil to deposit
in patches that are disconnected, especially if the spill occurred during high vegetation seasons.
Therefore, the fate of oil in the floodplain could be greatly affected by photooxidation, some of whose
products are soluble in water. The biodegradation of oil in the floodplain could also occur, provided the
sediments are sufficiently coarse to allow for oxygen replenishment from the atmosphere and also that the
nutrient concentrations are not limiting.

5.2.7 Wetlands

In general, oil does not penetrate much into wetlands due to the fine texture of wetland sediment.
However, channels and macropores due to bioturbation would constitute pathways to deeper penetration
of oil. In addition, oil tends to adsorb to plants and plant roots. But in general the bulk of the oil gets
trapped at the edge of the wetlands through adsorption to vegetation.

5.2.8 Underwater release

The release of oil under water due to a pipe rupture or well blowout causes the formation of a jet where
the spewing oil moves in the water column due to the momentum. Eventually, the buoyancy of the oil
starts to play a role, and the oil released at and beyond that location forms a plume. This is what happened
in the DWH. However, for practical purposes one could refer to the oil release as a jet/plume. Upon rising
through the water column, the vertical plume tends to get bent due to horizontal currents commonly
present in waterbodies. Models that simulate such releases include the CDOG model (Zheng et al. 2003)
and the Deepblow (Johansen 2000).

Jets/plumes generate a wide range of droplets whose transport and fate depends on the size. Small
droplets tend to remain within the plume even when it is bent by horizontal currents. However, large
droplets rise rapidly and could separate from the plume bent by horizontal currents (Figures 5.1 and 5.2).
Indeed, that was observed in the DWH spill, where the transport time of large droplets (size of
millimeters) to the surface was on the order of hours, while that of small droplets (< 70 µm) was on the
order of months (Kujawinski et al. 2011). For this reason, the DSD from a pipe rupture is of great interest
in terms of predicting where the droplets and plume will move. Unfortunately, the models CDOG and

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 191
DEEPBLOW do not generate the DSD; rather, they require it as input. Johansen et al. (2013) developed a
formula for predicting the D50 (median diameter) from an underwater blowout based on the orifice
diameter and the properties of the oil. Their model was calibrated to experimental data from vertical jets
whose orifice was 1 to 5 mm (Brandvik et al. 2013), and one case where the orifice diameter was 12 cm
(Johansen et al. 2001). Using their model to obtain the D50, the DSD would be obtained in the aftermath
by assuming either a Rosin Ramler distribution or a lognormal distribution. The VDROP model combined
with jet correlations resulted in the model VDROP-J, which provides the droplet size distribution from
jets/plumes (Figure 5.4a). The model was used to predict the DSD from the DWH in Zhao et al. (2015)
(Figure 5.4b). Zhao et al. (2014b) argued that there is no theoretical reason for the existence of a
unimodal DSD, as required when using the Rosin Ramler or the lognormal distribution.

(a) (b)

Figure 5.4 VDROP-J model: a) illustration of the fluid parcel moving along the jet trajectory, where VDROP
model is running for each fluid parcel location to obtain DSD along the plume. Of interest are the following
quantities: The number of droplets per unit volume “n”, the stream velocity “u”, the holdup φ (ratio of oil mass to
total mass), and the energy dissipation rate ε, which vary as the oil moves away from the orifice; b) simulation
results for DWH blowout4.

5.2.9 Ice

Lee et al. (2011) provided a thorough review of the behaviour of oil in ice-infested waters in the Arctic.
The findings of that report are summarized herein.

An obvious effect ice has on oil is to decrease the oil temperature causing the viscosity of the oil to
increase. However, weathering processes (addressed above) could still occur. In particular, for oil pooled
between blocks of ice, evaporation becomes the dominant weathering mechanism. Also high viscosity
causes the oil to spread less on water and to become less amenable to physical or chemical dispersion
(Fingas and Hollebone 2003). Oil spilled directly on ice would tend to spread less than on water as the ice
roughness would resist the spreading more than water (ITG 1983).

4
Images reprinted from:
a) Zhao, L., Torlapati, J., Boufadel, M.C., King, T., Robinson, B. and K. Lee. 2014b. VDROP: A comprehensive model for droplet
formation of oils and gases in liquids-Incorporation of the interfacial tension and droplet viscosity. Chemical
Engineering Journal 253: 93-106. Copyright Elsevier (2014).
b) Zhao, L., Boufadel, M. C., Adams, E.E., Socolofsky, S., King, T. and K. Lee. 2015. Simulation of scenarios of oil droplet
formation from the Deepwater Horizon blowout. Marine Pollution Bulletin. Accepted. Copyright Elsevier (2015).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 192
Oil spilled on ice can become encapsulated within the ice structure in winter conditions when new ice is
being formed. In some cases, this process can happen rapidly (within 18 to 72 hours). In salt water, the
formation of ice is accompanied by the release of salt, resulting in ‘brine channels’ and ‘brine pools’ (see
Figure 2.10), which contain highly concentrated solutions of sea salt. The channels provide pathways for
oil migration into the water beneath the ice, and the pools provide mini-reservoirs for oil accumulation
between the ice blocks (Bobra and Fingas 1986).

In terms of transport, studies have shown that as long as the ice coverage is less than 30% of the water
surface, the oil behaves more or less independently from ice. However, as the ice coverage exceeds 30%,
the oil is found to drift with ice (Ross and Dickins 1987; Venkatesh et al. 1990). In that case, one would
need to use ice models to track the movement of oil. Recently, major efforts have been placed on remote
sensing, namely using satellites to detect the presence and thickness of oil under the ice (Dickins et al.
2008, 2011).

5.3 Research Recommendations in Priority Order

 Better models need to be developed, using basic properties of waves, for the prediction of oil
dispersion by waves, especially for high viscosity oils.
 Models for blowouts need to generate the oil droplet size distribution (DSD) and not only the
underwater trajectory of oil. These would need to account for the effect of gas on the release.
Currently, the only model that can generate the DSD is the model VDROP-J (Zhao et al. 2014b),
which covers only the near-field.
 Predictive models are needed for the formation and fate of OPAs in freshwater and saltwater
environments. Existing models are phenomenological; they reduce the data but cannot make
predictions.
 Models for the biodegradation of oil droplets need to be validated by comparison with data
obtained from controlled experiments, especially when dispersants are used (Venosa and Holder
2007).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 193
CHAPTER 6: REVIEW OF SPILL RESPONSE OPTIONS

Abstract

Chapter 6 discusses the evolving nature of oil spill response. The implementation and effectiveness of the
various oil spill response options available are influenced by a variety of factors, such as oil types and
properties (presented in Chapter 2), environmental and ecological conditions in the vicinity of the spill
site (e.g., weather, wave, wind, daylight, ice conditions, visibility and sensitivity and distribution of
species) (introduced in Chapter 3), technical, logistical and financial factors (such as responders’
knowledge and skills, availability of personnel and equipment, time constraints, regulatory approvals,
health and safety criteria, cost and economic impacts), and other constraints (e.g., community
engagement). Some of those conditions and factors are temporally dynamic and interactive and should be
comprehensively considered in oil spill contingency planning and response decision-making.

This chapter is divided into four main topical areas, namely: 1) natural processes; 2) physical response
methods; 3) biological and chemical methods; and 4) factors affecting spill response and cleanup
effectiveness. Under each main section, up to eight subsections present in detail the strengths and
weaknesses of each response strategy to provide the reader with a clear understanding of how to
implement each method based on influencing factors, such as how each method is affected by the
environment and the type of oil being removed, and how each method is advocated or opposed for use
based on its known effectiveness in different environments or ecosystems (depending on factors such as
oil types, weather, wave impingement, ice conditions, daylight and visibility, ecological considerations
and technical and economic conditions). Section 6.1 describes natural processes, such as natural
attenuation, evaporation and photooxidation. Section 6.2 presents descriptions of physical cleanup
processes, including containment and recovery, sorption, shoreline types, vegetation cutting, removal of
oiled sediment, oiled sediment reworking, physical dispersion (influence of mineral fines) and in situ
burning. Section 6.3 is concerned with biological and chemical strategies, such as bioremediation,
phytoremediation, chemical dispersion, surface washing agents, solidifiers, herding agents and debris and
detritus removal. Section 6.4 presents a discussion of the factors that affect spill response and cleanup
effectiveness, such as oil types and properties, environmental and ecological factors (weather, wave
height, ice conditions, daylight and ecological factors, including fish, invertebrate and wildlife mortality),
and technical and economic factors impacting effectiveness.

Research recommendations for spill response needs and knowledge gaps are identified in priority order in
Section 6.5. The underlying theme of the recommendations is the need to conduct controlled field studies
to advance spill response strategies, especially for subsurface blowouts, Arctic oil spills and freshwater
shorelines. Most of what we know about oil spill response technologies has been developed from
laboratory studies, mesocosm test systems and case studies (with limited controls and treatment
replication). The stage for these studies should be primarily the Arctic, which has already been put forth
in the needs from Chapter 3, because that is where our spill response knowledge is most lacking. We need
to develop sufficient knowledge in understanding the fate, behaviour and impacts of different types of oils
(in varying weathered and treatment states) in snow or ice conditions. Controlled field studies will help
close this gap without significant negative impact on the environment if done properly. Furthermore, there
is a need to improve our understanding of what to do about anaerobic biodegradation in sediments and to
advance emerging technologies like bioventing, air sparging and/or dissolved air flotation to incorporate
oxygen temporarily into the anaerobic zone of sediments and stimulate aerobic biodegradation of
contaminating hydrocarbons. The results of such studies may aid in the cleanup of benthic environments
contaminated by sunken oil from accidental spills or deep sea blowouts. Finally, regarding dispersant use,
although some researchers have concluded that deep sea dispersant applications were successful during
the Deepwater Horizon spill, it remains unclear to what extent the oil was dispersed physically due to the
high velocity of oil emerging from the well into the water or chemically due to mixing with applied
dispersant. This question needs to be answered by controlled experiments to reduce the uncertainty of this

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 194
treatment strategy. Questions about the persistence of dispersed oil in both the deep sea and on the sea
surface need to be addressed with controlled empirical experiments.

Introduction

Oil spill response can be laborious, expensive and fraught with conflicting priorities. The implementation
and effectiveness of the various oil spill response options available are influenced by a variety of factors,
such as oil types and properties, environmental and ecological conditions in the vicinity of the spill site
(e.g., weather, wave, wind, daylight, ice conditions, visibility and sensitivity and distribution of species),
technical, logistical and financial factors (such as responders’ knowledge and skills, availability of
personnel and equipment, time constraints, regulatory approvals, health and safety criteria, cost and
economic impacts), and other constraints (e.g., community engagement). Some of those conditions and
factors are temporally dynamic and interactive and should be comprehensively considered in oil spill
contingency planning and response decision-making.

Many of the approaches and technologies that


Spill response strategies are varied and greatly have been used for treating oil spills were
affected by a variety of factors, such as the type of oil developed for marine shorelines, on-water-
(crude, refined, dilbit, etc.), the characteristics and spills and freshwater environments. These
remoteness of the spill site (water or ice, terrestrial, processes have been reviewed and described
marine or freshwater shoreline, riparian zone, arctic, extensively in a number of technical
permafrost, etc.) and even political considerations. documents, such as: Doerffer (1992), NOAA
(1992), NOAA and API (1994), U.S. EPA
(1999), NRC (2014), Canadian Coast Guard
(2005), IMO (2005), Fingas (2011) and
numerous ASTM manuals, monographs and data series
(http://www.astm.org/DIGITAL_LIBRARY/MNL/mnltocall.htm). The most commonly used spill
response cleanup options (Table 6.1) are briefly described in the following text. The rest of this chapter
summarizes these technologies in more detail.
Table 6.1. Spill response cleanup options with chapter section designations that discuss each
Category of Response Options Example Technology (Section)
Monitored natural attenuation (MNA) (6.1.1)
Natural processes Evaporation (6.1.2)
Photooxidation (6.1.3)
Containment and recovery (6.2.1)
Sorbents (6.2.2)
Removal from shorelines (6.2.3)
Vegetation cutting (6.2.4)
Physical processes
Removal of oiled sediment (6.2.5)
Oiled Sediment reworking (6.2.6)
Physical dispersion (6.2.7)
In situ burning (6.2.8)
Bioremediation (6.3.1)
Phytoremediation (6.3.2)
Chemical dispersion (6.3.3)
Chemical and biological processes Surface washing/flushing (6.3.4)
Solidifiers (6.3.5)
Herding (6.3.6)
Debris removal (6.3.7)

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 195
6.1 Natural Processes

6.1.1 Monitored Natural Attenuation (MNA)

In this section, monitored natural attenuation (MNA) is discussed in the context of circumstances that
would indicate it to be the method of choice for a given spill response. Examples of such cases include
spills at remote or inaccessible locations when natural removal rates are expected to be rapid, or spills in
sensitive wetland sites where cleanup
MNA or natural recovery is an option that should always be actions by human intervention (e.g.,
considered when determining which response option should be trampling) may cause more harm
implemented. It may be the best alternative especially in than good. Since some natural
certain situations where human intervention could result in weathering processes are slow
more harm than good. MNA is essentially a no direct response (Figure 2.4), the monitoring
option that depends on nature to degrade the oil. For some component of MNA is essential to
spills, such as those that have contaminated highly sensitive verify the loss of residual
wetland and salt marsh environments that serve as habitats for hydrocarbons and the recovery of
endangered species occupying several trophic levels, it is impacted plant, animal and other
probably more cost-effective and ecologically sound to let the species essential for an ecologically
site recover naturally than attempt to intervene. healthy habitat.

A primary mechanism of MNA is


biodegradation. This natural process
is particularly important in removing the non- and semi-volatile components of oil from the environment.
The ultimate fate of spilled hydrocarbons in the environment is dependent on the role indigenous
microbial communities play in the fate of the majority of hydrocarbons entering the biosphere (Leahy and
Colwell 1990; Prince et al. 2010; Atlas and Hazen 2011). This is a relatively slow process and may
require weeks, months or even years for microorganisms to degrade a significant fraction of oil,
particularly if other factors (discussed in Chapter 3) are limiting. In recent years, with the advent of
genomics and metagenomics research activities, a shift has occurred in our understanding of the
ecological significance of microbial adaptation in response to oil spills that identify the change in
structure and function of whole microbial communities in previously unknown ways (Macnaughton et al.
1999). Pre-genomics ecological studies relied on the use of microbial isolates to guide our understanding.
In contrast to the results of studies of isolates that represented only a small percentage of the community
structure developing in or near the spill site, recent advances in genomic studies have shown that
microbial activity in low temperature and deep sea conditions may be quite robust (Atlas and Hazen 2011;
Valentine et al. 2015).

6.1.2 Evaporation

Evaporation is an important natural cleansing process during the early stages of an oil spill, as the lighter
weight, higher vapour pressure components in oil are removed rapidly when exposed to the atmosphere
(Chapter 2). Depending on the composition and mass of the oil spilled, up to 50% of the more toxic,
lighter weight components of oil may evaporate within the first 12 to 24 hours following a spill (U.S.
EPA 1999).

6.1.3 Photooxidation

Photooxidation occurs when oxygen under sunlight reacts oxidatively with aromatic oil components (see
Chapter 2; NRC 2003). Maki et al. (2001) reported that when Arabian light crude oil was first
biodegraded and then exposed to sunlight, a substantial decline in the aromatics fraction with a
concomitant increase in the resins and asphaltene fractions was observed. The oxygen content in the oil
and thus the polarity increased as the irradiation was prolonged, and the bioavailability of the biodegraded

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 196
oil increased by the photooxidation.
Photooxidation aids in the removal of dissolved Photooxidation changes the physico-chemical
hydrocarbons in water. Aliphatic and aromatic properties of oil and its related components
hydrocarbons are oxidized photochemically to with the oxygenated forms being more polar,
oxygenated, more polar compounds that are more increasing the toxicity of the oil (Wang and
soluble than the parent compounds. Photooxidation Fingas 2003). The photooxidation products,
may also result in higher molecular weight products while toxic, would not likely occur at toxic
through condensation reactions, ultimately leading concentrations in water. However, evidence
to tar and gum residues. has shown that photooxidation within the
tissues of transparent aquatic species causes
rapid mortality (Chapter 4). Thus,
photooxidation may either produce potentially higher biodegradation rates due to increased hydrocarbon
bioavailability or potentially lower rates due to toxic byproducts, depending on the oil composition and
species sensitivity.

6.2 Physical Response Methods

6.2.1 Containment and recovery

According to Etkin and Tebeau (2003), a functional effectiveness


of 10 - 30% can often be realized using mechanical recovery, with
Booms and skimmers are usually
levels of 50% and greater being achieved on occasion (e.g. 95%
the first response method to an
was achieved following the 2007 Burnaby syncrude spill in
oil spill on water. It has met with
British Columbia [Crosby et al. 2013]).
varying degrees of success,
mostly limited, depending on During a spill response operation, the threat to sensitive near-
weather and sea state conditions. shore/coastal habitats by a spreading oil slick can often be
mitigated by deploying booms to trap and concentrate the
advancing slick so that the oil can be physically recovered from
the surface of the water by specially equipped boats or other devices (skimmers). Booms are floating,
physical barriers made of plastic (usually marine grade PVC or urethane with a molded polyethylene
shell), metal (for in situ burning [ISB]), sorbent fabrics or other materials having a cylindrical float at the
top with a weighted bottom skirt under the water, which slow and trap or contain the spread of oil (NOAA
2015). Typical photos of boom deployment are shown in Figures 6.1 and 6.2.

Booms are commonly placed across a narrow entrance to a larger mass of water so that oil is prevented
from passing through into marshland or other sensitive habitats. Deflection booms are used to deflect oil
away from sensitive locations, such as shellfish beds or beaches used by birds as nesting habitats. Fire
booms, discussed in more detail in Section 6.2.10, are used to collect oil from a slick and concentrate it
into a thickness suitable for burning. Booms for ISB are typically fabricated with floating metal cylinders
at the top that support a boom and skirt constructed with fire-resistant materials.

Although booms can be used in calm water (e.g. for streams, canals, ponds, lakes), open water (for
harbours and open ocean conditions) and some fast water environments (for rivers, streams, estuaries and
moving water lakes), their effective operational range is limited by rough weather and winds that induce
strong currents and breaking waves.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 197
Figure 6.1 Oil containment boom deployed in a saltwater application. Source: NOAA

Figure 6.2 Boom surrounding a set of floating pens at a salmon hatchery in Prince William Sound, AK, to
protect the pens from oil spilled from the Exxon Valdez tanker. Source: NOAA

Sorbent booms, also known as ‘spaghetti’ or ‘sausage’ booms, are used for collecting thick oil in flowing
waters. They are composed of absorbent strips in the interior of the boom contained by open-cell netting.
This type of boom allows oil to penetrate more deeply into the interior, which increases absorbency and
helps prevent spreading from the spill site. Importantly, the Oil Budget Calculator study during the
Deepwater Horizon (DWH) response (Federal Interagency Solutions Group (FISG) 2010) showed that
despite almost ideal weather conditions and substantial logistical support, physical recovery accounted for
only about 4% of the cleanup due to limited oil encounter rate between the oil from a continuous
discharge into an open water environment over a period of several months and mechanical recovery
systems with limited spatial coverage. This estimate was later modified to 4 - 10% (Fingas 2013c) due to
a calculation error by the FISG.

Skimmers are boats and other devices (e.g. weirs) designed to recover oil from the sea surface (e.g.,
Figure 6.3 and 6.4). Each type of skimmer has its own characteristics and limitations in dealing with
factors such as viscosity, sea state and debris. Typically, two boats will tow a collection boom to
concentrate the oil to facilitate its recovery. High waves may compromise the ability of boom
containment and skimmers to remain in contact with the oil. Devices that have a small mass, allowing
them to follow wave movements, should be part of the spill response inventory.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 198
Figure 6.3 A ship skims oil spilled after the Deepwater Horizon/BP well blowout in the Gulf of Mexico in April,
2010. Source: NOAA

Figure 6.4 Skimming the Delaware River in Philadelphia following the M/T Athos I spill on November 26, 2004.
Source: NOAA.

Weir skimmers use a different oil collection method from those described previously. The skimmers float
closer to the surface slick where the weir collects the oil while still allowing water to flow through. Some
weir skimmers use a flotation system
featuring four evenly spaced floats that
Skimmer systems such as weir skimmers may be more help the weir adapt to changing water
effective for high-viscosity oils, such as diluted levels. This type of skimmer may be
bitumen, whereas skimmers with oil-attracting coatings connected to a pump or vacuum system for
may be more appropriate for conventional crude oil rapid separation and removal of the
spills. Weir skimmers entrain large quantities of water, floating oil. Vortex skimmers use an
which enables them to cope with higher viscosity oils. induced eddy to separate the oil from the
water based on their differences in density.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 199
According to Wadsworth (1995), oil viscosity affects the efficiency of virtually all recovery devices.
Recovery is affected by the tendency of many oils to form water-in-oil emulsions, whereby, in addition to
making the oil considerably more viscous, the volume of material is increased by three- to four-fold. The
kinematic viscosity 1 may increase to as much as 100,000 cSt as a result of emulsion formation. The
injection of demulsifying agents2 can be used to mitigate this problem and minimize the storage volume
required for recovered oil. Conventional weir, vortex, oleophilic rope and disc devices are usually limited
to use with oils having a maximum viscosity of 10,000 cSt, whereas screw, belt and air conveyor devices
will often perform well with oil of viscosities up to 100,000 cSt.

Physical recovery is limited by high wave action, leading to the loss of oil from containment booms either
through oil splashing over the booms or to poor wave-following characteristics of the booms, resulting in
bridging between crests. Skimmers are also subject to limitations from wave action as failure to remain in
contact with the oil often results in the uptake of large quantities of water. In addition, turbulence can lead
to loss of oil under the skimmer. Thus, a recovery device should have a small mass to enable faithful
following of wave movements.

For oils that do not flow readily, such as heavy crudes and refined products, and ambient temperature
below the pour point, the oil will behave as if it were a solid and thus resist recovery. If conditions are
right, oil tends to form oil-in-water emulsions, increasing viscosity and volume of the emulsified material.
The recovery efficiency, defined as the relationship between the quantity of oil relative to water in the
collected material, is an important determinant of a system's overall performance. Some oleophilic
systems recover small quantities of water but are less effective at high oil viscosities. These systems are
best used with smaller vessels that are more effective in restricted water depths. Non-selective systems,
such as weir skimmers, entrain larger quantities of water, and this property often enables them to deal
with higher viscosity oils, such as dilbit or refined fuel oil.

6.2.2 Sorbents

This subsection is adapted from Region 10 Regional Response Team and the Northwest Area Committee
(2015). The purpose of using sorbents as a spill response tool is to aid in the cleanup and recovery of oil-
contaminated shorelines, sensitive habitats such as wetlands and salt marshes, and small spills on water.
The primary uses of sorbents include the following (adapted from Merlin and Le Guerroué 2009):

 Containment and recovery by rapid deployment in coastal areas, ports and harbours, estuaries and
on rivers;
 Containment of slicks in association with a standard boom (to improve watertight seal);
 Protection of areas difficult to clean (riprap, reed beds, mangroves, etc.);
 Immobilization or recovery of floating pollutants on lakes or in stagnant waters;
 Rapid application on terrestrial spills or ground surfaces to prevent or at least reduce infiltration
of the pollutant to the substrate;
 Sorption of leaks below a recovery worksite;
 Sorption of effluents from cleanup of rocks, structures and embankments;
 Sorption by filtration of pollutants suspended in the water column (water intakes, rivers);
 Cleanup or decontamination of personnel and equipment on cleanup sites; and
 Lining and protection of pathways.

1
Kinematic viscosity is the ratio of dynamic viscosity to density. The SI unit of dynamic or absolute viscosity is cen-
tipoise (cP) or mPa·s, and for water cP and cSt are equivalent. The kinematic viscosity of any fluid in cSt (cen-
tistokes) is the dynamic viscosity divided by the density of the fluid (in our case the oil). So, for oil, the kinematic
viscosity will not be equivalent to the dynamic viscosity because of the division by oil density.
2
Demulsifiers are a class of chemicals used to separate emulsions, e.g., water in oil, into oil and water phases.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 200
An adsorbent takes up significantly less volume of liquid than an absorbent (Brinkman 1998). In Canada,
sorbent pads and booms are frequently used in spill response. However, loose organic materials, such as
peat, clay, straw, feathers, etc., are less prevalent because of the creation of more oily waste that must be
burned or disposed to landfills. The preference is to collect bulk oil.

6.2.2.1 Selection of sorbents for floating oil spills on water

For fluid oil on open water with no current, bulk, pads and pillows should be used for containment. On
open water with current and smooth-surfaced embankment, pads, pillows or sorbent booms with ballasted
skirt should be used. On rough-surfaced embankments, either roll or sorbent booms with ballasted skirt
should be used. For slow current (< 0.2 m/s) and small quantities of oil, sorbent rolls are used. For
stronger current in a larger water body, standing floating booms or sorbent booms are used, but if these
fail, sorbent rolls reinforced with rope are used instead.

For viscous oil, such as cold heavy fuel oil, dilbit or weathered emulsified crude oil, in open water with
no current, standard floating booms or sorbent booms with ballasted skirt are used for containment or
bulk and mops for recovery. In open water with current, rolls or sorbent booms with ballasted skirt or
even mops are used if the current is slow. For deflection or recovery in faster current, sorbent booms with
ballasted skirt or mops attached to a rope are used.

Recently, a natural plant-derived sorbent (sugarcane bagasse) was investigated in laboratory microcosms
as a remediation strategy for low energy intertidal wetlands contaminated by crude oil (Chung et al.
2010). The results indicated that the use of this type of sorbent was beneficial not only in removing oil but
also in preventing further contamination. This technique has potential to stimulate biodegradation by
wicking oil out of contaminated anaerobic wetland subsurface to the aerobic zone on the surface where
rapid biodegradation can take place. The bagasse, being biodegradable, would not need to be removed
subsequent to treatment.

The American Society of Testing and Materials (ASTM) developed performance standards for
adsorbents (Method F726) and absorbents (Method F716). The combination of the liquid with the
solid causes the absorbent to swell, thus absorbing many times its volume. The liquid is no longer
available for release and the rate of vapour release is reduced by five to six times that of
adsorbents.

Three types of sorbents may be used in oil spill response: 1) products of mineral origin, such as
expanded perlite or glass wool; 2) products of animal or vegetable origin, such as coarse sheep’s
wool, peat or cellulose peat or cellulose; and 3) synthetic products and organic polymers, such as
polypropylene or polyurethane. Sorbents are selectively either hydrophobic, designed to recover
non-polar pollutants (non-miscible with water), or hydrophilic, designed to recover polar
products, such as water or materials soluble in water or non-polar products. Because these types of
sorbents do not necessarily float on water but soak it up, they are used only on land.

Recommendation: Sorbents are widely used as a cleanup tool to combat oil spills in water and on
shorelines. They have limitations, however. The need exists to determine if any advances have been
or can be made to improve sorption as an effective response tool, especially in the use of
biodegradable natural organic sorbents, such as bagasse that may possibly be left in place for
biodegradation to take place. Significant amounts of debris and solid waste are generated when
using sorbents for cleanup, so if natural sorbents can be developed that would sorb oil and then
remain in place for ultimate cleanup by biodegradation, this would be a significant advancement in
oil spill response.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 201
6.2.3 Oil removal from shorelines

Various beach washing techniques at ambient and high temperatures under both high and low pressure
settings have been used as part of containment and recovery operations to treat different types of substrate
contaminated by oil.

6.2.3.1 Sand beaches

When oil first arrives on a beach, deposits of fresh product must first be removed before cleanup can take
place by any means. One technique used for free product removal is scraping and either bagging or
trenching (Figure 6.5). Large volumes of bulk oil are removed quickly, without generating excessive
waste, by scraping the free product directly from the sand surface. Rubber squeegees are used to manually
concentrate the oil, which is then pushed into trenches dug parallel to the waterline. For long stretches of
contaminated sandy beach, the oil is collected by vehicles equipped with scraper blades and dumped into
temporary storage pits above the high tide line. The oil is collected by vacuum trucks for later treatment.

Figure 6.5 Workers clean a sandy beach using traditional methods of hand labour and bagging the oil and
debris. Source: NOAA

6.2.3.2 Pebble or shingle beaches

These types of beaches are ‘armoured’ with pebbles or small- to medium-sized cobbles as opposed to fine
sand. They are steeply sloped because waves are able to course through the highly porous surface, which
decreases the effect of backwash erosion and increases sediment formation. In these low energy beaches,
vehicle access is usually possible, and bulk oil is either collected into pools onto the beach surface or
flushed into man-made trenches. The recovered oil is then collected in vacuum trucks and either
transferred to temporary holding tanks or taken away for treatment. However, pebble beaches are difficult
to clean thoroughly because oil tends to penetrate into the subsurface, sometimes as far as the clay/mud
base. Thus, flushing into trenches requires later stages of cleanup, usually by pebble washing. In such
cases, a better cleanup method is to flush the oil off the beach at high tide into offshore booms, where the
oil can be collected by skimmers.

6.2.3.3 Cobble beaches

These types of beaches are characterized by higher energy wave action impinging on the surface. Thus,
surf washing (also known as berm relocation or oiled sediment reworking) is often used for treating

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 202
cobbles that have been exposed to beached oil (Lunel et al. 1996; Lee et al. 1997, 2003). Such cobbles are
placed in the high energy surf zone where they are scoured naturally by the wave action. Where removal
of large oiled cobbles has been difficult, large pits lined with plastic have been dug into the beach where
the cobbles are washed with an oil-releasing agent and water. Oil is then skimmed and the clean cobbles
are returned to the beach surface. However, this method usually leaves the cobbles oil-stained and is not
used in sensitive environments, such as salmon spawning streams, shellfish beds, nursery areas, etc.,
because of the potential adverse effects on those sensitive habitats from the increased oil runoff, sheening
or siltation.

Recommendation: Berm relocation, surf washing and sediment reworking have been used to
accelerate shoreline cleanup in the past. A comprehensive review is needed to determine
quantitatively how successful these technologies have been and where such techniques are best
applicable or not recommended. Testing protocols need to be developed to determine the
effectiveness and secondary ecological impacts of this response option.

6.2.3.4 Boulder beaches

Boulder beaches are usually found in high wave energy environments where clasts of these large
dimensions are released directly by erosion of bedrock, or where material is delivered to the shore zone
by slope movements, such as rockfall. Oil stranded in these areas can become mixed with sand, forming
an oil/sand coating on the rock surfaces. This coating can be removed relatively easily by surface
scrubbing, brushing or wiping (Figure 6.6). However, re-oiling often occurs in these types of areas due to
the constant movement of sand during tidal cycles, resulting in a protracted cleanup. Thus, wiping is a
highly labour intensive treatment technique. Other techniques, which may be more cost-effective than
wiping include high pressure cleaning with or without the use of surface washing agents. If the impinging
wave energy is high enough, natural recovery via physical abrasion from the waves is probably a better
option.

Figure 6.6 Workers clean oiled boulders by hand. Source: NOAA.

6.2.3.5 Bedrock shoreline with no beach

This shoreline type is frequently found along the coastal areas of Canada and common in many Canadian
lakes, including portions of the Great Lakes (e.g., Figure 6.7). Rock surfaces on marine shorelines are
often densely populated with kelp (and associated biofilm). They can have a thriving tide pool community
and therefore should not be regarded as ‘barren’. Furthermore, wet kelp may reduce oil adhesion to rock
surfaces. Figure 6.8 shows workers cleaning oil spilled following a collision involving crude oil tanker
Eagle Otome in Port Arthur, TX, on January 2010. Under inclement weather conditions following a spill,

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 203
waves may deposit oil up on the rocks, which may have a lichen/moss coating or be inhabited by other
vegetation, such as small willows or conifers that cling to the cracks in the rock.

Figure 6.7 North shore of Lake Superior. Source: NOAA.

Figure 6.8 Workers cleaning oil spilled following a collision involving crude oil tanker Eagle Otome in Port
Arthur, TX, on January 2010. Source: NOAA.

6.2.3.6 Low pressure washing or sediment flushing of a variety of beach types

Flooding a beach (Figure 6.9) with seawater can be used to flush fluid oils and oily debris from different
shoreline types, particularly in sensitive areas (IMO 2005). This approach is less damaging to flora and
fauna than most other methods. Booms are used to contain the flushed oil for skimming and removal.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 204
Figure 6.9 Flushing out oil buried in a sand beach using low pressure water supplied through lances and
perforated pipes. Photo Credit: ITOPF; http://www.itopf.com/knowledge-resources/documents-guides/response-
techniques/shoreline-clean-up-and-response/

6.2.4 Vegetation cutting

Vegetation is cut to remove oil-contaminated vegetation from shorelines, wetlands or salt marshes
(Figure 6.10) to prevent contamination of wildlife or remobilization of trapped oil. This strategy is based
on the removal of oil from impacted environments by the pruning and collection of the upper parts of
oiled plants for waste disposal. The cut vegetation is then collected and disposed. Any oil remaining near
the roots or around the stems can be flushed out and recovered. Cutting can result in a temporary loss of
habitat for small fish and invertebrates that dwell in those areas. Its feasibility depends strongly on the
season in which the spill occurs. In general, winter cutting of dead standing vegetation has little effect on
subsequent growth, but summer cutting could cause great damage to the regrowth of wetland plants and
result in shoreline erosion. The use of cutting should also be avoided immediately prior to an anticipated
rise in water levels because cutting followed by flooding could cut off necessary oxygen to plant roots
(Pezeshki et al. 2000). The plants may or may not recover, depending on the type of oil spilled and
differences in species and life stage sensitivity. This response method is generally used when large
amounts of potentially mobile oil are trapped in the vegetation or when the risk of oiled vegetation
contaminating wildlife is greater than the value of the vegetation that is to be cut. This response is
conditionally recommended for sheltered, vegetated, low banks and marshes. See Chapter 4 for a
discussion of the environmental impacts of vegetation cutting in the floodplain zone.

Vegetation cutting is also used for the removal of oil in the canopy of kelp beds, where thick layers of oil
may adhere to kelp fronds or collect under the kelp canopy. This procedure is similar to shoreline
vegetation cutting. The upper ~0.5 m of the kelp canopy are cut away by hand or with a mechanical kelp
harvester and removed for disposal. Vegetation cutting is used when a large quantity of oil is trapped in
the kelp canopy and the oil poses a risk to sensitive wildlife using the kelp habitat or when the
remobilization of oil to other adjacent sensitive environments is likely to occur.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 205
Figure 6.10 Cleanup workers removing oiled vegetation during the Deepwater Horizon spill and using walk
boards while cleaning to avoid causing further damage to the oiled marshes in Barataria Bay. Source: NOAA.

6.2.5 Removal of oiled sediment

This response technique is used to remove oiled surface sediments that cannot otherwise be cleaned in
situ. Oiled sediment is removed by use of hand tools (Figure 6.11) or motorized equipment. Oiled
sediment removal is restricted to the supratidal and upper intertidal areas to minimize disturbance of
biological communities in the lower intertidal and subtidal zones. After removal, oiled sediments are
transported for cleaning and/or disposal offsite. New sediments are not usually imported to replace those
that were removed; however, a variation of this response that includes sediment replacement (described
below) is used for beaches with low natural replenishment rates or high rates of erosion.

Physical removal of oiled sediment is frequently used on high amenity ‘tourist’ beaches (where pressure
is high for rapid cleanup) and is most effective when a limited amount of oiled sediment must be
removed. Close monitoring is required so that the quantity of sediment removed, siltation and the
likelihood of erosion may be minimized in all cases. Such operations are generally restricted in fish
spawning areas. Sensitive areas that are adjacent, and may be potentially affected by released oil sheens,
must also be protected. During the DWH spill at the Grand Isle State Park in Louisiana, remediation
operations included the excavation of the top few centimetres of contaminated sand for the removal of
bulk oil by physical/chemical extraction at a temporary facility constructed near the spill site and its
subsequent return to the beach (Figure 6.12) where the treated sediments were also surf washed for final
polishing.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 206
Figure 6.11 Cleanup workers manually remove oil following the M/V Westchester spill in the Mississippi River
near Empire, LO, on November 2000. Source: NOAA.

Figure 6.12 Oiled sand removal in Grand Isle State Park following the DWH spill. Source: K. Lee, DFO.

6.2.6 Oiled sediment reworking

The objective of this variation of oiled sediment removal is to rework oiled sediments to break up oil
deposits, increase surface area, facilitate physical oil dispersion and mix oxygen into deep subsurface oil
layers. This activity exposes the oil to natural removal processes and enhances the rate of oil
biodegradation. Under this protocol, oiled beach sediments are rototilled or otherwise mechanically mixed
with the use of heavy equipment or relocated from the upper intertidal area of the beach to the surf zone.
This latter procedure is also known as surf washing or berm relocation (Section 6.2.3). Generally,
sediment reworking is used on sand or gravel beaches where high erosion rates or low natural sediment
replenishment rates are issues. Sediment reworking may also be used where remoteness or other logistical
limitations make sediment removal unfeasible. Sediment reworking is not used on beaches near shellfish
harvest or fish spawning areas because of the potential for release of oil or oiled sediments into these
sensitive habitats. Sediment reworking is conditionally recommended for: (1) sand beach; and (2) cobble
and gravel beach habitats (Lunel et al. 1996; Region 10 Regional Response Team and the Northwest
Area Committee 2015).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 207
A new oil spill monitoring protocol called ‘poling’ was used in the Kalamazoo River spill in 2010 as a
means of finding oil in the sediment. Poling involves disturbing the river sediment with a pole every 50 m
or so to locate any remnants of oil residue remaining in the riverbed from the spill. Oil distribution in the
sediments was estimated from the observation of oil sheens on the surface after disturbing the riverbed by
poling. As a result of this survey method, identified areas with significant oil concentrations within the
sediment were physically agitated by water jet washes as a means to facilitate residual oil release to the
surface waters where it was physically recovered by skimming and the use of sorbent pads.

6.2.7 Physical dispersion (Oil mineral aggregates)

Suspended particulate matter within the water column, such as plankton and mineral fines, may form
nuclei around which oil droplets may interact and adhere (Muschenheim and Lee 2003) to form oil-
mineral aggregates (OMAs; Chapter 2). The subsequent ecological consequence of this process is
dependent on the oil concentration and sedimentation rates. If the material is retained within the water
column by physical oceanographic processes, oil toxicity may be reduced due to dilution of the residual
oil by dispersion and enhanced oil biodegradation rates (as the process may enhance the surface area of
the oil and nutrient availability; Figure 2.3). However, if the material rapidly sinks into the sediment and
accumulates at elevated concentrations, biological impacts may result and the oil may degrade very
slowly and thereby persist for long periods of time (see Chapter 3 for a discussion of anaerobic
conditions) (especially if entrained within anaerobic sediments; Figure 2.3).

Several in situ remediation technologies, bioventing, air sparging and dissolved air flotation (DAF),
developed over the last 20 years for the treatment of contaminated soils and groundwater have not been
tried for remediating bottom sediments contaminated by sunken oil from spills. These technologies are
characterized by injection of oxygen-rich air into the sediment to accelerate the aerobic biodegradation of
oil in that anaerobic environment. Specifically, DAF achieves contaminant removal by dissolving air in
the water or wastewater under pressure and then releasing the air at atmospheric pressure in a flotation
tank or basin. The released air forms tiny bubbles that adhere to the suspended matter causing it to float to
the surface of the water where it may then be removed by a skimming device.

Recommendation: Bioventing, sparging and DAF should be evaluated for their efficacy in treating
and overcoming the adverse effects of sunken oil in the anaerobic zone of river or lake sediments. If
successful, such techniques would greatly help in cleaning up benthic environments that have been
contaminated by oil.

6.2.8 In Situ burning

ISB, a technique used to rapidly reduce the mass of oil spilled at a site, involves the controlled burning of
oil that has spilled from a vessel or a facility at the location of the spill. ISB was successfully
demonstrated by Environment Canada in a large-scale field experiment, the Newfoundland Offshore Burn
Experiment (NOBE), on August 12, 1993. During each of two test burns, crude oil was poured into a
towed U-shaped fireproof boom and ignited. The first test burn lasted for an hour and a half and the
second for about an hour, with an average observed burning rate of 24 tonnes of oil per hour (Fingas et al.
1995a). ISB was conclusively validated as an operational physical oil spill response countermeasure
following controlled burns near a swamp (Figure 6.13) and at sea (Figure 6.14) during the DWH
incident.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 208
Figure 6.13 A view of one of the controlled burns to remove oil spilled in a wooded swamp outside of Baton
Rouge, LA, on January 19, 2013. Source: USCG.

Figure 6.14 After the Deepwater Horizon oil spill in the Gulf of Mexico in April 2010, ISB was one of the
techniques used to remove oil from the water. Source: NOAA.

During at-sea ISB operations, oil on the sea surface is captured within a boom towed by two boats in a U-
configuration and ignited using a hand-held igniter or an igniter suspended from a helicopter. Under
favourable conditions ISB is a fast, efficient and relatively simple way of removing spilled oil from the
water to minimize the adverse effect of the floating oil on the environment. Furthermore, it greatly
reduces the need for storage and disposal of the collected oil and the waste it generates. The burn will
continue only as long as the oil slick is thick enough—usually about 2 - 3 mm. Burning oil results in
residues (approximately 1 - 5% of the starting oil) from incompletely combusted oil and gaseous
emissions into the atmosphere that have raised environmental concerns. This is particularly true if the

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 209
residue sinks. Results of laboratory tests suggest the possibility that burn residues from ~40 to 60% of
crude oils may sink. However, whether results from laboratory tests can be extrapolated to large-scale
spills is not known (NOAA 2015).

To burn oil on the water, slicks must be thicker than 1 - 2 mm. During combustion the oil vapours ignite
and burn, rather than the liquid itself. When the oil layer is thinner than ~1 mm, the heat is lost to the
water, not enough vapours are released, and combustion ceases. Experiments have shown that ISB is best
used under relatively calm conditions, as choppy seas may extinguish the fire altogether (ITOPF,
http://www.itopf.com/knowledge-resources/documents-guides/response-techniques/in-situ-burning/).
When the current is stronger than about one knot, the boom cannot contain the oil, which splashes above
the boom or escapes beneath it. This resulting entrainment typically occurs when booms are deployed
perpendicular to the water flow (API 2015). Water-in-oil emulsions of over 50% will preclude ISB of
even light crudes or refined products, while much less than that is required for heavier crudes (NOAA
2015).

Studies of the emissions from ISB have shown fairly consistent results. According to Ferek et al. (1997),
about 85 to 95% of the burned oil is converted to carbon dioxide and water, 5 to 15% of the oil is not
burned efficiently and is converted to particulates, mostly soot, and the rest, 1 - 3%, comprises nitrogen
dioxide, sulfur dioxide, carbon monoxide, polycyclic aromatic hydrocarbons (PAHs), ketones, aldehydes
and other combustion by-products. The 5 - 15% estimated conversion to particulates differs from the 1 -
5% reported above, suggesting that a number of factors are in play when making estimates of ISB
conversions. No ‘exotic’ chemicals are formed. Rather, the burning of oil on water seems to be similar to
burning the oil in a furnace or a car’s internal combustion engine, with the exception that the burn is
oxygen-starved and not very efficient, so that it generates ample amount of black soot particulates that
absorb sunlight and create black smoke. ISB is not used if human populations are located near and
downwind from the site. For oil spills occurring in ice-covered water or wetland or salt marsh
environments, besides MNA, ISB may be the only spill response method available. ISB, however, should
complement, not exclude, other means of spill response. ISB has been used successfully on numerous
occasions when oil was trapped in ice or spilled into sensitive marshland.

Recommendation: ISB is fast gaining a foothold as a response technology for treating floating oil
slicks. More research, including peer-reviewed literature, is needed to ascertain harmful emissions
released to the environment during ISB, such as any incompletely combusted polycyclic aromatic
hydrocarbons toxic particulate matter, etc., and to determine the safety and effectiveness of ISB as
an oil spill response tool.

6.3 Biological and Chemical Methods

6.3.1 Bioremediation

Many hydrocarbons (normal and cyclic alkanes, most monoaromatics and some PAHs; Table 2.1) are
biodegradable under aerobic conditions, although many polar resins, most hopanes and asphaltenes are
resistant to microbial attack. Thus, bioremediation is a viable response technique for cleaning up a
shoreline contaminated by an oil spill, with the caution that light crude oils are much more easily

Bioremediation is the exploitation of the ability of microorganisms to convert pollutants, such as


petroleum hydrocarbons, into biomass, carbon dioxide, water and innocuous oxygenated end
products. The oil serves as food for the microbes, providing energy and carbon for reproduction and
growth. Biodegradation is the process microorganisms use to obtain that end. Thus, bioremediation is
a human intervention, whereas biodegradation is a natural property of microorganisms.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 210
biodegraded than diluted bitumen and heavy refined products, such as fuel oils. The passage of time has
allowed the evolution of numerous diverse microbes to utilize hydrocarbons as a source of carbon and
energy for growth (Atlas and Hazen 2011). It should also be noted that many of the oil deposits around
the world, especially those near the surface, are actually residues of millions of years of bacterial
degradation of organic, mostly plant-based material.

Two primary approaches are used in bioremediation: bioaugmentation and biostimulation.


Bioaugmentation is the addition of an exogenous supply of oil-degrading microorganisms (grown
offsite or in the laboratory) to the oil-impacted environment to supplement (augment) the existing
microbial populations and accelerate biodegradation. Biostimulation describes the addition of
nutrients or other growth-limiting chemicals, such as electron acceptors, to accelerate biodegradation
by the indigenous microbial communities that are already present. Both these approaches have been
extensively studied in the laboratory and the field.

The rationale for bioaugmentation is that indigenous microbial populations may not be capable of
degrading the wide range of potential substrates present in petroleum (Leahy and Colwell 1990), and that
addition of selected, competent microbes would permit efficient biodegradation. Other conditions under
which bioaugmentation may be considered are when the indigenous hydrocarbon-degrading population is
low, the speed of decontamination is the primary factor or when seeding may reduce the lag period to
start the bioremediation process (Forsyth et al. 1995). However, bioaugmentation has never been
demonstrated with statistical rigour to accelerate biodegradation to any appreciable extent in field studies.
McGenity et al. (2012) pointed out that the reasons for past failures of bioaugmentation have been: 1) use
of a pure culture rather than a community; 2) focus on biodegrading strains only, without partner
microbes; 3) microbes not adapted to the environment and unable to compete for nutrients with
indigenous microbes; 4) inadequate dispersion or access to the pollutant; 5) lack of protection from
indigenous grazers or predators (e.g., amoebae); and 6) other factors limiting biodegradation (nitrogen
and phosphorus).

Except for item 6, these factors were not in play in the statistically rigorous field research conducted by
Venosa et al. (1996) on the Delaware shoreline involving MNA, biostimulation and bioaugmentation. In
the bioaugmentation treatment, they applied both nutrients and indigenous microorganisms grown on site
in 55-gallon drums containing the same seawater and crude oil and grown on the same shoreline under the
same environmental conditions as the beach. The study design included quintuplicate plots of all three
treatments, and bioaugmentation was unable to further accelerate biodegradation beyond simple nutrient
addition, which did achieve statistically significant biodegradation rate enhancement. While inoculation
doesn’t seem to be an issue, as reported by Jacques et al. (2008), the future development of good
microbial consortia having complementary biodegradative pathways and bio-surfactant production
capabilities to enhance the bioavailability of residual hydrocarbons may someday make bioaugmentation
feasible on shorelines.

Biostimulation has been successful in a number of laboratory, mesocosm and field studies, making it the
strategy of choice in the open environment because of the ubiquity of oil-degrading bacteria (Venosa et
al. 1996; Lee 2000; Zhu et al. 2001, 2004). Microbial population density is generally not the primary
limiting factor over the time scales required for bioremediation. When high concentrations of oil are
suddenly introduced into the environment, the microbial communities are faced with a huge increase in
food (carbon) without a concomitant increase in the nutrients they need to grow at the maximal or near
maximal rate (i.e., nitrogen and phosphorus; Chapter 3). Thus, the rates of biodegradation will be
necessarily slow since growth of the microbes will be limited by the amount of nitrogen and phosphorus
available for the transformations. Hence, biostimulation is used to supplement the limited available

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 211
nutrients that indigenous microbes need to grow rapidly on the introduced carbon. This is true regardless
of the environment in which the oil spill occurred (shoreline, open water, wetland, land, etc.). Two
comprehensive guidance documents of oil spill bioremediation covering all aspects of microbial
interactions with spilled oil, published in 2001 and 2004 (Zhu et al. 2001, 2004), are still relevant today.

The Redfield Ratio (defined and discussed in Chapter 3) is the stoichiometric relationship among carbon,
nitrogen and phosphorus in plankton. To meet the theoretical demand for nitrogen and phosphorus during
biostimulation, approximately 150 mg of nitrogen and 30 mg of phosphorus are consumed in the
conversion of 1,000 mg of hydrocarbon to cell material (Rosenberg and Ron 1996). Therefore, a
commonly used field strategy is to add nutrients at concentrations that approach a stoichiometric ratio of
C:N:P of 100:10:1, which is similar to the Redfield Ratio of 106:16:1. However, the practical use of these
ratio-based theories remains a challenge. Particularly in marine shorelines, maintaining a certain C:N:P
ratio is impossible because of the dynamic washout of nutrients resulting from the action of tides and
waves. A more practical approach is to maintain the concentrations of the limiting nitrogen and
phosphorus within the pore water at a range optimal for rapid biodegradation (Bragg et al. 1994; Venosa
et al. 1996). Commonly used nutrients include quick release water-soluble, solid slow-release and
oleophilic fertilizers. Each type has its advantages and limitations. Zhu et al. (2001, 2004) and Venosa
and Zhu (2005) provide guidance on how to maintain these concentrations in the field. Importantly, the
optimum C:N:P ratio for anaerobic hydrocarbon biodegradation in situ has not yet been defined.

Recommendation: There is a need to quantify the persistence of oil of all kinds during and after
biodegradation and oil that has been chemically-dispersed in the environment following high-
impact spills to help stimulate innovations in bioremediation approaches. This is especially
pertinent to cold temperature and anaerobic environments.

6.3.1.1 Wetlands

Productivity in coastal marsh ecosystems is extremely rich,


Wetlands, salt marshes and mudflats and such systems provide habitat and breeding grounds to
are among the most sensitive and many vital species. Use of heavy equipment and human
difficult ecosystems to clean. trampling are not amenable to a viable response activity
because these cleanup actions may cause more damage than
the spill itself. If human presence is needed, wooden
walkways (Figure 6.10) or boat-based activities should be used as much as possible. Effective response
techniques for these sensitive environments may include less intrusive techniques, such as:

 Booms to contain and control the movement of floating oil at the edge of the wetland (Section
6.2.1);
 Sorbents for removal of the oil by adsorption onto oleophilic materials placed in the intertidal
zone (Section 6.2.2);
 ISB (but on a stringently limited basis) (Section 6.2.8);
 Bioremediation (Section 6.3.1);
 Phytoremediation (Section 6.3.2); and
 Low pressure flushing with ambient seawater at pressures less than 200 kpa or 50 psi to the water
edge followed by removal (Section 6.3.4).

Dispersants (Section 6.3.3) are unlikely to be appropriate for use near a coastal wetland or salt marsh due
to the lack of mixing energy to aid in effective dispersion into the water column, limited water depth to
effectively disperse the oil below toxicity threshold limits, and the unknown effect of either the dispersed
oil or the dispersant itself on ecosystems in the area. The use of dispersants in near-shore water could
have short-term toxic effects on adjacent coastal habitats, such as subtidal animal communities. Direct
spraying or contact of dispersants with wetland plants may also cause harmful effects on vegetation.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 212
6.3.2 Phytoremediation

Phytoremediation (Figure 6.15), which takes


Phytoremediation has been defined as the advantage of the mutually beneficial interactions
use of green plants and their associated between plants and microbes for environmental
microorganisms in the soil to degrade, cleanup (Cunningham et al. 1996)., is emerging as a
contain or render harmless environmental potentially cost-effective option for cleanup of soils or
contaminants. sediments contaminated with petroleum hydrocarbons
(Frick et al. 1999). As summarized by Macek et al.
(2000), the main advantages of phytoremediation
include less disruption to the environment, potential to treat a diverse range of contaminants and high
probability of public acceptance. Major concerns regarding this technology include dissolution and
migration of contaminants and relatively slow rates of remediation.

Degradation can be accomplished by both plants and associated microorganisms in the rhizosphere.
Plants can stimulate the growth and metabolism of rhizosphere microorganisms by providing root
exudates of carbon, enzymes, nutrients and oxygen, which can result in more than 100-fold increase in
microbial counts (Macek et al. 2000).

A principal factor involved in the enhanced degradation of hydrocarbons in soil is the ‘rhizosphere
effect3’, i.e., an increase in the numbers and activity of soil microorganisms in the plant-root zone. The
plant-root zone has a high surface area, which contributes to growth of oil-degrading microorganisms
(and other microbes) in response to high concentrations of hydrocarbons present in the rhizosphere
following oil incursion. In addition to elevated microbial community numbers, the contribution of plant-
root extracellular enzymes to the biodegradation of organic pollutants in the rhizosphere can be
considerable. This has been demonstrated by the findings of Gramss et al. (1999), who showed that the
roots of some plants release enough oxidoreductase enzymes to take part in the oxidative degradation of
certain soil constituents. Active involvement of plant peroxidase enzymes in the phytoremediation
process has also been suggested by several other authors (Kraus et al. 1999; Criquet et al. 2000; Chroma
et al. 2002), including PAH degradation (Muratova et al. 2009).

3
The rhizosphere is the zone of soil or sediment surrounding plant roots and roothairs that is directly influenced by
the living plant, e.g., exudates, moisture, nutrients, etc.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 213
Figure 6.15 Schematic representation of phytoremediation activities. “M” indicates microbial cells. Source: ©
2014 Favas, P.J.C., Pratas, J., Varun, M., D'Souza, R., Paul, M.S. Published in Soriano, M.C. (ed.) Environmental
Risk Assessment of Soil Contamination. InTech. Under CC BY 3.0 license. Available from:
http://dx.doi.org/10.5772/57469

Phytoremediation is mutually beneficial. The microbes can reduce the phytotoxicity of contaminants so
that plants can grow in adverse soil conditions. Co-metabolism 4 may also play an important role in
phytodegradation. Ferro et al. (1997) suggested that plant exudates might serve as co-metabolites in
enhancing the biodegradation of the 4-ring PAH pyrene in the rhizosphere.

Other major mechanisms of phytoremediation include retention of petroleum hydrocarbons and their
transfer from the soil to the atmosphere. Containment involves the accumulation of contaminants within
the plants, adsorption of contaminants onto roots and binding of contaminants in the rhizosphere through
enzymatic activities (Cunningham et al. 1996; Frick et al. 1999). Plants can also transport volatile
petroleum hydrocarbons to the atmosphere through leaves and stems. However, these effects are less
important than the rhizosphere effect during phytoremediation of petroleum hydrocarbons (Ferro et al.
1997).

4
Co-metabolism involves the partial oxidation of a chemical (e.g., hydrocarbon) by one organism (that receives no
benefit from the metabolism), and the product is further metabolized by another organism that does receive benefit
(e.g., can use the partially oxidized product for growth).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 214
Several full-scale studies in the 1990s and early 2000s highlight what is known about salt marsh
remediation. Lee and Levy (1991) conducted one of the first field trials on oil bioremediation in a salt
marsh environment. The study involved periodic addition of water-soluble fertilizer granules (ammonium
nitrate and triple super phosphate, providing the nitrogen and phosphorus) to enhance biodegradation of
waxy crude oil in a salt marsh in Nova Scotia dominated by Spartina alterniflora (saltmarsh cordgrass).
Results showed that the effectiveness of nutrient addition was related to oil concentration, i.e.,
effectiveness was high at 0.3% oil contamination by mass, but not at 3% oil contamination. Lack of
effectiveness at the higher oil concentration was attributed to the penetration of the oil into the reduced
soil layers where anaerobic conditions precluded rapid degradation.

Field trials of bioremediation (Mills et al. 1997) were conducted on a coastal brackish wetland in Texas
where two diesel pipelines, one gasoline pipeline and one crude oil pipeline ruptured, spilling ~4,700
tonnes into the tidal and coastal waters of the San Jacinto River. The effect of biostimulation was
investigated by evaluating the application of diammonium phosphate alone and diammonium phosphate
plus nitrate in replicated plots, providing nitrogen and phosphorus nutrients that otherwise would be
limiting. Results showed that the diammonium phosphate treatment significantly enhanced the
biodegradation rates of both total GC-MS-resolved saturates and total GC-MS resolved PAHs (Appendix
A), and the addition of diammonium phosphate plus nitrate only enhanced the biodegradation of total
resolved saturates.

Shin et al. (1999, 2000) investigated the effect of nutrient amendment on the biodegradation of a
Louisiana ‘sweet’ crude oil and oxygen dynamics in a Louisiana salt marsh vegetated by S. alterniflora.
Significant biodegradation of crude oil in the salt marshes occurred only when the tidal cycle exposed the
surface of the marsh to air (Shin et al. 2000). This study again indicated that oxygen availability appeared
to control the oil biodegradation process in salt marshes.

Phytoremediation was studied in a freshwater wetland on the St. Lawrence River, QC (Venosa et al.
2002) (Figures 6.16 and 6.17), and later on a salt marsh in Nova Scotia (Garcia-Blanco et al. 2007)
(Figure 6.18 and 6.19). Both studies intentionally released crude oil onto replicated experimental plots.
In the freshwater wetland study, the released oil was raked into the top two inches of sediment to simulate
penetration of the oil into the anaerobic or anoxic zone. In the salt marsh study, the oil was not raked into
the sediment but left on top of the aerobic sediment. Results indicated that addition of nutrients did not
result in enhancement of biodegradation of crude oil contaminating the freshwater wetland plots because
the oil was in the anaerobic zone where biodegradation is substantially slower. In such a wetland
environment, oxygen became limiting at depths within a few millimeters below the surface. However, in
the aerobic salt marsh study, the investigators found statistically significant depletion of the hydrocarbons
in nutrient-amended plots compared to those not receiving nutrients. This was a clear example of the
rhizosphere effect caused by the biostimulation from added nutrients.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 215
Figure 6.16 Photograph of crude oil being applied to one of the plots in the St. Lawrence field study of freshwater
phytoremediation. Source: Venosa et al. (2002)

Figure 6.17 Aerial photograph of the plot layout in the St. Lawrence River Field study of freshwater
phytoremediation. Source: Venosa et al. (2002)

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 216
Figure 6.18 Photograph of a team taking samples from a plot in an experimental phytoremediation project
conducted in 2001 in Dartmouth, Nova Scotia.

Figure 6.19 Aerial photograph of the experimental plot layout in the joint DFO-U.S. EPA salt marsh
phytoremediation project in the 2001 Dartmouth, Nova Scotia project.

Another field study on the performance of oil bioremediation in salt marsh ecosystems was carried out in
a tropical marine wetland located at Gladstone, Australia (Burns et al. 2000; Duke et al. 2000; Ramsay et
al. 2000). This study evaluated the influence of a bioremediation protocol on the degradation rate of a
medium range crude oil and a Bunker C oil stranded in salt marshes dominated by Tecticornia (a
succulent plant) and situated behind a mangrove forest. The bioremediation strategy used in this study
involved nutrient addition alone for the salt marsh sites. The addition of the fertilizer to the salt marshes
showed a stimulation of the degradation of the crude oil and resulted in about 20% more oil loss
compared to the untreated plots. However, the nutrient amendment did not significantly impact the rate of
loss of Bunker C oil in the salt marsh plots, likely because of its high proportion of recalcitrant
components (Appendix C).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 217
Recommendation: Most of what we know about oil spill response technologies has been developed
through results from studies in the laboratory, mesocosm test systems and case studies (with limited
controls and treatment replication). It is the consensus of the scientific community (and the Panel)
that controlled field trials are needed to advance our spill response strategies, especially for
subsurface blowouts, Arctic oil spills and freshwater shorelines.

6.3.3 Chemical dispersion

Dispersants are chemicals that reduce the oil-water interfacial tension (NRC 2005), thereby decreasing the
energy needed for the slick to break up into small droplets and mix into the water column. They are
typically applied by spraying from aircraft (Figure 6.20) and/or boats. The composition of most
commercially available dispersants is proprietary, but in general they consist of one or more nonionic 5
surface active agents (surfactants) dissolved in a solvent carrier (NRC 1989). Some dispersants also
include one or more anionic surfactants and other additives. Lower viscosity oils are more amenable to
natural and chemical dispersion than higher viscosity oils, such as fuel oils and dilbit.

The rationale for using dispersants as an oil spill cleanup tool is to reduce wind-driven transport of oil
to highly productive coastal waters and sensitive shoreline habitats by breaking the surface oil slick
into small droplets to facilitate the transport of the oil into the water column. This would also reduce
its exposure to surface water biota (e.g., sea birds). Furthermore, oil in the water column may be
diluted to concentrations below the toxicity threshold limits of resident biota. Because microbial
attack of oil occurs at the oil-water interface, droplet formation also enhances the biodegradation rates
of the residual oil.

Natural dispersion takes place when wave and wind action are sufficiently high (winds > 5 m/s, for
example) to overcome the oil-water interfacial tension and break the slick into droplets of various sizes
(Figure 2.6). A notable example of such occurred in 1993 when gale-force winds caused the grounding of
the tanker Braer in Shetland, UK, spilling a low viscosity crude oil into the water. Dispersion occurred
naturally, causing minimal shoreline impact. Under less energetic wave action, chemical dispersants
enhance natural dispersion, facilitating the breakup of the oil into small droplets that are transported into
the water column.

Another form of natural dispersion occurs when microorganisms, especially alkane degraders, produce
their own biosurfactants (Bodour et al. 2003). These biosurfactants are similar to the synthetic surfactants
in chemical dispersants in their mode of action, which is reducing the interfacial tension between oil and
water, solubilizing hydrocarbons and thus facilitating their uptake across the membrane into the cell for
metabolic purposes. Bodour et al. (2003) reported that biosurfactant-producing bacteria are widely
distributed in both undisturbed and contaminated soils. The majority of hydrocarbon-degrading bacteria
reported in the literature belong to the genus Pseudomonas (Widada et al. 2002; Bento et al. 2005). A
wide variety of metabolic and physiological factors is required for the degradation of different compounds
in oil (Frielo et al. 2001). All such properties are not found in one organism but rather in consortia. More
research is needed to augment our understanding of the effectiveness of these biosurfactants in enhancing
bioremediation of oil-contaminated environments.

Specially formulated products containing surfactants and solvents are sprayed (generally at concentrations
of 2 - 5% by volume of the oil) from aircraft or boats onto the slick. Surfactants are the active (i.e.,

5
Nonionic dispersants (e.g., some alcohols and ethers) do not have an overall electric charge. Anionic dispersants
are electronegative (e.g., have carboxyl, sulphate or phosphate groups) and cationic dispersants are electropositive
(e.g., have amino or ammonium groups).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 218
interfacial tension-reducing) ingredients in dispersants. Surfactant compounds are amphiphilic6 (i.e., have
hydrophobic and hydrophilic components within the same molecule), which causes them to accumulate at
oil:water interfaces (Porter 1991). Nonionic surfactants are most common in dispersants because they
have much lower aqueous solubility than do anionic surfactants (Porter 1991), and they are generally less
toxic and less affected by electrolyte concentration than are anionic and cationic surfactants (Porter 1991;
Myers 2006).

Figure 6.20 An aircraft releases chemical dispersant over an oil slick in the Gulf of Mexico in 2010. Source:
NOAA /US Coast Guard

Figure 6.21 is a schematic diagram showing the spraying of a dispersant onto a floating slick and the
slick breaking apart into small droplets that get driven into the water column by wave action. The water-
loving end of the surfactant is attracted to the water while the oil-loving end is attracted to the oil. The oil
droplet is surrounded by the surfactant molecules, forming what is called a micelle or a surfactant-
stabilized oil droplet. Eventually, as water mixes with the droplets, their concentration declines and the
bacterial communities in the sea are able to break down the oil quickly, leaving no lasting residue that
might exert toxicity to the seawater biota.

6
Amphiphilic is a term describing a chemical compound possessing both hydrophilic (water-loving, polar) and lipo-
philic (fat-loving, non-polar) properties.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 219
Figure 6.21 Schematic diagram of how oil is dispersed by dispersants.

A typical measure of the relative importance of the hydrophobic and hydrophilic characteristics of
nonionic surfactants is the hydrophilic-lipophilic balance (HLB), which ranges from 0 for completely
lipophilic or hydrophobic molecules to 20 for completely hydrophilic (charged) molecules. Figure 6.22
shows the HLB scale ranging from highly oil-soluble to highly water-soluble. Surfactants with low HLB
tend to stabilize water-in-oil emulsions, whereas those with high HLB stabilize the more desirable oil-in-
water emulsions (NRC 1989; Clayton et al. 1993). Commercial dispersants tend to have overall HLBs in
the range of 9 to 11, which is often achieved by combining surfactants having higher and lower HLB.
Although the industry consensus suggests that combining surfactants with different HLB improves
dispersant effectiveness (NRC 1989; Clayton et al. 1993), others have offered alternative findings (Fingas
et al. 1990a).

To date, most chemical dispersants have been formulated for marine use, as the dissolved divalent cations
in seawater (particularly calcium and magnesium) are integral to dispersant efficacy (Chapter 3).
Development of freshwater-compatible or -specific dispersants has lagged mainly because of regulatory
decisions against the use of dispersants near water intakes. A fundamental laboratory study was
conducted in 2008 to determine if it is possible to produce a dispersant for use in fresh water (Wrenn et al.
2009). The authors found that dispersants can be designed from scratch (rather than modifying existing
commercial dispersants) to drive an oil slick into the freshwater column with the same efficiency as in
saltwater as long as the HLB is optimum for that environment.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 220
Figure 6.22 Hydrophilic-Lipophilic Balance diagram showing the HLB scale. (Source: MHK;
https://en.wikipedia.org/wiki/Hydrophilic-lipophilic_balance. This figure is licensed under the Creative Commons
Attribution-Share Alike 3.0 Unported license)

The effect of using dispersants on an oil slick is not simply to dilute the oil. Rather, creating the
small oil droplets and driving them into the water column increases the surface-area-to-volume ratio,
thereby inducing significantly faster biodegradation by oil-degrading microbial communities within
the water.

The purpose of the solvent in a dispersant is to provide the surfactant mixture with an appropriate
viscosity, which ensures that it can be pumped through spray nozzles at environmental temperatures. The
solvent may be water-miscible7 (e.g., 2-butoxyethanol) or immiscible (e.g., normal alkanes) (NRC 2005).

Mixing energy driven by wind and wave action is required to achieve effective dispersion. Depending on
a site’s energy level, tiny droplets of oil (ideally < 70 µm in diameter) are mixed in the upper meter of the
water column creating a subsurface plume. This plume of dispersed oil droplets rapidly mixes and
expands in three dimensions (horizontal spreading and vertical mixing) down to as much as 10 meters
below the surface. The rise speed8 of these droplets is extremely slow and is overcome by the turbulence
in the upper depth of the sea, essentially resulting in neutral buoyancy of the droplets. The ideal size of
the droplets precludes re-coalescence and re-formation of a surface slick. As a result of this mixing, oil

7
Water-miscible solvents mix completely with water to form a single phase, whereas immiscible solvents form two
phases.
8
Rise speed is the rate at which submerged droplets will resurface (similar to the concept of terminal velocity for
objects falling through air), which depends on droplet size and coalescence. Large droplets may take several hours
or days to resurface, whereas small droplets may remain submerged for a long time regardless of their nominal
density and buoyancy.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 221
concentrations decrease rapidly from the initial peak concentrations, e.g., from 10 or 100 ppm down to 1
ppm or less, within hours (Lee et al. 2013).

Microorganisms act on the small dispersed droplets at the oil-water interface (Figure 2.3), so increasing
this surface area by reducing the droplet size increases the area for biodegradation to occur. Thus, the
ultimate fate of oil is its destruction. The transient effect might be negative to fish and other waterborne
animals (Chapter 4), but such an effect may be short-lived due to rapid dilution of the oil to
concentrations below toxicity threshold limits, enhanced oil degradation and resilience of the living
populations. Dispersants and dispersant applications are rarely 100% effective, however, so some oil will
likely remain floating on the water surface.

Dispersant effectiveness is limited by certain


In some cases, dispersing oil changes the trajectory physical and chemical variables, such as sea
of the oil plume from onshore to alongshore, as the state and oil properties. Sea state has already
transport of dispersed oil is less affected by the wind. been discussed briefly above. When wave
Therefore, oil dispersion may help protect sensitive energy is sufficiently high, re-coalescence of
shoreline environments, as wind usually is the the small dispersed oil droplets is minimized.
dominant environmental factor that carries floating Salinity is also an important variable.
oil to strand ashore. Dispersants in use today have been
optimized for use in seawater (salinity of ~35
g/kg). Lower salinities substantially lessen
dispersion effectiveness of commercial products unless the dispersant has been manufactured to work in
brackish or fresh waters. This is accomplished by manipulating the HLB as discussed previously.

The discussion above would not be complete without mentioning the 2010 DWH blowout in the Gulf of
Mexico. In that spill, at least 430,000 to 500,000 tonnes of light oil gushed out of a deepwater well for
almost three months before the well was finally repaired and the blowout contained (McNutt et al. 2012;
Fingas 2013). Approximately 3,000 tonnes of two chemically similar anionic dispersants (initially Corexit
9527, followed by Corexit 9500, which lacks the miscible solvent 2-butoxyethanol) were injected into the
subsurface oil plume as it escaped the well’s riser tube and a further 4,000 tonnes were applied to oil at
the sea surface (Gray et al. 2011). Significant amounts of data on fluorescence, droplet size distribution,
dissolved oxygen, total petroleum hydrocarbons (TPH) and volatile organics analysis (VOA)
measurements were collected by the U.S. Coast Guard, NOAA, U.S. EPA, BP, DFO and other
organizations deployed in the area and were published daily. Data were obtained from vessels deployed
on the surface of the water lowering instruments and samplers to the vicinity of the well and dispersed oil
plume, plus autonomous vehicles taking real-time data. These data were important to understanding the
hydrodynamic behaviour of the water and the direction of the currents carrying the plume of dispersed oil.
Initial findings suggested that the deep subsurface dispersed oil did not experience significant buoyant
rise or sinking after release and instead behaved as neutrally buoyant particles or as dissolved materials in
the water. All of the measurements indicated that the majority of the deep subsurface dispersed oil was
carried predominately to the southwest away from the wellhead by the prevailing currents at depths of
900 - 1,300 m. Analysis of TPH, VOA and coloured dissolved organic material (CDOM) fluorescence
data (Appendix A) indicated the presence of deep subsurface dispersed oil up to 100 km northeast of the
well.

Measurements of the dispersed oil plume in the deep sea during the DWH spill at 1,100 to 1,300 m below
the surface (Figure 6.23) showed that most of the plume consisted of particle sizes ranging from 2.5 to 60
µm in diameter. In the same area, a dissolved oxygen anomaly was observed signifying biodegradation
activity. Furthermore, light scattering measurements in the LISST particle size analyzer of water samples
collected from that plume displayed the characteristic bimodal distribution typical of chemically- as
opposed to physically-dispersed oil (Li et al. 2008, 2009). This suggests, but does not scientifically prove,
that the oil in the deep zone was likely chemically dispersed at least part of the time due to the injection of

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 222
dispersant into the oil exiting from the riser tube (injection of dispersant was not always active). Although
it is plausible that the extreme turbulence of the oil as it gushed from the well may have caused extensive
physical dispersion without the need for chemical dispersant use, review of reported data and information
in the literature suggested that it was less likely and that the application of dispersants in the deep sea was
successful in dispersing the oil at the source (Venosa et al. 2012). Later, however, Valentine et al. (2014)
reported finding oil contamination in the sediment downstream from the well at 1,500 m water depths.
The amount of contamination is still being investigated. For a detailed discussion of toxicological issues
associated with the use of dispersants in oil spill response, please refer to Chapter 4.

In addition to the dispersant injected at depth into the oil spewing from the wellhead, significant
quantities of dispersants were used on the Gulf’s surface during the DWH response. Some of the oil that
made it to the surface migrated to coastal habitats. The use of dispersants likely increased the
bioavailability of the oil (its ability to be taken up or acted upon by organisms) and enhanced the
opportunities for biodegradation. However, dispersants also may have increased the potential exposure of
sensitive organisms. The fate of the released oil and its many components remains poorly understood.
This is an active area of research among oceanographers, engineers and marine chemists. Certainly, the
use of dispersants at the sea surface reduced the exposure of the coastline to oil and application of
dispersants at the wellhead reduced the amount of oil reaching the surface, but effectiveness of both the
surface and the deep sea dispersion was not quantified.

Dispersants have been used successfully as an effective oil spill response technique. During the DWH
blowout in the Gulf of Mexico, dispersants were applied in deep water to the point of release for the first
time during an actual response operation. Although many have concluded that dispersants were successful
during the DWH spill, it is still unknown for certain to what extent the oil was dispersed physically due to
the high velocity of oil emerging from the well into the water or chemically in response to mixing with
applied dispersant.

Figure 6.23 Data collected during the Deepwater Horizon blowout showed clearly that at approximately 1200 m
below the surface, the dissolved oxygen spiked downward (red curve) suggesting biodegradation while the
fluorescence spiked upward signifying the presence of the oil droplets in that same area. Source: K. Lee, DFO

Recommendation: The relative importance of turbulence and chemical dispersion needs to be


measured in controlled experiments to reduce the uncertainty of the chemical dispersion treatment
strategy. Controlled empirical experiments, preferably in the field, need to determine the

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 223
persistence of dispersed oil in the deep sea and at the sea surface. The ecological and toxicological
consequences of dispersants and dispersed oil in wetland, salt marsh, mudflat and deep sea
environments need to be evaluated at cold and warm temperatures, in part by using mesocosms
and wave tanks. Analytical chemical techniques for monitoring dispersed oil or dispersants in the
water column at extremely low concentrations need to be improved, especially in fresh water.
Research is needed to develop, if practical, effective and non-toxic biosurfactants as alternatives to
traditional chemical dispersants currently on the market.

6.3.4 Surface washing by chemical washing agents

Surface washing removes free product stranded on the shoreline without removing sediment. Collected oil
is placed in containers and carried off the shoreline. This process is most effective on sand and gravel
beaches and riprap9, but can also be used on other types of terrain except soft mud substrates and mud
flats, where the oil would be pushed downward into the sediment causing a worse condition than before.
Surface washing is most useful for light to moderate oiling by medium to heavy oils. Removal of surface
oil is typically used only along the edges of low, sheltered or vegetated riverbanks and marshes and must
be closely monitored.

6.3.5 Solidifiers

Solidifiers are chemical agents applied as powders, granular mixes or gels that react with oil with minimal
volume increase to form a cohesive, solidified mass that floats on water (Ghalambor 1997). Use of
solidifiers for oil spills on water has been investigated since the early 1970s (Dahl et al. 1997). In some
cases, polymeric solidifiers are modified with other inorganic or organic additives, which serve as
chemical bonding agents. Due to the need for recovery and disposal of the immobilized oil, solidifiers are
generally impractical for wide-scale application to large spills and have been only used on small oil spills
on land or restricted waterways (Schulze and Hoffman 1993).

Solidifier booms have also been shown effective in removing sheen from wastewater holding ponds. The
U.S. EPA Region 4 authorized solidifiers to be used as an alternative to sorbents or mechanical recovery
when removing small or thin sheens from water or small amounts of oil from land (Michel et al. 2008).
As reviewed by Fingas (2013a), a solidifying agent manufactured by BP was tested by the Canadian
Coast Guard for treating oil-under-ice in the Beaufort Sea and oil in open sea conditions off
Newfoundland, with the conclusion that the product’s usefulness was equivocal in the first case and not
practical in the second case because of the large mass ratio of solidifier:oil required. Hand-spraying the
product onto an experimentally oiled Arctic marine shoreline on Baffin Island similarly yielded little
benefit (Fingas 2013a). Various other solidifiers have been laboratory-tested by Environment Canada.

Use of solidifiers for oil spill response has some drawbacks and thus has received little attention. For
example, their effectiveness depends on the type and composition of the crude oil tested (Fingas et al.
1990b; Dahl et al. 1997). Another disadvantage has been the large amount of material that needs to be
applied and recovered. It has been reported that 16 to over 200% by weight (solidifier-to-oil mass ratio) is
required to solidify crude oil (Fingas et al. 1995b). Moreover, practical application methods and
appropriate tests under various conditions and environments have been lacking (Delaune et al. 1999). In
recent reviews of oil spill solidifiers, Fingas (2008, 2013) noted a lack of rigorous scientific assessment of
their effectiveness at spill sites. Temperature may play an important role in the solidification of oil mainly
because of the variations in viscosity and API gravity of the petroleum product. At low temperatures
solidifiers have been found to be less effective (Fingas 2008, 2013a). In contrast, Pelletier and Siron

9
Riprap is rock, concrete or other material used to armor shorelines, streambeds, bridge abutments, pilings and
other shoreline structures against scour and water or ice erosion.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 224
The effectiveness of a solidifier product is dependent upon a number of incident-specific variables,
including oil type, oil amount and weather conditions, such as the sea state, sea temperature and
wind conditions.

(1999) tested a silicone-based solidifier on a light crude oil and reported that temperature did not appear
to have a major impact on effectiveness, although more testing would be required to confirm this finding.

In their review of environmental considerations for non-dispersant chemical countermeasures, Walker et


al. (1999) reported that the effectiveness of solidifiers probably decreases for emulsified, weathered, thick
and heavy oils because of the difficulty of mixing the product with viscous liquids. They also reported
that salinity does not affect the ability of solidifiers to solidify oil on water.

The density and viscosity of the oil reflect the chemical and physical variations of crude oils. Changes in
viscosity depend on environmental conditions, particularly temperature. Viscous oils tend to clog the
outer layer of the solidifier, thus decreasing further sorption of oil. Fieldhouse and Fingas (2009) reported
that product performance was similar at 0 and +15 °C, but at -15 °C longer contact times were required
for the samples to reach a state sufficiently solidified to prevent release of oil from the solidified material.
The study also reported that low temperature was a limiting factor for generating a solidified product,
while at room temperature, the treated material became a cohesive mass. Low temperature was also found
to result in longer solidification times or reduced effectiveness due to increased oil viscosity.

Sundaravadivelu et al. (2015) recently conducted an intensive, statistically rigorous laboratory study of oil
solidifier performance using a variety of products and oil types at 5 and 22 °C. The removal efficiency
and final consistency of the product depended on oil type, temperature, bulk density of the product and
application ratio. Products with lower bulk powder density had high specific surface areas, which
resulted in increased oil removal efficiency. For these products, the efficiency was significantly higher at
22 °C than at 5 °C (by 10 to 20%). However, variability was lower as a function of temperature for
products with higher bulk density and less removal effectiveness (only 4% increase at the higher
temperature). Overall, the best predictor of solidifier effectiveness was found to be its bulk density.

Recommendation: A laboratory protocol is needed to compare and test effectiveness of commercial


solidifier products. Research is needed to determine whether solidifiers are effective or feasible to
use on larger oil spills, and whether the product and oil can be recovered after such use.

6.3.6 Herding agents

Herding agents gather thin oil slicks together or move them (at the surface) towards a desired location for
collection or ISB, particularly in partial ice cover (Fingas 2013b). Buist et al. (2008) sprayed a herding
agent onto the seawater surface surrounding an experimental oil slick, which resulted in the formation of
a monolayer of surfactants on the water surface, a process first reported by Garrett and Barger (1972).
When the surfactant monolayer reaches the edge of a thin oil slick, it changes the balance of interfacial

A herding agent is a chemical containing a surfactant and a solvent, similar to a dispersant but in
different proportions with different functions.

forces acting on the slick edge and causes the oil to contract into thicker layers. Since herders do not
require a boundary to ‘push against’, they work even in open water.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 225
Surfactants considerably reduce the surface tension of the water surrounding an oil slick (from about 70
mN/m to 20 to 30 mN/m10). Although commercialized in the 1970s mostly to concentrate the oil for
skimming and recovery, herders were not used offshore because they worked only in very calm
conditions. When wind speeds were 2 m/s or greater, booms were needed to contain the slicks, and
breaking waves were observed to disrupt the herder layer. A recent strategy in loose pack ice is to herd
freely-drifting oil slicks to a burnable thickness (i.e., greater than about 3 mm). Once a sufficient
thickness is attained, the oil may be ignited for ISB. The drift ice acts as an additional containment
barrier to aid in reducing the time needed for the development of a slick thick enough to burn. Buist et al.
(2008) succeeded in burning oil in drift ice using a non-proprietary hydrocarbon-based cold water
herding agent (called USN).

Based on earlier studies (Buist et al. 2008; 2010a,b), work has continued on herding agents for ISB in
Arctic waters (Buist et al. 2011). Recent herding agents were capable of thickening a slick to > 4 mm,
suitable for ISB (Buist et al. 2010b), and when tested on experimental oil spills, > 90% of the oil was
combusted. Based on the results of earlier work (Buist et al. 2008), the researchers concluded that herding
agents work best on relatively fresh crude oil and light distilled product slicks that are still of relatively
low viscosity and remain ignitable. Slicks that have gelled or significantly emulsified and viscous residual
fuel oil slicks would not be good candidates for herder use. As with most oil spill response techniques,
rapid response improves the chances of success when using herders for ISB.

Recommendation: The utility of herding agents, especially for use in the Arctic, needs to be
quantified and properly tested to provide a scientific basis for their application to oil spills.

6.3.7 Debris and detritus removal

All spill incidents involve the production of huge quantities of recovered oil, wastewater, debris, detritus,
heavy metals and other solid materials during the spill response that are usually disposed in landfills.
Heavy metals are especially important because of their toxicity and persistence in the environment. These
cleanup materials include sorbents, debris (seaweed, trash, logs, mooring lines, etc.) and clothing (Figure
6.24). Interim storage and permanent disposal options have to be defined. It should be noted that disposal
of contaminated materials in landfills simply moves the pollution from the shoreline to another site.
Accordingly, the work plan should include provisions for reducing the volume of material to be disposed
in landfills and should address options, such as recycling, using waste oil in road asphalt and removing
excess liquid from the waste oil to decrease volumes. Oiled debris removal is recommended for sand
beaches, gravel beaches, sheltered rocky shores, man-made structures and sheltered rubble slopes. It is
conditionally recommended for exposed rocky shores, tidal flats, sheltered vegetated low banks and
marshes (Region 10 Regional Response Team and the Northwest Area Committee 2015).

In addition, incineration is one of the most efficient methods of disposal for recovered oil, used oil
sorbents and debris in a relatively short time and involving not much labour force. The general
disadvantage of incineration is high transportation cost of the disposed material to the incinerator facility
and regulatory prohibitions in many Canadian jurisdictions. A recent comprehensive review of all the
methods used for treating oily sludge in the petroleum industry was published by Hu et al. (2013),
including how to deal with heavy metals and hazardous liquid waste.

Recommendation: The change of heavy metal concentrations during different oily sludge treatment
processes is worthy of investigation in future research, and the integration of available heavy metal
control technologies into oily sludge treatment is recommended.

10
mN/m (millinewtons per metre) is a unit used to measure surface tension at an interface.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 226
Figure 6.24 Detritus removal during the Motiti Island oil spill off the coast of New Zealand, 2011. Source:
Maritime New Zealand

6.4 Factors Affecting Spill Response and Cleanup Effectiveness

6.4.1 Oil types and properties

Very light oils (e.g., jet fuels and gasolines) are highly volatile
(evaporate in one to two days) and have high concentrations of
The types and properties of toxic soluble compounds. Ventilation and/or absorption are
spilled crude oil are among the usually effective cleanup methods in conjunction with fire and
most significant factors inhalation prevention measures. Light oils (e.g., diesel, No.2 fuel
influencing the efficiency and oil, and light crudes, including Bakken crude) are relatively
effectiveness of oil spill response. volatile and thus easier to treat effectively than heavier oils.
Medium oils (e.g., most conventional crude oils) are generally
less volatile than lighter oils, increasing the potential for impacts
on waterfowl and wildlife. Heavy oils and diluted bitumen (e.g., Bunker C, weathered dilbit and railbit)
are receiving increasing attention due to their resistance to evaporation and biodegradation and their
potential long-term damage to the environment, waterfowl and fur-bearing animals. They have fewer
components that dissolve in water (Kok 2011; Dejhosseini et al. 2013). Cleanup is extremely difficult for
both marine and inland spills. For detailed information about crude oil types and properties, refer to
Chapter 2 and Appendix B.

The main physical properties that can affect the behaviour of oil are density, viscosity and flash point
(Fingas 2012; Yasin et al. 2013). Oils with high API gravity tend to be free flowing (low viscosity), have
a high proportion of volatile compounds and low flash points and pour points, which facilitate their
cleanup. However, the properties of oils with high wax or asphaltene content cannot be solely determined
by API gravity. Viscosity increases as the oil weathers (Demirbas et al. 2015). Volatility refers to how
quickly the oil evaporates into the air. High volatility usually can increase the amount of oil that
evaporates into the atmosphere. Flash point is the lowest temperature at which sufficient vapour exists
above the spilled oil to yield a flammable mixture (Gülüm and Bilgin 2015). Light crude and oil products
have low flash points that are suitable for ISB.

The fate and behaviour of the spilled oil will influence response effectiveness. Abiotic processes at water
surface and within the water column include spreading, evaporation, emulsification, oxidation, dispersion,

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 227
dissolution and sedimentation (see details in Chapter 2 and 3). Biotic processes in the water column
usually refer to biodegradation and uptake by living microorganisms. Emulsification (high viscosity) and
sedimentation can impede the effectiveness of response techniques (Cai et al. 2014b). Weathering
processes that influence changes in oil properties can be significantly influenced by ambient
environmental conditions. Thus, comprehensive consideration of the above oil types and properties and
their weathering processes, along with environmental factors, are necessary for making sound response
decisions and actions.

6.4.2 Environmental and ecological factors

6.4.2.1 Weather

The feasibility and effectiveness of oil spill response are heavily dependent on the weather conditions
prevalent at the oil spill site. Generally, a clear, sunny day without strong wind would be favoured by spill
responders, especially in offshore spill sites (Fingas 2012; Doerffer 2013). Such weather conditions
would encourage natural attenuation (e.g., evaporation), help the responders deploy equipment and
enhance the efficiency of booming, skimming, ISB and dispersant application (Verma et al. 2013).

In hot or warm weather and water conditions, evaporation usually dominates as higher temperature tends
to increase the volatility of spilled oil, especially for lighter products. The viscosity of water-oil emulsions
decreases with increasing temperatures. Therefore, mechanical cleanup techniques, such as booming and
skimming, are much easier to deploy and more effective in warm conditions. Temperate conditions also
favor ISB due to the ease of igniting the oil slick. While indigenous microorganisms are adapted to their
environment, their rates of activity are generally enhanced at higher temperatures in both the water
column and the sediment. Oil biodegradation rates and extent are generally higher in warm water
environments. For example, an increase of 10 °C in temperature can double the metabolic rate of an
animal, plant or microorganism (Tyagi et al. 2011; Hassanshahian et al. 2012). The bacterial metabolic
rate has been reported to be on the order of 800% higher in the Gulf of Mexico than in Alaska (Mortazavi
et al. 2013).

In contrast, oil spill cleanup is usually complicated by cold and harsh climatic conditions such as those
that prevail in northern Canada and the Arctic. Deploying response equipment and personnel is
challenging not only because of these harsh conditions but also the lack of access and the long distances
between the source of cleanup equipment and the spill. Physical constraints include the fact that during
winter months, it is often too cold to work and daylight hours are limited (or non-existent). In addition,
cold weather may cause health and safety issues due to factors such as potential hypothermia among
responders, brittle failure of metal components, freezing of sea spray causing vessel instability, etc. The
recovery of oil in extreme cold weather requires a different approach than in warm weather and the
likelihood of injury to responders increases significantly (Daling et al. 2010; Buist et al. 2011). Different
sets of skills and technologies are needed by responders to ensure effective spill response in harsh
environments (Oskins and Bradley 2005).

Cold conditions may also compromise the effectiveness of mechanical containment and other cleaning
efforts. For example, mechanical oil recovery, such as through the use of skimmers and pumps, may be
compromised because the equipment may freeze up and the increase in oil viscosity reduces recovery
efficiency. In the case of ISB, ignition may be difficult and burning may be slow and inefficient. In
addition, considerable debate continues over the effectiveness of dispersants on crude oil degradation at
low seawater temperatures (Daling et al. 2010). The main concern is that as the temperature decreases,
chemical and biological processes slow down and oil viscosity increases, making dispersion of oil more
difficult (EPPR 2015). In contrast, reduced oil spreading rates extend the ‘window of opportunity’ for
dispersant use (Nuka Research 2010; Arctic Oil Spill Response Technology Joint Industry Programme

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 228
2013), followed by a reduction in the size of the chemically-dispersed oil droplets, thus enhancing the
potential for effective biodegradation in a cold marine environment (Cai et al. 2014a; Greer et al. 2015).

In summary, the effects of weather conditions on oil spill response include:

Warm weather

 Evaporation dominates (higher temperature increases volatility and decreases oil viscosity),
resulting in better use of mechanical and ISB responses; and
 Microbial communities grow and metabolize hydrocarbons faster at higher temperatures,
speeding up biodegradation.

Cold weather

 Deployment of response equipment and personnel is challenging due to lack of access and
long distances to the spill site;
 Daylight hours may be limited or plentiful depending on the latitude of the spill site and
season;
 Injury potential to responders is heightened due to the harsh conditions;
 Different skill sets and equipment are needed (chain saws and ice augurs);
 Crude oil biodegradation rate is much lower at low temperatures;
 Oil viscosity increases, which reduces recovery efficiency;
 Cold temperatures may compromise the effectiveness of mechanical equipment;
 ISB ignition may be difficult and burning may be slow at cold temperatures;
 Reduced oil spreading rates extend the opportunity for dispersant use and reduce the size of
oil droplets;
 Cold conditions may cause health and safety issues to responders; and
 Mechanical equipment may freeze up.

6.4.2.2 Wave height

Wave height is an important consideration for any recovery method that relies on containment before
collection or removal. Waves can impact mechanical spill response by the following factors (Fingas 2012;
Andrade et al. 2013):

 Causing boom and skimmer failure or reducing their effectiveness;


 Making vessels difficult to keep on station;
 Creating unsafe conditions for crew to work on deck;
 Making deployment or retrieval of equipment challenging or impossible;
 Causing oil to submerge so that it is no longer available for recovery; and
 Limiting the ability to track and encounter oil.

Oil is often entrained beneath or splashed over booms in short-period wind-waves exceeding 1 - 1.5 m,
although booms can accommodate higher wave heights in a long-period ocean swell (Deng et al. 2013).
The issue of boom limitations in high waves also affects the practicality of using ISB in rough sea
conditions (Gong et al. 2014). Steep, wind-driven waves typically impart a more significant detrimental
effect on spill response operations than longer period swells. In addition to driving sea state conditions,
wind can also impede on-water mechanical response through direct limits to both operating vessels and

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 229
equipment (Wang et al. 2012; Li et al. 2014a). These conditions can be problematic in harsh northern
Canadian and Arctic regions, as well as large inland lakes, such as those in southern regions of Canada
and the Great Lakes, which can have severe wave heights during rough weather. However, it should be
noted that many Arctic areas experience less severe wind and sea conditions than most open seas and thus
may not significantly experience these problems (Dickins 2011). This is because the regional presence of
ice dampens wave action and often limits the fetch over which winds might otherwise create larger, fully
developed waves.

In summary, extreme wave height and wind can produce unsafe conditions to operate boats, vessels and
aircraft. They may also combine with low temperature to further cause wind chill (Gong et al. 2014). For
mechanical recovery, these conditions can move booms and skimmers away from oil slicks or make them
less or completely non-functional. In regards to ISB, they may impede oil containment and ignition and
compound unsafe conditions. Dispersant spray application, use and monitoring are also hampered by
these energetic conditions. However, high waves and wind do offer positive effects, such as potential oil
movement away from sensitive areas (if in a right direction) and extra mixing energy, promoting natural
dispersion and evaporation (Nuka Research 2010).

6.4.2.3 Ice conditions

In an environment with static ice and snow conditions, the first response timelines may be lengthened if
oil remains contained, concentrated and trapped for extended periods of time. In dynamic ice conditions,
the ‘window of opportunity’ for spill response operations
can be extended if the oil can be tracked until it appears
Existing trajectory models are limited in
naturally on the surface or until vessels and crews can
their capability to model oil fate and
access it at some distance from the original spill site
behaviour in the presence of a range of
(EPPR 2015). Tracking oil in or under ice and snow
sea ice conditions.
remains a challenge for in situ and remote monitoring
technologies. Furthermore, the choice for spill response
strategies is limited in harsh conditions, such as those encountered offshore in moving pack ice and rivers
in late winter and early spring when they are high and often over-topping river banks and ice floes are
moving rapidly or piling into ice jams that cause even more flooding.

Conventional booms and skimmers for mechanical containment and recovery of spilled oil become
increasingly ineffective as ice concentrations increase much beyond 10% or more (Lampela and Jolma
2011). For example, ice may tear or move containment booms or clog skimmers or reduce their efficiency
and maneuverability. The length of boom that can be deployed and maintained under freezing conditions,
even for a short period, depends on the severity of the ice conditions. Limited effectiveness is still
possible in very open drift ice (10 - 30% ice cover) and in isolated polynyas11 within closer pack ice
(Wilkman et al. 2014). However, even the presence of small fractions of ice interferes with boom
operation and quickly reduces flow to the skimmer head. Any limited containment of oil, which may be
possible in open drift ice, requires rugged, high strength booms to withstand contact with the ice. In
addition, ice may also create difficulties for deploying fire booms and reduce ISB effectiveness.

The fundamental limitations associated with maintaining and operating booms and skimmers in ice are
further complicated in polar areas by the lack of coastal infrastructure and approved on-land storage sites
(Pilipenko et al. 2013; Schmidt et al. 2014). The presence of ice may also preclude spill response vessels
from getting close to the spill site or to track spilled oil. Due to downtime associated with the clogging of
water intakes on vessels, additional ice scouts or ice management vessels may be needed during response
operations (Wilkman et al. 2014).

11
A polynya is an area of open water within the sea ice pack.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 230
Although ice conditions preclude the effectiveness of booming
To date, dedicated mechanical and skimming, many field observations and experiments have
systems for operations in pack ice shown that partial ice cover can also act as a barrier preventing
have not progressed beyond the oil from spreading as it forms a natural containment system.
prototype stage. With high ice concentrations and a fringe of fast ice, spilled oil
may be rapidly encapsulated within the ice leading to thicker
oil films and making burning, skimming and dispersant
application more effective (Dickins 2011). Conversely, ice cover can protect the oil from weathering
processes, such as evaporation, dispersion and emulsification, and reduce the mixing energy that is
needed for effective chemical dispersion (Nuka Research 2010). Although the deployment of fire booms
to support ISB operations may be hampered by the presence of ice, it may not be necessary as the ice
itself may also act as a boom.

NRC (2014) recently released a comprehensive review of the current state of the science regarding oil
spill response and environmental assessment in the Arctic region north of the Bering Strait. A committee
of industrial and governmental organizations tasked the NRC to review research activities and
recommend strategies to advance research and address information gaps, to identify opportunities and
constraints for advancing oil spill research, to describe promising new concepts and technologies, and to
assess the types of baseline information needed to monitor the impacts of an oil spill and to develop plans
for recovery and restoration. The report reviews oil spill countermeasures applicable to the Arctic,
including bioremediation, dispersants, ISB, chemical herders, mechanical containment and recovery,
detection and tracking and trajectory monitoring.

The report identified seven research and development needs for improved decision support, including
those listed below:

 Improving methods for ISB, dispersant application and use of chemical herders;
 Understanding limitations of mechanical recovery in both open water and ice;
 Investing in under-ice oil detection and response strategies;
 Integrating remote sensing and observational techniques for detecting and tracking ice and oil;
 Determining and verifying biodegradation rates for hydrocarbons in offshore environments;
 Evaluating the toxicity of dispersants and chemically-dispersed oil on key Arctic marine species;
and
 Summarizing relevant ongoing and planned research worldwide to achieve synergy and avoid
unnecessary duplication.

Recommendation: Initiation of multi-disciplinary research programs is needed to address specific


questions related to cold and harsh environments, such as developing sufficient knowledge on the
fate, behaviour and impacts of different oil types (in various weathering/treatment states) in snow
or ice conditions. Special training of personnel and development of equipment for harsh
environments or extreme weather are needed. Better understanding of the ecological effects in cold
environments of various oil spill cleanup measures would be beneficial to spill response.

6.4.2.4 Available daylight and visibility

Daylight duration and intensity directly affect oil weathering processes. A series of field spill experiments
in the Russian Arctic showed that photooxidation of spilled oil was a more significant process in the 24-
hour summer daylight than in more temperate climates (Serova 1992; Ivanov et al. 2005). However,
shorter daylight and lower solar energy in fall and winter, as well as the associated low temperatures,
can reduce evaporation and biodegradation activity by microorganisms.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 231
Visibility is an important safety factor during mechanical response operations in the field. Low horizontal
visibility can impede or prevent on-water mechanical recovery by making it difficult to see and track oil
slicks and by making vessel or air support operations unsafe. Low vertical visibility (cloud ceiling) limits
aerial observation from rotary or fixed-wing aircraft. Darkness precludes some aspects of mechanical
recovery and makes it infeasible to conduct ISB or aerial dispersant spraying (Nuka Research 2010;
Jacobs et al. 2015). Darkness also increases risk of accidents and makes it more difficult to deal with
emergencies, such as a man overboard (Genwest 2012; Salt et al. 2012).

6.4.2.5 Ecological considerations

Ecological concerns tend to become a major driving force for spill contingency planning and risk
management (U.S. Coast Guard 2001; Aurand and Coelho 2006a,b; Webler and Lord 2010). A drive is
underway to improve understanding of the ecological effects of various oil spill cleanup measures to
improve operational spill response guidelines. This need is highlighted by recent studies that provided
evidence of negative ecological impacts associated with the application of chemical dispersants in the
DWH oil spill (Goodbody-Gringley et al. 2013; Kleindienst et al. 2015) and spill mitigation techniques
used in the August 10, 2000, Pine River, BC, Pembina pipeline rupture. The environmental impact in the
latter spill included mortality to fish, insects and some wildlife. Moreover, the District of Chetwynd, BC,
had to discontinue use of the river, as well as many groundwater wells near the river, as a source of
potable water. See more discussion in Chapters 4 and 8.

It will take time to fully understand and quantify the impacts of dispersants in order to create best practice
guidelines. Nevertheless, some mitigation measures that should be considered include: using
mathematical tools to optimize the dose and operational processes or developing more potentially eco-
friendly dispersants based on biosurfactants12 (Zhong and You 2011; Cai et al. 2014b).

Conversely, significant research and practical efforts have been made in bioremediation and a
considerable amount of new knowledge has been developed. For example, indigenous microbial species
and nutrient levels are crucial factors that are often taken into account during design and application of in
situ bioremediation due to their impacts on oil fate and behaviours (JRP 2013). Consideration of
ecosystem vulnerability is discussed further in Chapters 7 and 8.

6.4.3 Technical and economic factors

A number of technical factors need to be taken into account when arranging and deploying response
equipment (Table 6.2). The specific operational conditions for each device are usually suggested by
manufacturers and/or determined by responders through drilling and field application, although some
special or uncommon environmental conditions (e.g., Arctic and extreme weather conditions) are not well
addressed mainly due to lack of sufficient knowledge or tests. However, the feasibility of adopting a
specific response technique depends on the proximity of equipment needed at the spill site and the time
required to transport it there (Li et al. 2014b). In addition, responders’ knowledge, experience and skills
should also be considered when choosing the technique or equipment (U.S. Coast Guard 2001). Proper
training of the response personnel can greatly increase response efficiency and reduce the possibility of
having health or safety issues.

12
Biosurfactants are surface-active chemicals produced by organisms. They may have more selective activities than
manufactured (synthetic) dispersants and are often biodegradable.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 232
Table 6.2 Examples of oil spill response equipment and their affecting factors. Sources: Transport Canada (2013)
and EPPR (2015)
Equipment Affecting Factors
Command equipment Weather and sea conditions
Dispersant systems Wave, ice and oil types/weathering degree
Air-inflatable booms Wave, ice and wind
Skimmers Wave, ice and oil types/weathering degree
Temporary storage equipment Capacity
ISB devices Weather and sea conditions, wind
Vessels Capacity and compatibility with onboard devices
Pumps and generators Efficiency and power
Subsea well intervention equipment Water pressure and sea conditions

For persistent spills, especially caused by heavy oils and diluted bitumen, the cleanup cost may include
long-term soil/shoreline remediation and groundwater restoration expenses. The lack of adequate support
funds may result in inappropriate or untimely response (Fingas 2011).

It is usually required by law that employers provide a safe and healthy workplace free of recognized
hazards, as well as health and safety plans for response crew. Employers must also provide training
and required protective equipment to enable the application of different response techniques
(Michaels and Howard 2012). Therefore, the conditions at the spill site and the logistic capacity
(e.g., in remote locations or under harsh weather conditions) can affect decisions about the most
appropriate set of techniques/devices.

A commonly encountered question, and often the most difficult issue to address, is “how clean is clean?”,
which for the most part is a site-specific question dependent largely on the species at risk. It also invokes
question about the intended end use of the spill site after cleanup: whether the site will be restored,
remediated, rehabilitated or reclaimed (decision categories more commonly applied to terrestrial or
shoreline sites than to open water) 13 . Michel and Benggio (1999) reported how to select the cleanup
endpoints to meet the spill-specific requirements. Generally, cleanup endpoints need to be determined in
the early stage of response according to general information and knowledge of the spill site and oil
type/properties (Sergy and Owens 2007). Ongoing monitoring is necessary to allow for the modification
of endpoints or techniques used, if required. Cleanup should proceed as long as it can accelerate the
recovery of the spill site and should be terminated if it would slow down the recovery or have no value
(Sergy and Owens 2008). The judgement of clean conditions depends on the availability and quality of
baseline data that are sometimes missing or incomplete (refer to Chapter 7).

6.4.4 Other factors

Response structure and communication may also influence the response effectiveness given that response
personnel, such as onsite response crew, offsite decision makers, technical consultants and other
stakeholders, need to communicate and follow up with guidelines and regulatory structure for

13
Restoration returns the site to the condition pre-dating the spill; remediation removes or reduces (hazardous)
wastes from the site to minimize current or future adverse effects; rehabilitation returns the site to a meet a previ-
ously determined use plan; and reclamation converts a disturbed site to former uses or to a different, but productive
state (Powter 2002).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 233
preparedness and response (Transport Canada 2013). The selection of best available response technique
should consider local communities and especially aboriginal groups in Canada in case response measures
affect traditional use of lands and waters in either the short- or long-term (e.g., fishing, tourism,
agriculture and aquaculture) (Colten et al. 2012). The extent to which each response technique is used
should be the subject of consultation with local communities. In addition to aboriginal groups,
consultations should include user groups (e.g., fishermen), scientists, community members, business,
media and non-government organizations and should include the potential for local people to be involved
in the cleanup activities (subject to safety and training requirements) (Hoffman and Jennings 2011).

6.5 Research Needs and Recommendations

6.5.1 High Priority Research Needs

6.5.1.1 Cold, harsh environments and controlled field studies

Current technologies, equipment and response personnel are needed to be fully capable of effectively
responding to major inland or offshore spills in Northern Canada and Arctic regions. Special skills and
equipment are needed both now and to be developed for inland or offshore oil spill responses under cold
and harsh conditions. Related to this, most of what we know about oil spill response technologies has
been developed through studies in the laboratory, mesocosm test systems and case studies (with limited
controls and treatment replication). It is the consensus of the scientific community and the Panel that
controlled field trials are needed to advance our spill response strategies, especially for subsurface
blowouts, Arctic oil spills and freshwater shorelines.

6.5.1.2 Sediments

Three in situ remediation technologies developed over the last 20 years for the treatment of contaminated
soils and groundwater have not been tried for remediating bottom sediments contaminated by sunken oil
from spills. These remediation strategies are bioventing, air sparging and dissolved air flotation (DAF),
all of which can be used to oxygenate anaerobic sediments to stimulate aerobic biodegradation of oil that
settled to the bottom. One or all of these techniques may be successful for treating and overcoming the
adverse effects of sunken oil in the anaerobic zone of a river or lake sediment. If such advances can be
made, it would be a huge help in cleaning up benthic environments that have been contaminated by oil
from accidental spills.

6.5.1.3 Dispersants

Although some have concluded that dispersants were successful during the DWH spill, it remains
uncertain to what extent the oil was dispersed physically due to the high velocity of oil emerging from the
well into the water or chemically in response to mixing with applied dispersant. This question needs to be
answered by controlled experiments to reduce the uncertainty of this treatment strategy. Also, questions
about the persistence of dispersed oil in both the deep sea and on its surface need to be addressed with
controlled empirical experiments. This can be partially ascertained through the use of mesocosms and
wave tanks. A further need is to determine if effective and non-toxic microbial biosurfactants, discussed
in Chapter 3, can be developed as alternative sources of dispersants.

6.5.2 Medium Priority Research Needs

6.5.2.1 Bioremediation, including phytoremediation and wetlands

Substantial new research on hydrocarbon biodegradation has been conducted in the Gulf of Mexico
following the DWH spill in 2010. There is an ongoing need to quantify the persistence of oil of all kinds
during and after biodegradation and oil that has been chemically dispersed in the environment following a

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 234
high impact spill to help stimulate innovations in bioremediation approaches. This is especially pertinent
to cold temperature and anaerobic environments.

6.5.2.2 Herding Agents and In Situ Burning (ISB)

Recent advances in the study of herding agents (Buist et al. 2008; Buist et al. 2010 a, b; Buist et al. 2011;
Lee et al. 2013) have been made that suggest they might be useful in corralling an oil slick for ISB. This
may be especially true in Arctic environments. The utility of herding agents for this purpose needs to be
quantified and properly tested to confirm this response approach. Such advances could measurably
improve the use of ISB in Arctic oil spills. ISB is fast gaining a foothold as an effective response
technology for treating a floating oil slick. More quantitative date are needed to determine the emissions
that are released to the environment from ISB that may affect downstream ecosystems adversely, such as
any incompletely combusted PAHs, toxic particulate matter, etc. It is also important to determine the
safety and effectiveness of ISB. More peer-reviewed articles published in the literature are needed in
support of ISB as a safe and effective oil response tool.

6.5.3 Low Priority Research Needs

6.5.3.1 Berm Relocation, surf washing, sediment reworking

Berm relocation, surf washing and sediment reworking have been used successfully to accelerate
shoreline cleanup in the past. Concerns about this process have been primarily linked to the potential fate
and ecological effects of the oil released during sediment relocation. The results of preliminary studies to
date suggest that the oil in near-shore waters is diluted to concentrations within regulatory guidelines for
ocean disposal for dredged sediments (Lee et al. 2003b). More research is needed to advance the status of
this cleanup methodology in terms of fate and effects or released oil.

6.5.3.2 Sorbents, surface washing agents (SWA), solidifiers

Sorbents are widely used as a cleanup tool to combat oil spills in water and on shorelines. They have their
limitations, however. The need exists to determine if any advances have been or can be made to improve
sorption as an effective response tool, especially in the use of biodegradable natural organic sorbents,
such as bagasse, that may possibly not need to be removed for biodegradation to take place. Significant
amounts of debris and solid waste are generated when using sorbents for cleanup, so if natural sorbents
can be developed that would sorb oil and then remain in place for ultimate cleanup by biodegradation, this
would be a significant advancement in oil spill response. There is additional need to determine if any
adverse effects on the ecosystem accrue from the use of SWA chemicals in spill response. In regard to
solidifiers, recently, new research has been advanced in the understanding of the mechanism of solidifiers
in removing oil floating on water. It is unknown if these materials can be made effective on larger oil
spills.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 235
CHAPTER 7: PREVENTION AND RESPONSE DECISION MAKING

Abstract

The Deepwater Horizon (DWH) incident in 2010 triggered scrutiny of the petroleum industry and
regulators across the globe regarding the response preparedness and capability for high-consequence,
low-probability spill events. Similarly, concerns are also growing over low-volume, high-probability spill
events, such as minor spills and leaks during routine operations, due to their contribution towards
negative cumulative effects. Preventing spills is a top priority and a regulatory goal, particularly in those
sensitive areas where spills can cause catastrophic or irreversible consequences, such as in the Arctic.
Meeting this goal requires identification of specific risks at every level of spill preparedness and response,
and targeting those risks with the right approaches to prevent spills or reduce impacts. Once a spill occurs
and triggers response actions, making sound and timely decisions becomes critical. This chapter reviews
the past and current development and practices in prevention and response decision-making with an aim
to deliver an improved integrated strategy for oil spill prevention, preparedness and response.

In terms of spill preparedness, with an adequate level of knowledge, higher-risk activities associated with
the future transport of crude and unconventional oils can be anticipated and mitigated or prevented. Under
this precautionary approach risk is reduced by the development of contingency plans that can provide the
level of capability (trained staff and infrastructure) to manage the potential risks identified. Furthermore,
an integrated spill response strategy should include the application of early warning and real-time
monitoring systems to minimize the release of contaminant oil into the environment. With advances in the
development and application of remote sensing technologies (e.g., surface radar, airborne and satellite
image analysis) and environmental forensics (analytical instruments and standard protocols), cost-
effective tools for detection and source identification in oil spill response operations are now
commercially available.

The current net environmental benefit analysis (NEBA) system for the selection of the most preferred oil
spill countermeasures is dependent on our ability to rapidly monitor, identify and predict the fate,
behaviour and transport of the contaminant hydrocarbons accidently spilled into the environment with
consideration of supporting baseline data for environmental and ecological conditions. Advances in
information technology (e.g., geomatics analysis, artificial intelligence and computer visualization), are
being coupled with oil spill fate/transport and risk assessment models to enhance the decision-making
process for spill response operations. However, the implementation of optimal operational oil spill
response measures in Canada has not been fully achieved due to current limitations in scientific
knowledge (e.g., the behaviour of dilbit under various environmental conditions and the secondary effects
of oil spill response countermeasures such as oil dispersants) and site-specific data. Most existing sources
of essential baseline data are not well coordinated, accessibility of the data varies, and geographic
coverage can be spotty, particularly for remote areas such as the Arctic. Inadequate decision support has
been identified as one of the major challenges that limits the efficiency of current response practices. The
content of this Chapter clearly articulates the need for the development and operational acceptance of oil
spill response decision support systems, which can dynamically and interactively integrate monitoring
and early warning, spill modeling, vulnerability/risk analysis, response process simulation/control, system
optimization and visualization.

7.1 Prevention and Preparedness

All pre-spill strategies emphasize prevention as a key emergency management activity. Prevention
focuses on reducing the likelihood of a spill, thus reducing risk. The basic principle of spill prevention is
that risk reduction through prevention is much more effective than risk reduction through management of
consequences. Prevention is based on understanding the science and technologies associated with oil
operations and potential oil spills, and applying that understanding to the specific environment in which

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 236
an activity takes place, with the goal of managing risk throughout an operation (US GAO 2012). With a
high level of knowledge, many potentially high-risk activities can be anticipated and the risk of spills may
be reduced. For example, when an oil well or tanker is properly designed for the range of anticipated
risks, established procedures are followed, equipment is properly inspected and maintained and incident
response training is also provided. Thus, spill incidences may decline as suggested by ExxonMobil (2013)
and the tanker spill statistics shown in Chapter 1. Furthermore, if accidents occur, the causes and
outcomes can be analyzed providing new information to be used to update and improve prevention
policies and measures.

7.1.1 Prevention

7.1.1.1 Tanker safety and spill prevention

Possessing the world’s longest coastline with a total length of more than 243,000 km, Canada ships 80
million tonnes of oil per year off its east and west coasts (Transport Canada 2015e). As more shipping and
oil exploration are considered for the Arctic, tanker traffic in the Arctic region may increase significantly
from the current low levels of less than 15 tankers per year. Therefore, tanker safety and spill prevention
are among the top priorities.

Safety provisions for oil tankers include navigation safety focused on preventing tanker accidents,
including collisions, groundings and explosions. Navigation safety regulations are set by international and
national agencies and include measures such as:

 Speed regulations;
 Passage plans that respect safe navigation and the environment;
 Mandatory pilotage zones;
 Appropriate navigation technology; and
 Reporting requirements regarding vessel routing measures (e.g. traffic separation, two-way
routes, recommended tracks, precautionary areas and areas to be avoided) (Transport Canada
2014).

In Canada, the Northern Canada Vessel Traffic The Northern Canada Vessel Traffic Services
Services Zone is established where ships carrying Zone is established under the Canada Shipping
over 500 tonnes must check in with officers for Act, 2001. It consists of: a) the shipping safety
inspections of safety and navigation equipment control zones prescribed by the Shipping
and for examination of potential pollutants. Safety Control Zones Order; b) the waters of
Compulsory pilotage and tug escort services for Ungava Bay, Hudson Bay and Kugmallit Bay
difficult waterways are provided by the four that are not in a shipping safety control zone;
Pilotage Authorities in Canada (WSP 2014). c) the waters of James Bay; d) the waters of
Moreover, joint industry-government guidelines the Koksoak River from Ungava Bay to
for the control of oil tankers and bulk chemical Kuujjuaq; e) the waters of Feuilles Bay from
carriers in ice control zones of Eastern Canada Ungava Bay to Tasiujaq; f) the waters of
were established in 2012 (Transport Canada Chesterfield Inlet that are not within a shipping
2015b). These guidelines address the special safety control zone, and the waters of Baker
risks of ice damage in certain waters off the east Lake; and g) the waters of the Moose River
coast during winter and spring and include from James Bay to Moosonee (Government of
provisions, such as: having an ‘ice advisor’ on Canada 2010).
board when transiting an active Ice Control
Zone; travelling at moderate speeds commensurate with visibility and ice conditions; being equipped
with searchlights for night time navigation; and obtaining current ice information and a recommended
route to follow.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 237
Since 2001, Canadian regulations under the Canada
Shipping Act have required double hulls for oil tankers The Canada Shipping Act was established
built after 1993. The Canadian phase-out date for in 2001 respecting shipping and navigation
single-hulled large crude oil tankers was 2010. The and to amend the Shipping Conferences
phase-in period for double-hulled smaller oil tankers Exemption Act, 1987 and other Acts.
was up to the end of 2014. The International Maritime
Organization phase-in period for double-hulled tankers
worldwide will be fully implemented in 2015.

The results of improved tanker safety are reflected in data on the frequency of oil spills due to tanker
accidents. Globally, even with increased tanker traffic, there has been a consistent downward trend in
large oil spills (Figures 7.1). In and around US waters, the average annual oil spillage (spills greater than
0.14 tonnes or one barrel) by oil tankers and tank barges has decreased by 91% and 67%, respectively, in
the decade 1998-2007 compared to previous decades (Figure 7.2). In Canada, the most significant oil
spill in eastern coastal waters occurred in 1970, when the tanker Arrow spilled over 10,000 tonnes of
Bunker C fuel oil off Chedabucto Bay, NS (also refer to Chapters 1 and 8). Since then, no large spills
greater than 700 tonnes have occurred in Canadian coastal waters. The most significant spill in the last 20
years off the British Columbia coast was in 2006 when the ferry the Queen of the North sank with 240
tonnes of oil on board, including 225,000 litres of diesel fuel, 15,000 litres of light oil, 3,200 litres of
hydraulic fluid and 3,200 litres of stern tube oil (BC Ministry of Environment 2015). The annual
Canadian ship-source spill frequency for crude oil from 2003-2012, based upon the data from the
Canadian Coast Guard Marine Pollution Incident Reporting System, was close to zero for all spill sizes
(WSP 2014a,b). The recent WSP reports (2014a,b) quantified the risk of marine spills in Canadian waters
and gave an optimistic estimate of future ship-source spills as shown in Table 7.1.

Figure 7.1 Seaborne oil trade and number of tanker spills seven tonnes and over, 1970 to 2013 (Crude and Oil
Product; vessels of 60,000 DWT and above; barges excluded.) Image from ITOPF (2015)

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 238
Figure 7.2 Average annual large oil spills by oil tankers and barges in US. Data from API (2009)

Table 7.1 Future Ship-source spills in Canadaa


Volume (m3) 10-100 100-,1000 1,000-10,000 >10,000
Annual 0.022 0.014 0.019 0.004
Frequency
Return Period 46.4 69.2 51.6 242.3
(yrs)
a
Data from WSP (2014a,b)

7.1.1.2 Pipeline safety and spill prevention

The major existing pipelines carrying crude oil in Canada are (Figure 7.3): the Trans Mountain from
the Edmonton area to Vancouver and Anacortes, WA; the Keystone pipeline from the Edmonton area
through southern Saskatchewan and Manitoba; the Enbridge pipeline from the Edmonton area through
southern Saskatchewan and Manitoba before entering North Dakota and then to a network delivering
oil to destinations, such as Chicago, Sarnia (via Michigan) and the Gulf of Mexico; the Express
pipeline from Hardisty, AB, through southeastern Alberta to Montana and then east/southeast where it
joins the network delivering oil to the Gulf of Mexico; and the Enbridge Line 9 running between
Sarnia and Montreal. Other feeder pipelines include the Western (Pembina) between Kamloops and
Edmonton, the Rainbow pipeline from Norman Wells and pipelines from the Fort McMurray, Cold
Lake and Lloydminster oil sands and heavy oil producing areas. With the exception of Enbridge Line
9, all crude oil pipelines in Canada are in the four western provinces.

Due to the strong growth of oil and natural gas production in Canada and the US, pipeline capacity is
expected to expand to provide access to new markets. A number of pipeline projects are being proposed
to connect the growing supply with replacement markets in eastern Canada and beyond to growing global
markets (CAPP 2015). Although the statistics indicate no significant increase in numbers and volumes of
pipeline incidents in Canada since 2008 (Figure 7.4), concerns are growing due to major ongoing or
proposed projects and the recent frequent spills in North America, such as Refugio Oil Spill in California
on May 19, 2015 (Panzar et al. 2015), Yellowstone River oil spill in Montana on January 21, 2015 (Yan
2015), Mid-Valley Pipeline Spill in Louisiana on October 13, 2014 (Maykuth 2014), and Hiland Crude

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 239
Pipeline Spills in North Dakota on March 21, 2014 (MacPherson 2014). The National Energy Board
(NEB), which regulates pipelines in Canada, recently introduced new pipeline performance measures
(NEB 2014b). These measures include training and competency assurance, ‘integrity management’
(pipeline conditions, equipment inspection, assessment of pipeline hazards, shutdowns for hazard control
etc.), integrity inspections, environmental inspections, contractor awareness and damage prevention (e.g.,
unauthorized activities by other industries or municipalities). The conditions attached to the recent NEB
approval of the reversal of Northgate Line 9B reflect these measures (e.g., analysis of threats to integrity
of the pipeline, a remaining life analysis, repairs to all features in pipeline sections identified by
engineering assessments and a hydrostatic testing program) (NEB 2014a).

Figure 7.3 Canadian Energy Pipeline Association (CEPA) liquids pipelines map. Image used with permission
from CEPA.

The NEB requires its regulated companies to promote a positive safety culture and expects safety culture
programs to apply not only to worker health and safety but to process safety (NEB 2015a). Process safety
focuses on preventing catastrophic incidents associated with the use of chemicals and petroleum products.
Safety culture is a vital component of spill prevention and includes the empowerment of employees to
take immediate action (e.g., closing of valves) without waiting for authorization up a chain-of-command.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 240
In 2012, Enbridge filed an application with the NEB to reverse the flow on the segment of Line 9
between North Westover, ON, and Montreal (Line 9B) and that the entire Line 9 capacity from Sarnia,
ON, to Montreal be increased with a revision to regulations to allow transportation of heavy crude
(NEB 2015b). The NEB had previously approved flow reversal between Sarnia and North Westover in
July of 2012. Line 9 originally flowed eastward after its construction in 1975. Flow was reversed in
1998 as oil from areas such as West Africa and the Middle East became more affordable. However,
western Canadian crude is now priced lower than foreign oil, leading Enbridge to propose to reverse
the flow of Line 9 once again to provide oil to two refineries in Quebec, which represent 20% of
Canadian capacity. The NEB approved the reversal on March 6, 2014, subject to 30 conditions which
are currently being addressed by Enbridge (NEB 2014a). However, concerns regarding the potential
risks and regulatory requirements (e.g., environmental review) associated with using aged facilities
have been raised.

Figure 7.4 Total volumes of liquids released vs. number of liquid releases in Canada between 2008 to 2015.
Image adapted from NEB (2015c)

Under the NEB Onshore Pipeline Regulations (OPR), companies must immediately notify the NEB of
any incident that relates to the construction, operation or abandonment of a pipeline. Generally, spills of
crude oil from pipelines in Canada have been infrequent, small (< 1 to 30 m3) and rarely enter water
bodies, although there are variations in volumes and frequencies among provinces and companies (AER
2015). However, some larger incidents have revealed the need for an increased focus on prevention.
Regulatory examination of larger releases in the range of 1,000-5,000 m3 in Alberta since 2011 indicated
issues, such as: failings in the identification of risks related to corrosion and corrosion repair; inadequate
leak detection alarm and response processes; and lack of training and supervisory oversight (ERCB
2013). For example, a large spill caused by a ruptured year-old pipeline was observed on July 15, 2015, at
Nexen Energy's Long Lake oil sands site in Alberta. It released about five million litres of bitumen, sand
and produced water and affected an area of about 16,000 m2 along the pipeline's route (The Canadian
Press 2015a). During the writing of this report, Alberta's energy regulator issued an order on August 21 to
immediately suspend the company’s operations of 95 pipelines in northeastern Alberta due to non-
compliance with the Pipeline Act surrounding pipeline maintenance and monitoring in its Long Lake oil
sands project (The Canadian Press 2015b).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 241
These recent events further raised the attention and importance of how to prevent and monitor spill
incidents in transporting oil by pipelines. Research investment is inadequate to support the development
of optimal response strategies and enhancement of response capacities in preventing and remedying
spills from aging pipelines for crude oil transport.

Note that Canada-based Enbridge Partners will stop using its No. 5 pipeline under the Straits of Mackinac
to transport heavy crude oil under an agreement reached September 3, 2015, with the state of Michigan
(Dalbey 2015). This binding legal agreement places a permanent ban on use of the aging pipeline to
transport heavy crude and formally implements the first recommendation of the Michigan Petroleum
Pipeline Task Force Report released in July to “prevent the transportation of heavy crude oil through the
Straits Pipelines” (Department of Attorney General 2015). The Report gave the recommendation based on
the Coast Guard’s public acknowledgement that they lack capacity to effectively respond to spills of
heavy crude oil in the Great Lakes (Matheny 2014). In addition, Governor Rick Snyder of the State of
Michigan issued an executive order creating the Michigan Pipeline Safety Advisory Board on the same
day of signing the agreement.

Information regarding the nature and extent of effects of both smaller and larger pipeline incidents, and
the corresponding response actions, is presented in Chapter 6 and later in this chapter.

7.1.1.3 Oil transport by rail: safety and spill prevention

On August 3, 2005, 43 cars derailed just west of Edmonton spilling 700,000 litres of Bunker C fuel oil
and 70,000 litres of a wood preservative into the waters of Wabamun Lake, a recreational mecca for many
Edmontonians and home to the Paul Band First Nation (Benoit 2007). As admitted by Alberta Premier
Ralph Klein at that time, his government was unprepared to deal with the oil spill into Lake Wabamun
caused by a train derailment (CBC 2005). A Transportation Safety Board report into the derailment found
that CN did have an emergency plan to respond to the spill, but pointed out that it would have been more
efficient had there been closer coordination with other agencies (Benoit 2007). The Government of
Alberta formed a Commission to investigate the incident and the most significant among the
Commission’s findings was the absence of a regulatory framework to govern emergency response to
major spills of environmentally hazardous goods that fall outside the purview of the Transportation of
Dangerous Goods Act (Benoit 2007).

Authorities have learned from recent spill


events and are gradually improving response
After the Lac-Mégantic spill, further measures
management. For example, in response to the
required by Transport Canada have included
derailment and explosion at Lac-Mégantic, QC,
(Transport Canada 2015a):
in July 2013, Transport Canada, which
• Removal of the least crash-resistant DOT-111
regulates the transport of crude oil by rail, took
tank cars from dangerous goods service;
immediate action by establishing a two-person
• New safety standards for DOT-111 tank cars,
minimum for locomotive crews on trains
and requiring those that do not meet the new
carrying dangerous goods, and by imposing
standards to be phased out by May 1, 2017;
stricter rules for securing unattended trains
• Slower trains transporting dangerous goods and
(Transport Canada 2015a). Transport Canada
introducing other key operating procedures;
has begun a process of enhanced inspections,
• Emergency response plans for even a single tank
documentation and follow-up for rail safety,
car carrying crude oil and refined products; and
including more frequent inspections at sites
• Setting up a task force that meets regularly and
where petroleum products are transferred from
brings municipalities, first responders, railways
one mode of transport to another (e.g., from
and shippers together to strengthen emergency
truck to rail). The department has also
response capacity across the county.
proposed or introduced fines, mandatory
certificates and new safety and transportation

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 242
of dangerous good regulations and reporting requirements on railways and shippers. Through Protective
Direction 32, railway operators must provide municipalities and first responders with information on the
dangerous goods being transported through their communities. The department has also updated
regulations to bring cross-border consistency to the way dangerous goods are identified, providing
emergency personnel with a clear understanding of the risks posed by goods being transported.

Railway companies are now expected to have safety management systems, which include safety goals,
performance targets, risk assessments, responsibilities and authorities, rules and procedures, and
monitoring and evaluation processes. The systems can help companies better comply with federal
legislation. Transport Canada monitors compliance through formal safety management system audits and
detailed inspections and takes enforcement action in case non-compliance is identified.

According to the Transportation Safety Board of Canada, there were 174 railroad accidents involving
dangerous goods in 2014, up from 145 in 2013 and a previous five year average of 131 (TSB 2015). Only
one accident in 2014 involved the release of petroleum crude oil. The number and type of rail incidents in
Figure 7.5 illustrates the range of causes, any of which could apply to crude.

Figure 7.5 Number of rail incidents by type in 2014. Image from TSB (2015).

New prevention measures target these causes, as evidenced by Transport Canada’s intention to expand
mandatory requirements under the Railway Safety Management System Regulations. The proposed
changes (Transport Canada 2015a) include processes to:

 Encourage employees to report accidents to senior management;


 Analyze data and trends to identify safety concerns;
 Manage organization knowledge so employees can perform their duties more safely;
 Improve work scheduling to prevent employee fatigue; and
 Create annual safety targets and develop tools to achieve them.

Recommendation: Continued efforts are needed to understand the associated impacts of, and
improve response management for, spills caused by railroad accidents.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 243
7.1.1.4 Oil transport by truck: safety and spill prevention

The Motor Carrier Division of Transport Canada is primarily responsible for facilitating the reduction of
accidents involving large commercial trucks in Canada. Provincial and territorial regulations govern the
operation of the commercial truck industry except for the rules and regulations set by the Division in
support of the safe operation of extra-provincial motor carriers that carry goods across a provincial or
international boundary (Transport Canada 2015c). However, provinces and territories are allowed to
regulate extra-provincial truck carriers under the Motor Vehicle Transport Act on behalf of the federal
government. In addition, the federal Transport of Dangerous Goods Act applies to crude transport by
truck. However, the provinces are responsible for inspecting vehicles under the Act. Fifteen standards for
motor carrier safety are included in the National Safety Code and are supported by provincial regulations.
A Safety Management Systems approach is emphasized by the Motor Carrier Division, which is the same
approach used to oversee other modes of transport (marine, rail) by Transport Canada. This approach
expects the industry to develop a safety culture aimed at identifying and managing risks.

As outlined in Chapter 1, while the total volume of crude oil transported by truck in Canada is very small
relative to transport by pipeline, the number of incidents per volume of crude transported is an order of
magnitude greater. Experience with truck-related oil spills indicates that the largest threats are:
contamination of nearby water bodies; effects on drinking water supply or water use for industrial or
agricultural purposes; fire and explosion and economic impacts caused by property damage; disruption to
traffic due to road closures; and disturbance to habitats (Christopherson and Dave 2014). Therefore, how
to decrease the number and severity of crude oil spills involving truck transport is an important issue.
Suggested prevention measures could include: improved monitoring and maintenance of key
infrastructure at delivery points and fuel loading terminals; identification of high-risk sections of common
routes (dense population or highly sensitive environmental areas); and improved truck design and data
collection/accessibility for risk management.

7.1.2 Oil spill monitoring

7.1.2.1 Onsite monitoring

Canada conducts regular aerial surveillance of marine shipping under the National Aerial Surveillance
Program. Transport Canada reported that in 2011-2012, surveillance crews observed more than 12,000
vessels and detected 135 pollution occurrences nationally, with an estimated total volume of 1,014 litres
of oil (Transport Canada 2015e). There is also an obligation for owners of vessels and operators of oil
handling facilities to report marine spills to the Canadian Coast Guard.

During and after spill cleanup operations, it is essential to continuously monitor the sea or site state and
weather conditions, as well as the movement and behaviour of the oil. In marine environments, oil spills
can be tracked using specially-designed buoys to follow oil movement. Fingas (2012) reviewed oil spill
tracking buoys and devices from 1970 until the present and pointed out that the buoys were being
beneficially used during exercises and on occasion to simulate oil spills to improve preparedness. Visual
monitoring can be undertaken from vehicles, vessels and/or, ideally, from spotter planes or even
unmanned aerial vehicles, which can also ensure that the heaviest concentrations of oil are being targeted
(Ferraro et al. 2009; Pisano 2011). Lieske et al. (2011) performed a comparative and retrospective
analysis of the spatial pattern of oil pollution for the Canadian east coast based on the data obtained from
visually unaided aerial surveillance, side-looking airborne radar-assisted aerial surveillance, as well as
remote sensing images. The results provided insights into Environment Canada’s surveillance programs
and offered a useful technique to combine information from different surveillance regimes. Observers
should look out for a visible reduction in oil coverage, as well as a change in the appearance of the oil,
often seen as a coffee-coloured plume within the water column in good viewing conditions (Chapman et
al. 2007; Tan 2011). Submerged flow-through systems using ultraviolet fluorescence spectrometry

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 244
(UVFS) have also been used to monitor oil concentrations and plumes in water to indicate the
effectiveness of dispersion (Lambert 2003; Chatterjee 2015). Although UVFS cannot measure the exact
amount of oil, it is able to qualitatively determine whether or not dispersed oil in the water column has
increased significantly and therefore whether or not the application of chemical dispersants has been
successful (Chapman et al. 2007). The laser fluoresensor is one of the most useful and reliable
instruments to detect oil on various backgrounds, including water, soil, weeds, ice and snow. It is the only
reliable sensor to date to detect oil in the presence of ice or snow (Brown and Fingas 2003). However,
detection and monitoring of oil spills are always challenging for a number of reasons, such as the
substantial land and sea area to cover, low probability and high variability of discharge detection, and
limited efficiency, high uncertainty and logistic difficulties due to harsh conditions (e.g., dense vegetation,
snow and ice coverage, low visibility, rough waves and darkness).

In recent years, autonomous underwater vehicles (AUVs) and remotely operated vehicle (ROVs) have
been proven especially productive and efficient in spill monitoring. They have emerged as the platform of
choice for many surveying tasks, such as bathymetric, near-bottom geophysical and optical surveys.
Hundreds of shallow water AUVs (e.g., REMUS) and ROVs (Allen et al. 1997) are operated around the
world. Fewer deepwater AUVs and ROVs exist. However, these larger, more expensive systems are
routinely used for industrial, military and scientific applications. Notable examples of these systems
include the Hugin (Marthiniussen et al. 2004) and Autosub (Griffiths et al. 1999). After the Deepwater
Horizon (DWH) oil spill, a new 4,500 m depth-rated AUV sentry was developed and applied in two
separate cruises (WHOI 2010).

7.1.2.2 Remote sensing

Remote sensing technologies have been tested and applied in oil spill responses for decades with mixed
results (Goodman 1994; Fingas 2011; Salt et al. 2012). Their incorporation into oil spill response requires
confidence in the robustness and reliability of the information, based on capturing oil spatial patterns
consistent with oil spill response processes. Beyond technical challenges, considerations of cost-
effectiveness and difficulties in conducting field trials have limited actual field applications to data
processing (Leifer et al. 2011).

Recent technological advances in remote sensing, multi-spectral airborne imaging and echo-sonic tools,
along with increased efficiency and lower cost, have created a new set of potentially practical tools for
application in oil spill monitoring, preparedness and responses. For example, Ying et al. (2011)
discussed recent research and progress in monitoring sea surfaces covered by oil using microwave remote
sensing technology and also proposed a simple and effective filtering technique based on synthetic
aperture radar (SAR) for sea oil slick observation. Leifer et al. (2011) reported a new spectral library
approach that uses near infrared imaging spectroscopy data from the airborne visual/infrared imaging
spectrometer (AVIRIS) instrument on the NASA ER-2 stratospheric airplane. Extrapolation to the total
slick used MODIS (moderate resolution imaging spectroradiometer satellite visual-spectrum) broadband
data, which observes sunlight reflection from surface slicks. Ermakov (2013) summarized the results of
field studies of the effect of damping of short wind waves due to surface films conducted using radar and
optical methods. The results showed that film slicks can be characterized by their specific “spectral
contrast sign” and can be discriminated from some look-alikes like wind depression areas. Some radar-
based oil spill detection systems such as the Oil Spill Detection (OSD) system of Rutter Incorporated
have been proven to be effective through extensive field tests and adopted on various platforms, such as
floating production storage and offloading vessels, offshore workboats, patrol vessels and specialty
cleanup vessels. Etkin et al. (2013) foretold that with more field testing and reductions in costs, some of
these technologies will eventually become more ‘tools in the toolbox’ for spill response.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 245
Recommendation: Develop improved in situ and remote oil leak detection systems to detect spills
soon after an incident and post spill monitoring, including autonomous, remotely operated
detection devices (e.g., AUVs and ROVs).

7.1.2.3 Shoreline monitoring

When shoreline impact occurs or is likely to occur after a spill, shoreline monitoring and assessment are
critical components of the response program, and they provide essential information for setting goals and
priorities and making decisions for an effective shoreline response (IPIECA 2015). The Shoreline
Cleanup Assessment Team (SCAT) approach, which originated from the Exxon Valdez oil spill (EVOS)
response in 1989, has been widely used in many spills and continuously developed to meet specific spill
conditions (Owens and Sergy 2004; Owens and Reimer 2013). A SCAT survey uses a set of specific and
standard terminologies to define and assess shoreline oiling conditions, although the process itself is
diversified and the assessment activities can be designed to match the individual spill conditions (Owens
and Sergy 2000; IPEICA 2014). The surveys may be applied to spills in both marine coastal and
freshwater environments and have become integral components of spill response in Canada, the US and
several other countries. An effective SCAT or shoreline assessment program should be integrated into a
shoreline response program and provide critical information to support the contingency planning,
response decision-making and implementation/closure processes.

The scope and scale of a SCAT survey is governed by a range of parameters that include: the coastal
character and configuration; the type and amount of spilled oil; the size of the affected area; and the needs
of the response organization. The basic principles that govern a SCAT survey are (Owens and Sergy
2000):

 A systematic assessment of all shorelines in the affected area;


 A division of the coast into geographic units or ‘segments’;
 A set of standard terms and definitions for documentation; and
 A team of interagency personnel to represent the various interests of the responsible party, land
ownership, land use, land management or governmental responsibility.

The first complete manual for SCAT surveys was developed by Environment Canada (1992), followed by
production of a more simplified Field Guide (Owens and Sergy 1994). Other agencies, including the US
National Oceanographic and Atmospheric Administration and U.K. Maritime and Coastguard Agency,
have adopted the approach and produced similar field forms and manuals (US NOAA 1994, 2013; MCA
2007). The Second Edition of the SCAT manual was developed by Environment Canada in 2000 with
improved guidelines for the description and classification of oiled shorelines that were directly
compatible with the format and content of forms used by the NOAA. In 2004, the Arctic Edition of SCAT
Manual was published by Environment Canada. It is based on the Second Edition but provides new
material on the unique shoreline types, characters of the various forms of snow and shore-zone or near-
shore ice, cold climatic conditions, and behaviour of oil in arctic regions, as well as on the SCAT
activities in these environments (Owens and Sergy 2004).

Recommendation: Future improvements may focus on how to more effectively and dynamically
integrate SCAT into the entire response system covering offshore to shoreline operations, as well as
the survey data, into the response modeling and decision-making processes.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 246
7.1.3 Oil and sample analysis

7.1.3.1 Physical analysis

Efficient determination of oil properties can enable effective oil spill response and countermeasures.
Some frequently used analytical methods for major oil properties (viscosity, density, specific gravity,
flash point, pour point and interfacial tension) are summarized in Appendix B.

7.1.3.2 Chemical analysis

Crude oil generally contains hydrocarbons (e.g., paraffins, aromatics, naphthenes and asphaltenes) and
heterocompounds (e.g., sulfur, oxygen, nitrogen and metal ions) (Yang 2013a). Being able to
unambiguously characterize, identify, categorize and quantify all sources of hydrocarbons entering the
environment is important and necessary for oil spill response. Analytical techniques have been applied
extensively in the laboratory to measure and quantify oil constituents belonging to four primary
categories: saturated hydrocarbons, aromatics, resins and asphaltenes (Wang et al. 2003) (In-depth
discussion can be found in Chapter 2). A summary of the frequently used instrumental techniques is given
in Appendix A.

Advances in environmental forensics analysis have played an ever increasing role in oil spill analysis
(Wang and Stout 2006). Biological markers or biomarkers (see definition in Chapters 2 and 4) are among
the most important hydrocarbon compounds in petroleum for chemical fingerprinting (Wang et al. 2002,
2003; Yim 2011; Radović et al. 2014). Biomarkers have also been proven useful in identification of
petroleum-derived contaminants in marine and aquatic environments. Consequently, biomarker
fingerprinting techniques have been gradually recognized and increasingly used in characterization,
correlation, differentiation and source identification in oil spill investigations (Wang et al. 2012).
Biomarkers can be detected in low quantities (µg/L and sub-µg/L levels) in the presence of a wide variety
of other types of petroleum hydrocarbons by the use of the gas chromatography/mass spectrometry
(GC/MS). For example, Wang et al. (2011) presented integrated forensic oil fingerprinting and data
interpretation techniques to characterize the chemical composition and determine the source of the 2009
Sarnia, ON, oil spill incident. Other examples in the literature confirm the advantages of forensic
fingerprinting using these biomarkers (Yim et al. 2012; Yang et al. 2013a, b; Wang et al. 2013, 2014).
Recently, efforts were also reported on identifying suitable biomarkers and applying them for dispersed
oil fingerprinting (Song et al. 2014).

Recommendation: Develop and refine environmental forensics technologies for analysis of oil
components of interest. Establish standard fingerprinting protocols including new biomarkers to
track dispersed oil, diluted bitumens and other petroleum hydrocarbons of interest. Develop
protocols to identify and monitor oil and its compositional changes over time in contaminated
sediments (on river or sea beds), particularly in the presence of ice and harsh marine conditions.

7.1.3.3 Biological monitoring

Biological monitoring (or biomonitoring) is the use of a biological community to provide information on
the quality or ‘health’ of an ecosystem (NYSDEC 2015). Basic parameters for biological monitoring
include (Reilly and York 2001):

 Mortality of large animals and plants, which are easier to monitor (identification of mortality
among smaller organisms and plants need special input);
 Sublethal effects, such as bioaccumulation, gene activation, CYP1A induction, behavioural
changes or histopathological effects (e.g., presence of disease or injury Chapter 4), which need
specialist input;

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 247
 Changes in community structure, such as changes in species diversity; and
 Tainting after fish ingest the oil and incorporated into its fatty tissues.

Existing response technologies cause various levels of impacts on the environments. For example,
recently interest has increased in the impact of dispersants and dispersed oil particularly focusing on the
DWH spill. Judson et al. (2010) discussed the potential for toxicity of the dispersants used in that spill
and potential alternatives at the spill site, especially given the limited toxicity testing information that is
available. Using a series of in vitro high-throughput assays, eight commercial dispersants were analyzed.
White et al. (2012) examined deepwater coral communities three to four months after the well was capped
to assess the potential impact on offshore ecosystems. The results showed the presence of recently
damaged and deceased corals beneath the path of a previously documented plume emanating from the
Macondo well providing evidence of impact on the deepwater ecosystems. Paul et al. (2013) presented
data on general toxicity and mutagenicity of upper water column waters and, to a lesser degree, sediment
pore water of the northeastern Gulf of Mexico and west Florida shelf at the time of the spill in 2010 and
thereafter. See more discussions in Chapter 4.

Biological monitoring is a proven approach to help collect baseline data before a spill event and evaluate
the effectiveness and impact of response and cleanup actions after the event. It can also provide
assistance to development of more eco-friendly and cost-effective response technologies.

Recommendation: More research on monitoring parameters/protocols and technologies, as well as


result interpretation and integration with decision-making processes, are desired.

7.1.4 Baseline data

Many environmental and ecological databases and archives hosted by governments have been used in oil
spill responses in Canada. Continued efforts have been made to establish and maintain the sources of
baseline information at the federal, provincial and local levels. Some examples are as follows:

 National and interagency programs, such as:


 The GIS Services Group of Canadian Coast Guard maintains databases containing geographic
locations of features (e.g., light stations, aids to navigation, pollution response sites) and
events (e.g., search and rescue incidents). The databases are incorporated into a GIS
(Dollhopf et al. 2014); and
 The Canadian Aquatic Biomonitoring Network (CABIN) maintained by Environment Canada
to support the standardized collection of biological monitoring information and assess the
health of freshwater ecosystems in Canada (Environment Canada 2015a, b).
 Provincial government monitoring programs, such as:
 GeoBC , which provides a one-stop shop for spatial information for the province, including
base maps, specific datasets such as conservation status information for 7,400 plants and
animals and fisheries sensitive watersheds, and digital imagery;
 The former BC Burrard Inlet Environmental Action Program was an interagency partnership
that compiled baseline data on habitat types, intertidal vegetation, number of bird nests,
outfalls, docks, impervious surfaces and numerous other coastal conditions in the inlet
(Chang et al. 2014). These could be instrumental in ranking environmentally sensitive areas
to inform a hierarchy of response efforts in the event of an oil spill;
 The Alberta Environmental Monitoring, Evaluation and Reporting Agency, a science-based,
comprehensive, arm’s-length organization responsible for coordinating province-wide
environmental monitoring and evaluation with emphasis on the oil sands region; and
 The Canada-Newfoundland and Labrador Offshore Petroleum Board (CNLOPB), which has
continuously produced and compiled the environmental assessment reports conducted by

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 248
authorized consulting company and made them open for access by the public (Lin et al. 2009;
Payne et al. 2012).
 Local conservation authorities, such as the Grand River Conservation Authority in Ontario, which
manages water and other natural resources on behalf of 39 municipalities and close to one million
residents and operates a network of automatic gauges throughout the watershed to measure river
and stream flows, weather conditions and water quality (GRCA 2015).
 Provincial compliance monitoring data for specific permitted facilities which can provide some
relevant data near the spill site, such as basic water quality, biological monitoring data and/or
facility information.
 Community-based monitoring programs, such as:
 Streamkeepers training program, a comprehensive education and awareness program
supported by Fisheries and Oceans Canada to protect and restore local aquatic habitats
(Pacific Streamkeepers Federation 2003); and
 The Atlantic Coastal Action Program, a unique community-based program initiated by
Environment Canada to help Atlantic Canadians restore and sustain local watersheds and
adjacent coastal areas (ACAP 2008).

However, challenges and gaps still exist in the baseline data, particularly in acquisition, integrity,
sharing and uncertainties. For example, the British Columbia baseline data could not address species’
population levels within the Burrard Inlet, which would be
helpful for determining species most at risk and essential for Comprehensive baseline
quantifying impacts of a spill event. Stable legislative and assessments of natural resources
financial mechanisms are lacking in driving long-term data and socioeconomic entities are
acquisition and management at both federal and provincial important for informing
levels. Access to baseline data and data sharing are still limited environmental and economic
among government, industry and academia, which constrains oil impact studies, supporting
spill response research and practices (Lieske et al. 2011; Marty contingency planning and response
and Potter 2014). Uncertainties unavoidably exist and come from decision-making, and facilitating
a variety of sources during data acquisition, processing and accurate monitoring after a spill
manipulation, leading to the need for quantifying the impacts has occurred.
and developing mitigation methods.

In addition, most of those existing sources of baseline data are not well coordinated, accessibility of the
data varies and geographic coverage can be spotty, particularly for remote areas. Besides further
extending the baseline information, there is a strong need for a thorough examination of available data
applicable to establishing a defensible baseline dataset for high-risk aquatic systems adjacent to crude oil
shipment routes, whether by marine tanker, pipeline, rail or road. One possible place to start might be a
national directory of monitoring information with links to databases, GIS maps and reports.

7.1.5 Sensitivity and vulnerability

7.1.5.1 Sensitivity

Sensitivity is regarded as one of the central concepts in ecosystem protection (Holt et al. 1995; Tyler-
Walters and Jackson 1999). It is axiomatic that all ecosystem features have either evolved (in the case of
biotic features) or been formed (in the case of abiotic features) within a certain range of environmental
conditions.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 249
Sensitivity can be measured using one or more indicators (of species, communities and habitats) that
respond to oil spills. These responses are potentially nonlinear and are likely to include interactions
among stressors (Adler and Inbar 2007). In this context,
Zacharias and Gregr (2005) defined sensitivity does not inherently assume the characteristics of
sensitivity as the degree to which fragility or intolerance with which it is often associated.
features respond to stresses, such as There is no implied judgement that an increased association
spills, which are deviations in between the indicator and the stressor reduces a feature's
environmental conditions beyond the probability of persistence. Nevertheless, as exposure to a
expected range. chronic perturbation or stress increases, the persistence of
that feature is diminished.

7.1.5.2 Vulnerability

Vulnerability is the likelihood of exposure to a relevant external stress factor, combined in some way with
the exposure (duration, magnitude, rate of change) to that stress. Zacharias and Gregr (2005) provided a
quantitative methodology for identifying vulnerable areas based on valued ecological features, defined as
biological or physical features, processes or structures deemed by humans to have environmental, social,
cultural or economic significance.

Responding to an oil spill is extremely challenging in any


environment, especially those regions where extreme weather Vulnerability is defined as the
prevails (Pilipenko et al. 2013). For example, the offshore probability that a feature will be
operating season in the Arctic, and therefore the period when exposed to a stress (e.g., oil spills) to
it would be possible to clean up an oil spill, is restricted to which it is sensitive (Adger 2006;
four to five months by darkness, heavy ice and extreme cold Metzger et al. 2006).
(Reich et al. 2014). These severe conditions would make it
impossible to attempt an oil spill cleanup for half of the time during the operating season and 100% of the
time during the winter. In this intervening time, the Arctic environment would be extremely vulnerable to
any possible oil spill event. Due to the low temperature and harsh climatic conditions, the natural
attenuation rate is extremely low in this region (Margesin and Schinner 2001). Given the Arctic region
has limited biodiversity, damage to some vital species can lead to widespread changes that cascade
through all components of the social-ecological system (Chapin et al. 2004).

7.1.5.3 Vulnerability analysis

When responders are making decisions regarding the hierarchy of vulnerable areas to be safeguarded after
an oil spill, they need to compare the biological impact and recovery potential, the relative value of the
population among other impacted populations and the technical possibilities to help the population.
Incorporation of vulnerability analysis into risk assessment has been increasingly adopted in
governmental initiatives, programs and practices. For example, in 2013, Transport Canada conducted a
nationwide risk assessment to determine the risks associated with ship-source spills in Canadian waters
south of 60° north latitude and in the Arctic (Transport Canada 2015d). Vulnerabilities have been
considered as inputs to risk assessment. On the other hand, growing efforts have been made to develop
more effective methods to classify and evaluate vulnerability to oil spills. For example, Reich et al. (2014)
developed a model of environmental vulnerability to determine the relative vulnerability of broad
geographic regions to spilled oil. The model was based on the underlying vulnerability of habitats and
representative species present in each region and season assessed. It can be applied to any geographic area
of interest and provide a simple, yet comprehensive and expandable methodology for determining
baseline environmental vulnerability of large geographic regions. Ihaksi et al. (2011) presented an index-
based method that can be used to make decisions concerning which populations of classified natural
organisms should be primarily safeguarded from a floating oil slick with oil booms. Olita et al. (2012)
proposed a new model for evaluating the hazard of oil slicks contacting shorelines based on the

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 250
geomorphology of the intertidal area. The hazard index layer can be used as a solid basis to assess the risk
and support oil spill response decision-making. Recently, Li et al. (2014a) developed a Monte Carlo
simulation-based two-stage adaptive resonance theory mapping model to classify a given site into
distinguished zones representing different levels of an offshore oil spill vulnerability index (OSVI).
Integrating Monte Carlo simulation with the adaptive resonance theory and mapping approach addresses
the uncertainties that widely exist with the environmental conditions. The classification results can also
provide the least desired number of zones that can sufficiently represent the levels of offshore OSVI in an
area under uncertainty and complexity, saving time and budget in spill monitoring and response (Li et al.
2014c). However, limited studies focused on the Arctic and sub-Arctic environments are reported in the
literature. Vulnerability analysis based on field measurement and integrated with spill modeling and
response decision-making are again lacking.

7.1.6 Preparedness and contingency planning

A management strategy or contingency plan is a set of instructions that outlines the steps that should be
taken before, during and after an emergency. It helps minimize potential damage to human health and the
environment by ensuring a timely and coordinated response (Chen et al. 2012).

7.1.6.1 Preparedness and response regime

In Canada, the designation of the lead agency for a spill event may be based on legislation, an interagency
agreement, a Cabinet decision and/or custom or precedent. There can be more than one lead agency
represented under a unified command, as well as the party that is taking responsibility for impact
mitigation (e.g., cleanup and response management) and generally referred to as either the spiller or
polluter. Table 7.3 shows a list of the lead agencies in Canada for different types of oil spills based on the
National Environmental Emergencies Contingency Plan (Environment Canada 1999).
Table 7.3 Lead agencies for different types of oil spills
Federal Legislation on Oil Spills Lead Agency
Deposition of oil into migratory bird habitats
Environment Canada
(Migratory Birds Convention Act)
Oil pollution from ships Department of Fisheries and Oceans;
(Canada Shipping Act) Transport Canada
Cleanup, sampling and analyzing oiled wildlife Environmental Conservation Service
Oil loading and unloading from ships
Canadian Coastal Guard
(Canada Shipping Act)
From loading facilities connected to oil production platforms
National Energy Board
in Canadian Atlantic

For example, for marine oil spills, Transport Canada is the administrative agency that ensures an
appropriate level of preparedness is available for any oil spill incidents occurring within prescribed time
standards and operating environments. The Marine Oil Spill Preparedness and Response Regime,
established in 1995, is built on a partnership between government and industry (Transport Canada 2013).
The regime is designed to ensure that marine shippers pay for spill preparation, response and cleanup. The
regime also includes requirements for rigorous contingency planning. The regime is equipped to handle
up to 10,000 tonnes of oil spilled in Canadian marine waters. The Canadian Coast Guard monitors and,
where necessary, augments or assumes management of the response when it is in the interest of the
public.

Privately-funded certified Response Organizations (RO) have the responsibility to respond to oils spills
from vessels with which they have arrangements and when contracted to respond. Ship owners are

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 251
required to have such an arrangement with one or more ROs, depending on the intended destination(s) of
the ship and the area covered by each RO. However, there is no legal obligation to enact the arrangement
and engage the services of the RO. Alternative arrangements could be made using other resources, if
deemed appropriate. Below 60 °N, Canadian waters have been divided into two principal areas: west and
east coasts, with an RO established for each. In addition, two further ROs have been established to cover
specific parts of the eastern coast region. Each RO has a Response Plan establishing the resources and
strategies needed to respond to a range of spill scenarios within its jurisdiction (ITOPF 2013).

Although it is prudent to have some level of preparedness for a worst case spill, the primary focus of the
ROs should be preparing for the types of spills that are likely to occur within their area of response. The
challenge is how to determine what level of preparedness is appropriate for each area of response.
Transport Canada commissioned risk assessments for marine spills in Canadian waters (WSP 2014a, b) as
mentioned earlier. These reports estimated the risk of pollution from marine oil spills to specific sectors
of the Canadian coast (Pacific, Atlantic, Estuary/Gulf of St. Lawrence and the Great Lakes/St. Lawrence
Seaway System). The results can be used as a valuable base for ROs to determine the level of
preparedness and develop their response strategies.

No comprehensive national framework is in place for training and exercises for ship-source oil spill
preparedness and response in Canada that involves all key stakeholders. Although Canada’s
preparedness, at the national level, is within an oil spill preparedness and response regime, it still lacks
sufficient capability to address and manage the risks existing along the coast line, particularly within the
Great Lakes and the St. Lawrence Seaway (Transport Canada 2013). This lack of capability may be the
reason why Canada’s current regime has been questioned sometimes in terms of providing the best
approach to mitigating the impacts of potential future spills (Marty and Potter 2014; Rise 2014).

7.1.6.2 Technical preparation

Sound preparedness and contingency planning should rely on knowledge of oil spills and the associated
impacts and technologies for response and cleanup. Technical preparation involves not only an
examination of spill risks, but also thorough analysis of the nature of the spill events that might occur
(source, oil type, spillage rate, location and timing) and the way in which the nature and impacts of spills
may change over space and time. For example, Silliman (2014) stated that many environmental factors
affect the fate of spilled oil sands products in aquatic environments because bitumen, a large component
of oil sands products, has a density greater than fresh water. By analyzing specific factors in areas at risk,
responders can better prepare for and expect submergence of oil sands product spills.

If areas identified have low salinity, rough sedimentation, high turbidity, strong sunlight exposure, high
temperatures, and strong currents that cause a high risk of submergence, the response teams in these areas
should have submerged oil recovery equipment readily available for rapid deployment (Soomere 2012).
According to the Joint Review Panel for the Enbridge Northern Gateway Project (JRP 2013a), the
advances in technology and modern pipeline design, materials, construction and operating practices
would greatly reduce environmental risks from pipeline spills.

A spill of any oil product could have serious effects in any environment. Compared to lighter crude oils,
heavier oils and diluted bitumen usually take longer to degrade naturally. The information about oil spill
fate, influencing factors and impacts can directly guide the development of contingency plans and the
pipelines’ design, location and management systems to minimize the potential volume released and keep
oil from reaching waterways (Ramseur et al. 2014; Taylor et al. 2014).

Insufficient information could lead to difficulties in making sound contingency planning and response
decisions. Many recent publications and conferences, such as the 2015 AMOP Technical Seminar
(Environment Canada, 2015c), have revealed knowledge gaps and research needs especially related to

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 252
the fate and behaviours of diluted bitumen, as well as the implications for response methods under
different spill and environmental conditions (especially in cold regions or waters). Also refer to the
relevant discussions in Chapter 9.

7.1.6.3 Contingency planning

Oil spill prevention planning requirements are determined by the potential source of the spill, which for
transported oil primarily include storage facilities, vessels, pipelines, rail cars and trucks. The designated
federal agency must assess the capacity of the responsible party to effectively respond to a spill, which
may include providing oversight of response plans, maintaining contingency plans at various levels and
personnel training (Ramseur et al. 2014). Transport Canada (2013) proposed an Area Response Planning
model that includes a national risk assessment, an identification of area of response, a regional risk
assessment, a response requirement, an area response plan, a certification and a continuous improvement
plan. Moving from static national standards to a risk-based model represents an important shift for the
Canadian Regime. The more rigorous planning process inherent in this model may strengthen the links
between the public and private elements of the regime and build public confidence in Canada’s ability to
respond to an oil spill. Provincial governments are also playing a more important role in oil spill
contingency planning. For example, the British Columbia Ministry of Environment (2014) recently
introduced the preparedness and response guidelines for land-based oil spills, particularly from heavy oil
pipelines. The large volume of technical information and stakeholder feedback gathered by the Ministry
suggested that world leading practices for oil spill prevention, response and recovery systems to manage
and mitigate the risks and costs of heavy oil pipelines are highly desired.

Mathematical models and computer tools have been effectively applied to facilitate contingency planning,
although significant amounts of high-quality data, computational resources and maintenance efforts are
usually required. Integration with state-of-the-art risk/readiness assessment and geomatic analysis
methods has recently been gaining interest and research efforts. For example, Etkin et al. (2011)
recommended a risk-based approach to contingency planning and spill risk management that takes into
account the complex factors on both the probability and consequence sides of the risk equation. The
approach may include relevant cost-benefit and cost-effectiveness analyses of spill response measures to
help plan for the most beneficial strategies to reduce impacts on sensitive natural and socioeconomic
resources. Etkin et al. (2013) further suggested that the determination of scope of a contingency plan
should be based on the risk of spills within the geographic area that the plan is intended to cover.
Lamarche et al. (2013) developed a new cartographic system, based on Google Earth, to support the
planning effort in the event of an incident. It allows users to access the system on stand-alone or portable
devices, providing a more efficient planning and response tool for spill responders and the responsible
party. A variety of geomatic analysis tools and GIS-aided management systems have been increasingly
used by Canadian ROs, such as Eastern Canada Response Corporation and Canadian Coast Guard. Taylor
et al. (2014) reported that, in 2011, the Regional Association of Oil and Gas Companies - Latin America
and the Caribbean developed the Oil Spill Response Planning and Readiness Assessment Manual and its
assessment tool, the Readiness Evaluation Tool for Oil Spills (RETOS™). The tool provides a general
guide to industry and governments to assessing their level of contingency planning and readiness
management in relation to pre-established criteria, which are commonly agreed upon by the involved
institutions and consider international best management practices. These criteria are not intended to
reflect or add any legal or regulatory requirements. However, their voluntary use by governments and
industry can help guide improvements in management of oil spill preparedness and readiness.

7.1.6.4 Legal requirements and assessment

The responsibility for drawing up contingency plans at a local level, for example, for an individual
facility, port or stretch of coastline, and at a larger district or national level, will be dependent on the
relevant domestic administrative arrangements. The plan holders should be involved from the outset if

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 253
plans are to be realistic and practical. Responsibility in the US for ensuring that all plans are compatible
usually falls to a national agency, either the US Coast Guard for marine shorelines or the US EPA for
inland spills (US Coast Guard 2001). In Canada, to associate industry with government, a legislative
framework has been created by an amendment of the Canada Shipping Act 2001 (Minister of Justice
2001). See Section 7.1.6.1 for details about the lead agencies responsible for various types of oil spills in
Canada.

Meanwhile, to guide and enhance response preparedness for transboundary spill events, a number of
international contingency plans and legal instruments have also been developed, such as Canada-United
States Joint Inland Pollution Contingency Plan, Canada-United States Joint Marine Contingency Plan and
the International Maritime Organization (IMO)’s International Convention on Oil Pollution Preparedness,
Response and Co-operation (OPRC) (IMO 1990; Environment Canada 1999; Martini and Patruno 2005).

7.1.6.5 Training, drills and education

Training programs should be developed for all response levels and include marine and shoreline response
teams and interested parties (US Coast Guard 2001). Regular and realistic exercises will help to ensure
that contingency arrangements function properly and that the roles and responsibilities of all parties are
thoroughly tested and understood. Equipment should be mobilized and deployed regularly to assess its
availability and performance. Such exercises also ensure that contact details and equipment listings are
current (Doerffer 2013). Plans should be periodically reviewed and, if appropriate, amended in the light
of lessons learnt from exercises or actual incidents. All those involved need to be made aware of any
changes to the plan (Fingas 2012). Nuka (2010) recommended that field exercises and drills be conducted
for training and practice purposes. They can provide valuable information on how well the response
equipment perform, the limits to the response systems posed by real-world conditions and the cooperation
among various personnel and response organizations. However, exercises are still inherently artificial as
the equipment and vessels are pre-assigned and the personnel are pre-notified or even pre-positioned. An
unannounced true spill response drill is sometimes necessary to test the operator’s response capacity to
contact, move and deploy the whole response team.

Meanwhile, education through social media is being used and should be further encouraged in order to
improve the way the community is informed and engaged during oil spills about the situation and how
decisions are made. Proper education can help the involved communities improve their attention to
preparedness, agility in responding to spills and recovering ability from spills (Merchant et al. 2011).
Johnson (2014) suggested a strategy to provide education to local communities in order to help the public
understand oil spills and their impacts.

Recommendation: Educational outreach to enhance community awareness of preparedness and


response plans should be a key part of the training programs.

7.2 Response Decision Support

The response to an oil spill usually consists of a series of dynamic, time-sensitive, multifaceted and
complex processes subject to various constraints and challenges (Chen et al. 2012). The success and
effectiveness of a response must rely on how efficiently the information (location, oil properties, weather,
currents, etc.) and response resources (devices, manpower, money, etc.) can be used and how optimally
the decisions and actions can be made. Even though the policy or regulations focusing on framework or
infrastructure are relatively consummate, inadequate decision support may be one of the major challenges
that limit the efficiency of spill response. In the past decades, many models have been developed mainly
focusing on individual oil spill response activities, including early warning and detection, spill simulation,
cost-benefit analysis, risk and impact assessment, cleanup technology selection and cleanup optimization
and performance evaluation (Christoph 2013; Lamarche et al. 2013; Paul et al. 2013; Li et al. 2014b,c;

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 254
Silliman 2014). However, to date, the integration of the above into an integrated response decision
support system is still lacking. This section will review the current status of research and practice in the
area of oil spill response decision-making.

7.2.1 Early warning

A reliable, integrated system of early warning and real-time monitoring can significantly improve the
effectiveness and efficiency of oil spill emergency response. Similar to medical diagnosis, such a system
is able to identify, diagnose and react promptly to minimize the oil discharge into the environment at an
early but critical stage of the emergency (Li et al. 2014a).

7.2.1.1 Spill imaging and analysis

Most existing satellite methods of oil spill monitoring tend to be economical, but cannot eliminate the
need of systems for early detection and direct monitoring. The integration of in situ and remote
monitoring hardware with pollution analysis software is of necessity and has been gaining attention.
Barenboim et al. (2013) proposed an automated monitoring system for oil spills in waterbodies. It consists
of a remote sensing subsystem using fluorescent LIDAR, a network of automatic monitoring stations and
an oil pollution identification subsystem based on hydrocarbon contents, alteration of radioactivity and
water conductivity. The system provides an efficient tool for early warning and monitoring of oil spills
from oil and gas facilities. Rapid distribution of aerial images helps responders at both tactical and
strategic levels. In addition, aerial imaging systems that can share information over low cost data links
greatly enhance response capabilities by providing real-time information to response teams.

Recent technology advancements enable aerial images to be communicated over low-cost data links to
stakeholders at different geographic locations within minutes of acquisition. Sweeten et al. (2012)
incorporated a proof-of-concept test into an emergency response drill that confirmed the ability to
transmit images to stakeholders and responders rapidly during oil spill response events.

Recommendation: Comprehensive and in-depth analysis of real-time spill images, along with
baseline and historic data, are highly recommended to provide more effective and meaningful
information for spill response decision-making in a timely and effective manner.

7.2.1.2 Early warning indicators

One efficient way of improving the ability to prevent and reduce major oil spill events and impacts is to
use early warning indicators. Table 7.4 categorizes the typical early warning indicators and compares
their strengths and weaknesses. The use of indicators can be seen as a regulatory requirement and/or a
means to avoid unwanted events. Investigation reports of accidents, such as Longford (Hopkins 2000) and
Texas City refinery (US Chemical Safety and Hazard Investigation Board 2007), recommended the use of
indicators. Øien (2010) compared the different approaches for developing early warning indicators and
suggested that people should be flexible with respect to the choice of methods and preferably use more
than one method. One challenge with the establishment of early warning indicators for potential minor
events is that these events are not included in the current quantitative risk analysis used in the offshore
petroleum industry (Skogdalen et al. 2011). Most existing response systems lack or are weak in effective
early warning function/ability to identify, diagnose and react promptly to minimize the oil discharge into
the environment at an early but critical stage of the emergency. Furthermore, it is challenging to
incorporate the selected indicators into the early warning and real-time monitoring systems and into the
spill response decision-making processes.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 255
Table 7.4 Summary of early warning indicators and their strengths and weaknesses
Early Description of approach Advantages Disadvantages References
warning
indicator
Safety  Describes the safety level within an organization,  Is favorable when it comes to  The risk significance Jennings and
performance- activity or work unit practicality, simplicity and and the relative Schulberg
based  Starts with a set of factors that have potential documentation importance between (2009);
indicator effects on safety  Is very relevant as an early the chosen influencing Skogdalen et
 Becomes not only useful for describing safety warning factors are unknown al. (2011);
levels, but also applicable to early warning Christoph et
al. (2013)
Risk-based  Utilizes risk models as bases, and the development  Provides indicators for major  Rather resource Khan et al.
indicator of risk models are part of the method accidents intensive, especially (2002); Pula et
 Regards risk control as the main function of the risk  Easily determines the risk for organizational risk al. (2005)
indicators significance indicators, which are
 Becomes preferred with sufficient data  Depends on either accident most relevant as early
 Particularly focuses on organizational risk investigation or occurred events warnings
indicators in the case of early warning  Indicates potential scenarios
without the occurrence of
accidents
Incident-  Depends on detailed analysis of incidents or  Closely relates to major  Requires very Jennings and
based accidents accidents thorough review and Schulberg
indicator  Assumes that if the contributing factors are efficient  Easily communicates with documentation (2009); OGP
then the incident or accident has not been analyzed stakeholders based on a factual  The risk significance (2011)
and similar ones have not occurred incident or accident and relative
 Mainly focuses on identifying and measuring the importance of the
factors that contribute to the incident or accident underlying causes are
with the use of indicators unknown
Resilience-  Questions capability of recognizing, adapting to,  Focuses on positive signals with  The risk significance Hollnagel and
based and coping with unexpected events by providing failures that may be lack of data and relative Woods (2006);
indicator specific approaches to manage risk in a proactive  Does not rely on information importance of the Øien et al.
manner from occurred events and the influencing factors are (2010);
 Indicates the engineering resilience in organizations indicators are relevant as early unknown Paltrinieri et
and safety management approach with methods, warnings al. (2012)
tools and management approaches under
complexity

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 256
7.2.2 Response technology screening and evaluation

7.2.2.1 Technology standards

The DWH oil spill revealed a variety of regulatory failures in terms of implementing a regulatory regime,
which mandates adequate safety and cleanup technologies in deepwater oil exploration (Bush 2011). To
help address the problems, Best Available Technology (BAT) standards are being advocated (Hout et al.
2014; Leifer et al. 2015). In fact, BAT standards have been adopted by some agencies since 1990 and
Alaska is among the first. In early 1990, following the EVOS incident in Prince William Sound, Alaska’s
Department of Environmental Conservation (ADEC) established regulations establishing new oil
discharge prevention and contingency planning statutes, which require BAT standards and rigorous
technical and economic assessment of BATs (ADEC 1993). However, such an assessment requirement
only applies for mechanical oil spill response equipment (Shepherd 2014).

Under Alaska state law and ADEC regulations, each blowout prevention contingency plan for marine
waters in Alaska must mandate adequate secondary relief well capacity, or similar blowout
prevention and response tools in Cook Inlet. Such plans must include BAT standards "consistent with
the applicable" statutory criteria, including: a) whether each technology is the best in use in other
similar situations and is available for use by the applicant; b) whether each technology is transferable
to the applicant's operations; c) whether there is a reasonable expectation each technology will
provide increased spill prevention or other environmental benefits; d) the cost to the applicant of
achieving BATs, including consideration of that cost relative to the remaining years of service of the
technology in use by the applicant; e) the age and condition of the technology in use by the applicant;
f) whether each technology is compatible with existing operations and technologies in use by the
applicant; g) the practical feasibility of each technology in terms of engineering and other operational
aspects; and h) whether other environmental impacts of each technology, such as air, land, water
pollution and energy requirements, offset any anticipated environmental benefits (ADEC 2015).

In promulgating a BAT regulatory standard, regulators must identify the best available technology to
adequately respond to the particular risks associated with each industrial category. The identification
should be based on scientific understanding of the effectiveness and technical/economic constraints,
especially in cold and harsh environments. Consequently, regulators must proactively investigate the
safety records of oil industry members, the safety regulations and standards, and the industries that are
interested in advancing cleanup technology. However, industry usually tends to bypass the standards to
reduce the costs (Prendergast and Gschwend 2014). Economic variance can affect the viability of a BAT
standard as increased technology requirements often necessitate increased expenditures by industry.
Therefore, the oil industry may consider replacing the BAT standard with other strategies to avoid excess
costs (Hout et al. 2014). Meanwhile, depending on the location of a spill and weather conditions, response
technologies may encounter various constraints to effectiveness (Andrade et al. 2013), especially in
Canadian harsh environments where rough weather, high seas, poor visibility, high wind, freezing
temperature and ice conditions prevail (Chen et al. 2012). There have been difficulties in selecting the
BAT according to both equipment capacity and working conditions. Laborde et al. (2015) proposed a
framework to develop an oil spill response gap map model based on high resolution historical data at
large spatial and temporal scales. The model would show how often environmental factors exceed the
operational capacity for in situ burning, booming-skimming and dispersant application in a web-based
GIS environment.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 257
7.2.2.2 Screening and performance evaluation

A spill response strategy often starts with the identification of appropriate technologies that accommodate
the site-specific environmental conditions and increase the probability of a favourable outcome.
Numerous factors must be considered for the selection of oil spill response procedures for use in
contingency plans and emergency response operations. These factors include but are not limited to the
probability of an oil spill, the possible volume, type and properties of crude oil or refined product that
might be spilled, environmental factors influencing the fate and behaviour of the hydrocarbons that could
be released, the sensitivity of the most valued ecosystem components (VECs) to oil pollution, the
potential impacts from the application of oil spill countermeasures, and the time needed for habitat
recovery (Board of Ocean Studies & Marine Board 2014). Leschine et al. (2015) introduced a scenario
analysis for oil spill contingency planning and response technology screening for vessels, pipelines, tank
farms, berthing docks and many other oil-industry infrastructures. Liao et al. (2012b) proposed an oil spill
emergency preparedness system that recommends the BAT based on a comparison with similar cases.
Still one of the key challenges in selecting and developing suitable response strategies and cleanup
technologies is insufficient knowledge about their technical limitations, influencing factors and ecological
impacts, especially related to some emerging concerns, such as diluted bitumen, aging or subsea
pipelines and spills to freshwater ecosystems and Arctic conditions. Furthermore, there is a need for new
or updated regulations to govern the application of response technologies based on our improved
knowledge. Further discussion about influencing factors and performance evaluation for response
technologies are provided in Chapter 6 and later in this chapter.

7.2.3 Spill physical and numerical simulation

Besides the gaps in modeling oil spills as discussed in Chapter 5, challenges also exist in how to use
modeling tools to more effectively and dynamically support response decision-making (Chen et al. 2012;
Li et al. 2012a, 2014b; Li 2014). The linkage between modeling of oil slicks (weathering and trajectory)
and cleanup processes (e.g., in situ burning and skimming) is lacking, as well as their dynamic coupling
with ecological/environmental risk assessment to provide decision-makers more accurate and
comprehensive information. Furthermore, harsh environments prevailing in northern regions and the
Arctic can pose a large amount of uncertainty and difficulty in using the traditional physical and
numerical simulation models and their integration with the corresponding response simulation and
decision support systems.

7.2.4 Net Environmental Benefit Analysis (NEBA)

Net Environmental Benefit Analysis (NEBA) has become widely used to assess oil spill countermeasures,
both active (e.g., in situ burning and mechanical recovery) and passive (e.g., monitoring of natural
attenuation processes). NEBA can help determine whether or not additional environmental damage could
be caused by specific response actions (Fritt-Rasmussen et al. 2013). NEBA can provide decision-makers
with a reliable strategy for deciding what response options are appropriate at a specific spill location
based on the analysis of environmental trade-offs that may occur from the use of the various responses
(Coelho et al. 2013; Coolbaugh et al. 2014).

The generic NEBA framework is outlined in IPIECA (2000). A typical NEBA process involves: a) review
of previous spills and experimental results that are relevant to the area and to possible countermeasures;
b) assessment of likely environmental outcomes if the proposed countermeasures are implemented
compared to outcomes if the area is left to recover naturally; and c) comparison and weighing of
advantages and disadvantages of various potential responses with those of natural cleanup. The NEBA
process can be used to establish the most important resources at risk before or in an oil spill based, for
example, on status as protected species, ecosystem service or ecological relevance, economic value or
human use (Le Floch et al. 2014; Prince 2015).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 258
The determination of priorities will depend on the preparation of a list of local environments at risk, the
results of predictive oil weathering and trajectory models, and in situ or remote monitoring data, etc.
NEBA accounts for the nature of the spilled oil and changes it may undergo during weathering and
spreading, which may influence the level of environmental, biological and socioeconomic effects (Daling
et al. 2014). However, NEBA can only process environmental variables and does not account for other
considerations in the strategy-forming process. A strategy that is environmentally or ecologically ideal
but does not recognize political or legal factors or operational feasibility may not be ultimately
implemented (Coelho et al. 2013). This should be better addressed in the NEBA.

NEBA tends to be an integral part of contingency plans because post-spill decisions are best and most
quickly made in light of pre-spill analyses, consultations and agreements involving all of the appropriate
organizations and parties (Daling et al. 2014). Use of NEBA in contingency planning offers several
advantages, including extended timeframes for analysis, consideration of spill scenarios covering a wide
range of environmental factors (e.g., seasonal changes in species diversity and ice cover), time for
identification and collection of scientific data and stakeholder involvement (Cox 2014). NEBA can also
facilitate the identification of potential conflict areas and possible solutions through consultation before
any spill occurs. Other important considerations during NEBA include the logistical constraints that are
likely to be encountered during oil spill response operations, which will influence the efficacy of current
countermeasure strategies (Bejarano and Mearns 2015).

Cumulative impacts refer to the effects of an action that are added to or interact with other effects in a
particular place (e.g., an oil spill site) and within a particular time (US EPA 1999). The combination of
these effects and any resulting environmental changes caused by response actions should be considered in
a cumulative impact analysis, which has been incorporated into environmental impact assessment
processes. When introducing cumulative impact analysis to NEBA for oil spill response, difficulties may
be encountered when identifying/quantifying cumulative impacts of response actions (e.g., application of
dispersants), especially for long time periods (Paul et al. 2013; Davies and Hope 2015).

Regardless of what and how response methods are applied, there will be some level of environmental
impact. Many studies have conclusively shown that the application of aggressive cleanup operations may
delay the rates of habitat recovery by causing additional damage beyond the oil spill itself (Judson et al.
2010; Kirby and Law 2010). For example, in the aftermath of the EVOS incident, excavation and washing
of rocks to remove surface and subsurface oil were shown not to offer a net environmental benefit
because the procedures altered shore structure and delayed biological recovery (US NOAA 1990). The
recent DWH spill involved around 47,000 people and 7,000 vessels during in situ burning, mechanical
recovery and application of dispersants, yet long-term side effects of those activities, especially some
caused by the use of dispersants, have been reported (Graham et al. 2011). In fact, a response strategy
that provides protection for one environmental resource (e.g., chemical dispersion of oil slicks to protect
seabirds) may increase risks to another (e.g., toxicity of dispersed oil to fishes in the water) (Graham et
al. 2011; Payne et al. 2012; Paul et al. 2013). Decision-makers need to select the optimal response
strategy based on the protection of prioritized environmental resources and the countermeasures that
offer them the greatest protection, which always leads to some level of trade-off. The value (effectiveness
and impact) of a particular method will depend on the situation, including the type and properties of oil
spilled, weather conditions, organisms and ecosystems that are impacted and availability of response
support.

A NEBA process can help identify the relevant factors affecting the effectiveness of options and select the
best response strategy. However, prioritizing response options in NEBA usually involves human inference
and judgement in making the trade-offs (Prince 2015). Qualitative evaluation approaches, including
questionnaire surveys, are usually used and unavoidably introduce uncertainties into the results due to
incomplete information and subjective judgement based on personal knowledge, experience and opinion
(Reynolds 2014).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 259
Recommendation: More unbiased measures and better understanding of the influence of human
judgement are required in NEBA and subsequent decision-making processes. Furthermore,
socioeconomic factors play an important part in response decisions and should be reflected in the
NEBA process in terms of impact on a region’s environment, its economy and the wellbeing of its
people.

Many oil spill countermeasure technologies (e.g., chemical agents including dispersants and facilitation of
oil mineral aggregate formations) may not be fully understood in terms of their environmental and
ecological impacts and their effectiveness in harsh environments such as ice conditions (Payne et al.
2012; Coelho et al. 2013). Until sufficient information is available on the population dynamics of VECs,
the fate and environmental effects of oil and the efficacy of various response technologies, the
effectiveness of NEBA will vary and outcomes will be uncertain (Fitzpatrick et al. 2012; Coolbaugh et al.
2014). However, application of NEBA is still necessary to assess some newly emerged concerns (e.g.,
dilbit), countermeasures (e.g., use of biosurfactants as dispersants) and special environmental conditions
(e.g., ice coverage, permafrost and peat lands).

In terms of damage assessment, an ecosystem services approach has been suggested (Menzie et al. 2012;
NRC 2013). It can account for an event’s impact on all aspects of human wellbeing and can help
remediate the damage to natural resources caused by oil spills. Chapters 4 and 8 present the definition and
detailed discussion about ecosystem services modeling.

In addition to the contingency planning stage before a spill, NEBA has also been used to choose the
intensity of the response countermeasure, as well as the decision on when to conclude the response
(DeMicco et al. 2011). Decisions are typically made based on potential environmental and socioeconomic
impacts of the spill, the potential impacts of the cleanup operation and the possibility and ability of
natural or human-enhanced recovery of the impacted environment. The 2012 report of the American
Petroleum Institute (API), Spill Response in the Arctic Offshore (Potter et al. 2012), identified where a
range of different cleanup intensities and conditions might be appropriate. It was noted that the most
intense efforts to remove oil are appropriate in areas with high levels of human activity or environmental
sensitivity, the risk from oil is high and the impact from intense cleanup measures is low. By contrast,
monitoring may be sufficient in situations where the spilled oil poses little risk, the potential for natural
oil degradation and natural recovery are high and the risks posed by cleanup efforts are high. Another
example from the API report showed that, where areas adjacent to the spill site are highly sensitive to oil
and the oiled area is only moderately sensitive to cleaning, the appropriate cleanup efforts may include
removing most of the visible oil but allowing traces of oil to remain.

7.2.5 Performance evaluation and post-event management

7.2.5.1 Performance evaluation and impact assessment

The oil types, applied response methods, performance and environmental impacts of some major oil spills
in Canada and the US are summarized in Table 7.5. The summary presents significant variations of
performance and impacts of similar response options in different spill events. Post-event comparable
studies are lacking to help develop more comprehensive quantitative methods for response performance
evaluation in a holistic manner. Consequently, lessons from past spill responses should be better analyzed
and reflected in planning and decision-making processes for future spill responses. Also, effective
environmental monitoring and impact assessment practices should be in place, using the most
appropriate scientific techniques, to enable the response, conservation and scientific communities to
properly assess the real significance of spills and subsequent response activities and to learn lessons for
future events. The requirements for improved onsite and post-incident monitoring and assessment
continue to increase. Although most risks can usually be estimated during or right after spill response,
the variability of the situation may require some degree of onsite or post-event assessment to continuously

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 260
assess the risk of oil spill and response activities (Leong 2012). Such variability can include, but not be
limited to, weather changes and subsequent work activities, etc. Onsite and post-event assessment usually
includes visits to key risk areas, direct observation, data collection and analysis, questionnaire survey and
interview and long-term site monitoring (Dix et al. 2011; Shupe et al. 2014). Bostrom et al. (2015)
applied a mental models survey approach to assess public thinking about oil spills and oil spill response.
The decision model considered controlled burning, public health and seafood safety. The results
illustrated opportunities to reframe discussions of oil spill response in terms of trade-offs between
response options, and new possibilities for assessing public opinions and beliefs during events.

7.2.5.2 Adaptive management

Adaptive management, as a systematic process, has been widely used to continually improve management
policies and practices by adjusting subsequent actions based on both past lessons learned and advanced
knowledge from new scientific and socioeconomic information on the systems being affected (Rost et al.
2014; Kulapina et al. 2015). The key steps usually include learning and reducing uncertainties, using
learned knowledge to change policy and practice, focusing on improving management effectiveness,
applying trial management and being formal and systematic (Cluzel et al. 2012; Hausberger et al. 2012).
The application of adaptive management, especially with integration of risk analysis, has been limited in
the oil spill response field. Peltier et al. (2014) developed a new approach that used multiple indicators,
including oil slick drifting, to provide monitoring information on dolphin populations and to setup
adaptive management strategies. Adaptive management should be further encouraged to continually
improve management policies and practices by adjusting subsequent actions based on learning from past
events and new knowledge development.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 261
Table 7.5 Representative examples of oil types, applied response methods, performance and environmental impacts of some major oil spills in Canada and
the U.S.
Oil Spill Case Oil Type Date Amoun Responses and Performance Key Impacts Referen
t (US ces
gallon)
Nova Scotia Mud, gas and February 126,000  Well crews evacuated  Strong hydrocarbon odour was Gill et al.
Uniacke G-72 Oil condensate 22, 1984  No immediate response present as far as 10 km away (1985)
Spill, Nova  Well recapped 10 days later  Oiled seabirds and seals were
Scotia, Canada observed
 No evidence of condensate on the
shoreline
Exxon Valdez Oil Crude oil March 24, 11-38  Controlled burning (with ship and  Immediate effects - the deaths of Morris and
Spill, Alaska, 1989 million boom) - 113,400 L of oil reduced to 100,000 to as many as 250,000 Loughlin
USA 1,134 L of removable residue seabirds, at least 2,800 sea otters, (1994);
 Mechanical cleanup (with boom and and many other animals Boehm et
skimmer) - only 10% oil recovered  Long-term effects: more losses of al. (2014)
 Skimmer: low efficiency (thick oil, species than expected
heavy kelp, plenty of permanent  Bioremediation was somewhat
container, bad weather, etc.) effective in accelerating
 Chemical dispersants ineffective due to biodegradation as a cleanup option
lack of wave energy
 Shoreline washing
 Bioremediation of shorelines
Pine River Oil Light crude oil August 1, 260,400  118,800 gallons removed from the river  Extensive mortality of fish and Anderson
Spill, British 2000  Vacuum oil and set up containment some wildlife and
Columbia, booms of 12, 24 and 28 miles from the  Water supply to the District of Schwab
Canada rupture Chetwynd was ceased and the use (2012);
 119,560 gallons removed from soil of many groundwater wells nearby Chung et
 Removal of large woody debris from the river was discontinued al. (2012)
river channel due to it being coated with  Disruption of the river channel and
oil destruction of fish habitat by
seasonal flooding following the
removal of woody debris that
stabilized the river morphometry
Terra Nova Oil Waxy crude November 43,650  5% oil recovered  An estimated 10,000-16,000 direct McGrath
Spill, 2004 kills of birds (2014)
Newfoundland,
Canada

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 262
Wabamun Lake Heavy Bunker August 3, 343,000  185,000 gallons of Bunker C and  Residents were told to stop cleaning Newbrey
Oil Spill, Alberta, C fuel oil 2005 Imperial Pole Treating Oil removed wildlife and to completely avoid use et al.
Canada  Early response by sorbent booms of all lake water and well water, (2012);
(ineffective) even for watering gardens and Martin et
 Containment boom (inadequate lawns al. (2014)
quantity)  Access to the lake for recreation
 Later response by SCAT process and and fishing was limited for months
cleanup of individual shore segments following the spill
based on geomorphology and degree of  Extensive damage to reed beds
oiling caused by oiled vegetation removal
 Extensive removal of shoreline oiled and recovery of oil and tar balls
reed bed vegetation and mechanical  Mortality of reed bed fish and
recovery of oil from reed bed water wildlife and reduction of nesting
column and sediments habitat for western grebe
 Residual oil and tar balls in
shoreline and marsh areas for
several years
 Evidence of oil contamination of
whitefish spawning shoals at
concentrations sufficient to impair
embryo survival
Deepwater Crude oil April 20, 210  33 million gallons recovered  Most impacts were on marine Allan et al.
Horizon Oil Spill, 2010 million  Initially: ROVs, 700 workers, 4 species – by November 2, 2010, (2012);
Gulf of Mexico, airplanes and 32 vessels 6,814 dead animals, including 6,104 White et
USA  Involved 47,000 people and 7,000 birds, 609 sea turtles, 100 dolphins al. (2012);
vessels in total and other mammals Barron
 Chemical dispersants - Corexit (1.07  Actual number of mammal deaths (2012);
million gallons on water surface & due to the spill may be as much as NRC
721,000 US gallons underwater) 50 times higher than the number of (2013)
 In situ controlled burning (13 million recovered carcasses
US gallons)  Environmental concerns based on
 Over 60 open water skimmers evidence of oil transport to
(33,000,000 US gallons or 120,000 m3 sediments
of tainted water removed)  Beach erosion and disruption of
 Extensive shoreline remediation, plant and animal life-cycles were
including shaping shorelines to direct oil observed due to oil toxicity and
away from sensitive areas and oil cleanup actions
recovery from beaches and coastal
wetlands

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 263
Kalamazoo River/ Cold Lake July 26, 819,000  Over 1.1 million gallons of Line 6B oil  Eighty miles of shoreline and Dollhopf
Enbridge Pipeline Blend (with 2010 recovered adjacent lands were contaminated and Durno
Spill, Michigan, benzene  Excavation and dredging  Submerged oil was assessed and (2011);
USA diluent) recovered at over 25 locations McGowan
 Over 100 residents were relocated et al.
due to air quality (benzene) (2012)
concerns
 Residual oil in sediments producing
sheen when disturbed
 Extensive severe habitat damage
and destruction due to removal of
riparian vegetation, construction of
roads on riparian lands up to the
river banks, removal of islands that
had been over-washed by oily
water, and continual disturbance
and removal of contaminated
sediments
Lac-Mégantic Bakken oil July 6, 26,455  Removed oil and damaged rail cars from  Heavily contaminated with benzene Ecosocialis
Rail Disaster, 2013 downtown Lac-Mégantic causing heat and toxic conditions t Network
Quebec, Canada  Excavation of 57,250 m3 of  Most-contaminated areas might (2013);
contaminated soils never be habitable Lacoursièr
 49,494,000 L of oily water and oil were  30 km stretch of the Chaudière e et al.
recovered and treated till December 19, River has residual oil in sediments (2015)
2013 and along the banks
 Run-off water management - oil and
water recovery trenches and pumping
stations
 About 100,000 L spilled into the
Chaudière River – shoreline cleanup
Gogama Oil Spill, Synthetic March 7, Over  Containment booms installed  The spill contaminated a nearby Canadian
Ontario, Canada crude oil 2015 264,000  Skimmers and vacuum trucks deployed lake, river and soil National
 The municipal water system was Railway
impacted by oil Company
(2015)

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 264
7.2.6 Risk assessment

7.2.6.1 Spill risk assessment

Risk assessment underpins all preparation and


planning for oil spill response and includes the
The main components of a relative risk
assessment of both the likelihood of a spill occurring
assessment of oil spills are:
and the consequences or effects caused by a spill
 Estimation of the probability of various
(Maritime New Zealand 2006). Substantial research
sizes of spills based on past and projected
has been done and progress made to develop analysis
future incident rates;
tools and evaluate risks of oil spills across the world
and especially in many maritime countries, such as  Vulnerability of the environment to oil
Canada, US, Norway, United Kingdom and Australia spill impacts; and
(Turner 2010). Refer to the discussion about oil spill  Selection of the oil spill scenario(s) to be
risks and case studies in Chapter 8. assessed; e.g., maximum most probable
discharge and/or worst-case discharge.
WSP (2014a,b) have examined the potential
frequency of spills in Canadian waters and the The assessments are conducted on a regional
potential consequences associated with these spills. and seasonal basis by oil type. The level of
Potential spill frequencies were estimated using a detail used depends on the spatial and temporal
combination of Canadian and worldwide spill scope of the assessment and the availability of
statistics. Risk values generally increased in coastal data.
areas and the overall highest risk was observed for
small size spills due to their high frequency. Much of the conceptual foundation for these methods is
provided by the Washington Compensation Schedule (Washington Administrative Code 173-183), which
provides a relative impact score for oil spills, considering sensitivity of the locations oiled, relative
density and seasonal distributions of sensitive biota, and factors related to oil type. The methods used in
these reports are similar to those developed and demonstrated by Etkin (2012a), Reich et al. (2014) and Li
et al. (2014a).

The general approach used by WSP (2014a,b) in their assessment of spills in Canadian coastal waters was
as follows:

 Coastal waters were divided into four main sectors which were in turn divided into smaller subsectors
for a maximum number of 77 zones;
 Shipping densities, as well as vessel types and size distribution, in each zone were estimated;
 Oil spill frequencies for ships were obtained from the most recent 10 years of worldwide data;
 The behaviour of oil spills (surface area over time) was estimated from simple transport and fate
models which depend on the oil type, the spill size and location characteristics;
 The environmental sensitivity index (ESI) was calculated based on physical, biological and human
metrics, which were further mapped to illustrate their spatial distribution in each zone; and
 The overall ESI was determined using a spreadsheet calculation and mapped to present its spatial
distribution using GIS tools (Figure 7.6).

The key findings of the WSP (2014a) report on spills in southern Canada were as follows:

 The largest marine traffic volumes are in the Pacific sector where the probability of small size fuel
spills is the highest. The zones with the highest probability of a large spill occurring were the waters
around the southern tip of Vancouver Island, the Cabot Strait including southern Newfoundland, the
eastern coast of Cape Breton Island and the Gulf of St. Lawrence and the St. Lawrence River;

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 265
 ESI results showed that the zones of highest potential impact were in the Estuary and Gulf of St.
Lawrence, as well as in the southern coast of British Columbia including Vancouver Island. Higher
ESI scores were observed in near-shore zones compared with intermediate and deep sea zones;
 The highest relative risk values were for small spills due to their higher frequency of occurrence. The
risk of large spills is generally low in Canada. The risk generally increases in near-shore zones
compared with deep sea zones with the exception of the Pacific sector where US traffic may increase
deep sea probabilities. The increase in risk in near-shore zones is related to an increase in
environmental sensitivity; and
 The Estuary and Gulf of St. Lawrence, the St. Lawrence River, the southern coast of British
Columbia, as well as three subsectors in the Atlantic sector, are at the greatest risk from large oil
spills. For the rest of the study area, the risk posed by spills over 10,000 m3 were much lower. Risk is
higher from small and medium spills in every sector, especially spills in the 100 to 999 m3 range.
These smaller spills can also cause significant damage and are likely to happen much more frequently
than the larger spills.

These results can be used to tailor spill preparedness for each sector and subsector. It would be useful, as
a first step, to evaluate the current level of preparedness under the Marine Oil Spill Preparedness and
Response Regime in light of relative risks in order to identify deficiencies.

WSP used a similar methodology to assess the relative risk of oil spills in Canadian Arctic coastal waters
(WSP 2014b). Physical Sensitivity Indicators (shoreline type, wetlands, ice coverage), Biological
Resource Indicators (ecologically and biologically significant areas and bird distribution and Human-Use
Resource Indices (coastal population, tourism, national/international freight tonnage) were combined to
produce an overall ESI for each of 18 subsectors. Spill frequencies were very low compared to southern
Canada because of the current low amount of marine tanker traffic. The higher ESI scores occurred in the
southern Arctic subsectors, including in the south of James Bay, the south of the Beaufort Sea and at the
Mackenzie River Delta. Because the current probability of oil spills is low, relative risks were also low
throughout all subsectors.

The WSP assessment of oil spills in Arctic waters did not include prediction of future risks due to
increased traffic associated with the opening of routes due to climate change. The assessment also did not
appear to account explicitly for the difficulty inherent in cleanup of Arctic spills, which would potentially
produce more long-term risk.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 266
Figure 7.6 Overall ERI for crude oil spills South of the 60th Parallel. Image from WSP (2014a)

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 267
Etkin (2012a) and Li et al. (2014a) proposed new modeling approaches to calculate the probability and/or
vulnerability of oil spills by considering environmental conditions, shipping routes, incident cause, vessel
type and/or nearby ecological reserves with uncertainties reflected in the system. The approaches were
applied to the Cook Inlet vessel traffic and the southern Newfoundland offshore areas, respectively, where
concerns of oil spillage exist due to shipping operations and potential impacts on the sensitive
environment of the area.

Recently, Reich et al. (2014) assessed the spill risk in Alaska/Arctic and identified the three highest
relative risk regions within the scope of their study for maximum probable spills and worst case spills as
shown in Table 7.6. They pointed out that the relative risk model was highly data-intensive. Therefore,
the quality of the results was dependent on the quality of input data. They cautioned that some
environmental vulnerability data inputs were known to be of poor quality and should be updated. In
particular, bottom habitat and submerged aquatic vegetation data coverage were lacking for much of the
area assessed. Sensitivity testing of the model suggested that the number of species used for the study was
sufficiently robust, but the addition of more species could refine the risk scoring. This would be
important, for example, if species of high traditional use value to aboriginal people were a priority for
assessment. Reich et al. (2014) also pointed out that incident rates and potential spillage volumes
forecasted for 2025 were subject to considerable uncertainty. Notwithstanding the limitations of the
assessment, the results identified broad regions of Alaska with high relative risk, with implications for
spill prevention and preparedness.
Table 7.6 Summary of the relative risk rank and regions. Data from Reich et al. (2014)
Relative Risk Max Probable Worst Case Max Probable Worst Case 2025
Rank Current Risk Current Risk 2025 Risk Risk
1 Southeast Alaska Southeast Alaska Beaufort Sea Beaufort Sea
2 Aleutians Kodiak/Shelikof Aleutians Aleutians
Strait
3 Kodiak/Shelikof Cook Inlet Southeast Alaska Southeast Alaska
Strait

7.2.6.2 Environmental/ecological risk assessment

The general environmental/ecological risk calculation is expressed as the probability of spills multiplied
with the potential impacts on environmental or ecological systems (US EPA 1999, 2004; NRC 2013). In
general, the first step of risk assessment is hazard identification, which determines the qualitative nature
of the potential adverse consequences caused by the action or contaminant. The second step is dose-
response analysis, which determines the relationship between dose/frequency of action and the probability
or the incidence of effect (dose-response assessment). The third step is exposure quantification, which
measures the amount of a contaminant (dose) that individuals and populations will receive. And the last
step is risk management, including coordinated and economical application of resources to minimize,
monitor and control the probability and/or impact of risks (Board of Ocean Studies 2014; Li 2014).

Many assessment methods and practices for oil spill events have been reported. For example, French-
McCay (2002) developed and validated an oil toxicity and exposure model (OilToxEx) for estimation of
impacts to aquatic organisms resulting from acute exposure to spilled oil. The author further proposed a
coupled oil fate and effects model for the estimation of impacts to habitats, wildlife and aquatic organisms
resulting from acute exposure to spilled oil. The physical fate model estimated the distribution of oil (as
mass and concentrations) on the water surface, on shorelines, in the water column and in the sediments.
The biological effects model estimated exposure of biota of various behaviour types to floating oil and
subsurface contamination, resulting percent mortality and sublethal effects on production (somatic
growth). Impacts were summarized as areas or volumes affected, percent of populations lost and

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 268
production foregone because of a spill’s effects (Liu et al. 2009). Recently, Etkin (2012b) developed a
simple and defensible model to provide a relative quantitative measure of the environmental and
socioeconomic impacts of a pipeline or facility spill as a function of key location-specific attributes. The
model was intended to reflect significant differences in the sensitivity of different types of environments
to spill-related damage without requiring a complex site-specific environmental impact assessment.
Stephenson et al. (2015) studied hypothetical spills of dilbit in the Kitimat River and concluded that the
potential for chronic ecological risk was the greatest for wildlife species occupying the floodplain or
riparian habitat and having relatively small home-range sizes.

7.2.6.3 Health risk assessment

Health risk assessment is used to provide individuals with an evaluation of their health risks and quality
of life. The main objectives include assessing human health, estimating the level of health risk and
providing feedback to participants to motivate behavioural change to reduce risk. The results can reflect
the impacts of oil spills and response activities on human health and therefore aid in decision-making as
to whether or not and to what degree responses should be conducted (Crosby et al. 2013; Nicholls, et al.
2014; Bostrom et al. 2015).

Baars (2002) estimated the health risk for people involved in beach cleaning, sunbathing and swimming
on the coast of Brittany, France, due to the wreckage of the oil tanker Erika in 1999. The results
suggested that the risks were limited only to people who had bare-hand contact with the oil. Aguilera et
al. (2010) reviewed the possible consequences of exposure to spilled oil on human health. Barenboim and
Saveka (2012) introduced the biological impacts of individual petroleum hydrocarbons into the
determination of the risk of oil spills for biota and humans. Michigan Department of Community Health
(2014) evaluated the health risk associated with an oil spill on the Kalamazoo River and found that
chemical levels found in surface water were not expected to cause long-term harm to people’s health. It
was also deemed that the concentrations of hydrocarbons found in fish from the Kalamazoo River and
Morrow Lake would not harm people’s health. Yee et al. (2015) carried out a Human Health Risk
Assessment (HHRA) to assess both potential acute and chronic risks to human receptors in the unlikely
event of a full bore pipeline break. The model was tested by a hypothetical spill with the proposed
pipelines used for transporting oil extract from the Alberta oil sands to refineries or to coastal terminals
for international export. The results showed that risk management strategies targeting the protection of the
exposure pathways with relatively high contribution to human health risk (e.g., consumption of below
ground and root vegetables and fish ingestion) during the remediation would be important to reduce the
risk.

An emerging concern has been raised on the psychosocial impacts of a major oil spill. Grattan et al.
(2011) used a community-based participatory model to conduct the standardized assessments of
psychological distress (mood, anxiety), coping, resilience, neurocognition and perceived risk on residents
of fishing communities who were indirectly impacted (n = 71, Franklin County, FL) or directly exposed
(n = 23, Baldwin County, AL) to coastal oil. They found the residents of both communities displayed
clinically significant depression and anxiety but no significant differences between community groups.
They further pointed out that current estimates of human health impacts associated with the oil spill might
underestimate the psychological impact in Gulf Coast communities that did not experience direct
exposure to oil. Income loss after the spill might have a greater psychological health impact than the
presence of oil on the immediately adjacent shoreline. Gill et al. (2014) compared the psychosocial
impacts caused by the 2010 DWH oil spill and the 1989 EVOS based on surveys of local residents
collected a certain period of time (12 and 18 months, respectively) after each of the spills. The analysis
revealed similarly high levels of psychological stress, tied to renewable resources, concerns about their
economic future, worries about air quality and safety issues regarding seafood harvests in oiled areas.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 269
It is valuable to incorporate a comprehensive examination of ecological and human health risks of oil
spills into contingency planning and response decision-making (Cao and Fan 2013; Crosby et al. 2013;
Bostrom et al. 2015). It would allow spill contingency planners and decision-makers to determine optimal
risk-based response strategies, calculate the costs and benefits under different scenarios and choose the
best scenario (or make the optimal decisions) to guide response actions (Gala et al. 2009; Etkin et al.
2011; Etkin 2012a; Fernandes et al. 2013). Linkov et al. (2006) suggested that decision analytical
methods could incorporate risk assessment into integrated frameworks. A key attribute of the methods is
the ability to engage stakeholders, providing for an opportunity for a shared understanding of the issues,
and the importance of particular assumptions and model or data results on the assessment of alternatives
(von Stackelberg 2013).

7.2.7 Cleanup process simulation and control

Process control is defined as an engineering discipline that deals with mechanisms and algorithms for
maintaining the output of a specific engineering process within a desired range (Jing et al. 2012, 2015).
For oil spills, the knowledge and prediction of dynamic response processes to the variations of
environmental conditions and operational factors are critical to ensure an optimal operation of the
response process. A clear understanding of the mechanism of a response process (e.g., booming, in situ
burning, skimming, dispersion and bioremediation) will help to quantify the direct relationships among
the inputs (e.g., number and types of skimmers) and outputs (e.g., recovery rate), as well as the indirect
relationships, such as the time-series correlation (Jing et al. 2015). Modeling of oil behaviour, effects and
fate, along with the influence of spill response measures (e.g., skimming or dispersion), has been
recognized as an essential component and foundation of successful process control strategies (Li et al.
2014b). A numerically simulated response process can provide the decision-makers with a dynamic
means to optimize oil spill recovery and cleanup operations. Particularly with the real-time aid of
process simulation and control tools during spill response actions, the response efficiency and
effectiveness can be promoted and the overall time and cost of recovery can be minimized.

7.2.7.1 Gaps in cleanup process modeling

In dealing with an oil spill event, responders always require fast and accurate estimates of the spill to
make timely and effective decisions in deploying skimmers, applying dispersants or conducting other
response activities before, during and after the spill event. However, complexity and dynamics inherently
exist during oil spill response processes emanating from the fate and transport of spilled oil, existing
climatic and oceanic conditions, operation of cleanup equipment, as well as their interactions, all of
which cause significant challenges in response decision-making. Traditional physics-based spill models
are weak in providing a good solution. A few studies have explored the possibility of simulating oil
recovery processes based on empirical oil weathering models and artificial equipment performance
settings (Buist et al. 2011; El-Zahaby et al. 2011; You and Leyffer 2011; Zhong and You 2011), but
technical and knowledge gaps persist in how to obtain accurate and timely forecasting results under
varying environmental conditions. This is especially true in harsh environments where the window of
opportunity for spill response is significantly shorter, and more efficient decisions and actions are highly
recommended. Chen and his group (Jing et al. 2012; Li et al. 2012a, 2014b,c, 2015) proposed a set of
simulation-based dynamic nonlinear programming approaches to optimize the number and operation of
oil skimmers during a spill in a fast, real-time and cost-efficient manner based on the consideration and
modeling of oil weathering processes. Nonetheless, the modeling of response processes when using
different cleanup techniques, such as booming, skimming, in situ burning, dispersant application and
bioremediation, still face many challenges due to the lack of background data, high nonlinearity and
various uncertainties from oil properties and weather conditions.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 270
7.2.7.2 Special considerations for harsh environments

Cold and harsh environments in northern Canada and the Arctic are usually characterized by a wide
range of wind speed and direction, limited visibility, low temperature, rough water surface, ice coverage,
etc., posing substantial challenges for oil spill cleanup process simulation and control (Oskins and
Bradley 2005; Keller and Clark 2008). For example, considerable amounts of fixed or floating ice may
occur in the winter in both inland waters and offshore areas (Cleveland 2010). The presence of ice is a
key factor affecting the ability to respond to a spill, as well as modeling efforts (DeCola et al. 2006). The
physical distribution and condition of spilled oil under, within or on top of ice plays a critical role in
determining the most effective response strategies at different stages in ice growth and decay cycle. As
pointed out by the National Energy Board of Canada’s State of Knowledge Review of Fate and Effect of
Oil in the Arctic Marine Environment (Lee et al., 2011), understanding the influence of these
environmental variables/factors and the influence of various levels of ice coverage on oil fate and
behaviours and the associated environmental impacts is needed. A vast amount of basic information is
available from previous laboratory, mesocosm and field studies, as well as lessons learned from past
spills. However, knowledge gaps still exit on the fate and behaviour of oil in water in ice conditions
(solid, slush and frazil ice) and during active periods of formation and breakup of annual and multi-year
ice, as well as the impacts on ecosystems. Consequently, the effects of applications of different response
methods under Arctic conditions are hard to forecast and manage in a more timely, eco-friendly and cost-
effective manner. The recent report by Arctic Oil Spill Response Technology Joint Industry Programme
(AOSRT JIP 2014) provided a comprehensive review in this regard.

In-depth knowledge and modeling tools for oil fate and transport and effective response technologies are
lacking for oil spills in ice and harsh conditions. Existing models have insufficient ability to accurately
predict how oil and ice interact and then how oil transforms and moves where ice is present. Most
mathematical equations used to simulate oil cleanup/recovery are likely based on empirical
approximations and assumptions and are subject to time step and grid limitations. Furthermore, current
response technologies are either ineffective or require more scientific and field validation in dealing with
oil in ice conditions. These challenges should be addressed to support more effective response decision-
making and operations.

Recommendation: Research is needed to develop physical and numerical models for simulating,
predicting and optimizing response processes and evaluating their individual and collective effects
on response decision and effectiveness, particularly in terms of accuracy and adaptability to
different environmental conditions. The dynamic links with spill modeling and decision-making
should also be realized.

7.2.8 Response operation optimization

The response to an oil spill is a dynamic, time-sensitive, multifaceted and complex process subject to
various constraints and challenges. It is dependent on a variety of factors, including quantity and
properties of the spilled oil, location, environmental conditions and availability and utilization of response
resources at various degrees of oil weathering (Ornitz and Champ 2003). Response operations usually
need to be undertaken within a limited time window and improper decisions may compromise the
efficiency of oil recovery and waste resources. It is highly desirable to develop and implement an
optimized strategy to better coordinate different types of operations. You and Leyffer (2011) developed
an approach for oil spill response planning with integration of the physiochemical evolution of oil slicks.
The approach includes a dynamic oil weathering model and can simultaneously predict optimal coastal
protection plans and oil spill cleanup schedules with different types of mechanical, burning and dispersant
application equipment. The results demonstrated the importance of integrating an oil transport and
weathering model in response planning. Zhong and You (2011) addressed the optimal planning of oil spill
response operations under the constraints of economic and responsive criteria, with consideration of oil

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 271
transport and weathering processes. A multi-criteria, multi-period linear programming model was
developed that minimizes the total cost and response time-span and simultaneously predicts the optimal
time trajectories of the oil slick’s volume and area, transportation profile, usage levels of response
resources, oil spill cleanup schedule, and coastal protection plan. Li et al. (2012a, 2014c) proposed a
simulation-based nonlinear optimization approach to provide sound decisions for skimming spilled oil in
a fast, dynamic and cost-efficient manner, which especially targeted harsh marine environments. The
model was further integrated with the optimization method to determine the optimal strategy to achieve
the maximum oil recovery within constraints of time and resources. The approach was tested using
hypothetical case studies showing its capability of efficiently incorporating the process simulation and
optimization into the same modeling framework to improve skimming effectiveness. Efforts are being
made to incorporate these approaches into existing modeling software or new decision systems with user-
friendly interfaces to become effective tools for onsite use and training by responders.

Recommendation: In general, response decision-support systems are rare and only present limited
degrees of integration between oil spill modeling, cleanup process simulation and/or risk
assessment, but their dynamic and interactive features need to be significantly enhanced, and more
field validation efforts are urgently required to increase the confidence for implementation.

Uncertainty, which is one of the major hindrances to improving the efficiency of cleanup process
simulation and control, may arise from a variety of sources. Such sources include, but are not limited to,
incomplete information, errors in sampling, subjective judgement, random variations of and dynamic
interactions among operating factors, approximations and assumptions in measurement, and changes of
environmental conditions (Huang et al. 2001; Lin et al. 2009; Liu et al. 2009). Uncertainties lead to
difficulties in developing optimization models for supporting decision-making in oil spill response and
impair confidence of decisions.

7.2.9 Information technologies for response decision making

7.2.9.1 Geomatics, remote sensing, and geographical information systems

Geomatics (or geomatic technology or geomatic engineering) is a hybrid field of gathering, storing,
processing and presenting geographic information or spatially referenced information (Zhu et al. 2013;
Song et al. 2014). The techniques of geomatic analysis, remote sensing and GIS have been widely used in
oil spill response. The Central Gulf of Mexico Ocean Observing System (CenGOOS), along with remote
sensing and GIS, were adopted in monitoring and forecasting the DWH spill (Howden et al. 2011; Li et
al. 2012b). Morović and Ivanov (2011) discussed the properties of synthetic aperture radar (SAR)
imagery as the most reliable source of oil spill information and the possibilities of a combined SAR-GIS
approach for oil spill monitoring and management for the protection of the Adriatic Sea. Pan et al. (2012)
used multi-source remote sensing data, including MODIS data, and advanced SAR images to study the
aftermath and delineate the extent of the 2006 Lebanon oil spill. Dollhopf et al. (2014) introduced a
multiple-lines-of-evidence approach for assessment and recovery of submerged oil by integrating with a
robust GIS tool. The approach helped optimize assessment and recovery, inform decision-making related
to transitions and endpoints, and improve cleanup efficiency while limiting the long-term ecological
damages from the recovery activities.

7.2.9.2 Data mining and artificial intelligence

Knowledge Discovery (KD) effectively uncovers hidden but subtle patterns from large and diverse
datasets and outperforms traditional statistical techniques. Data mining, a major stage in the KD process,
is the analysis of datasets that are observational, aiming at finding out hidden relationships among
datasets and summarizing the data in a manner that is both understandable and useful to the users. Many
artificial intelligence (AI) tools have been used for data mining, such as neural networks, fuzzy logic, ant

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 272
colony algorithms and genetic algorithms. Given these features, data mining and AI tools have received
increased interest and efforts in addressing the complexity and dynamics of oil spill and response
operations and seeking optimal solutions for risk assessment and decision-making (Cai et al. 2009;
Akinyokun and Inyang 2013). Some new methods, such as fuzzy neural network (Cao and Fan 2013;
Inyang and Akinyokun 2014) and predictive Bayesian model (Echavarria-Gregory and Englehardt 2015),
have been developed and applied to assess oil spill risks and characterize their patterns.

7.2.9.3 Results visualization

Preferably, an integrated oil spill response decision support system would include visualization of not
only past and current situations, but a dynamic probability envelope for the future spill positions and the
associated risks, time and costs under different response scenarios, much as is done in the
hurricane/typhoon and flooding forecasting community (Sanyal et al. 2010). However, present oil spill
visualization systems are generally based on producing either a snapshot map of representative oil spill
particle trajectories or a movie of an evolving point cloud of particles (Kang et al. 2013; Hou and Hodges
2014). Furthermore, as geomatic technologies become more powerful and usable over desktops and even
mobile platforms (e.g., smart phones, tablets), oil spill visualization systems should use GIS standard
formats for output data to allow web-based access for emergency response personnel (Yu and Yin 2011).
Standardization within a GIS can allow spill trajectories to be linked to existing index systems (e.g.,
environmental or ecological sensitivity index) that classify sensitive coastal areas by their degree of
exposure and vulnerability.

Fernandes et al. (2013) described recent updates of the oil and inert spill modeling component of MOHID
model, including interfacing with meteorological and oceanographic data and the integration with
decision support systems. This system revealed a good stability and strong performance in visualization
of weathering processes and properties. The rich image resources of satellite remote sensing Google Earth
was selected as the client during constructing the system. The system solved data exchange and data
storage of KML between Google Earth and Oracle by connecting them, which realized the visualization
of the oil spill system. Hodges et al. (2015) introduced methods to integrate existing oil spill models,
servers, connections to online data services and visualization tools to support response decision-making.
To improve deepwater oil spill emergency response, Wei et al. (2015) built a high-resolution
hydrodynamic data field and a 3D visualization system based on oil behaviour and fate.

The oil spill response community will greatly benefit when current technical challenges and knowledge
gaps are overcome to enable the coupling of advanced modeling and visualization techniques to address
questions such as:

 How to integrate a dynamic spill modeling system with remote sensing, GPS and GIS to turn
real-time information into real-time decisions and then visualize them;
 How to combine oil fate and transport modeling, environmental risk/impact assessment and
cleanup process simulation and optimization with information technology to achieve the best
response efficiency while minimizing time and costs; and
 How to visualize the levels of uncertainties within models to evaluate their risk.

7.3 Summary

This chapter reviewed the current development and practice of oil spill prevention and response decision-
making in Canada and beyond. Prevention and preparedness have been emphasized as the key factors to
proactively reduce the potential damage of oil spills. With a high level of knowledge, many potentially
high-risk activities can be anticipated and may be mitigated or prevented. Cutting-edge monitoring and oil
analysis technologies, as well as advanced vulnerability assessment and contingency planning methods,
can help identify and even prevent unexpected oil spills in an efficient, timely manner. Baseline data of

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 273
environmental and ecological conditions before spills take place in areas of high spill probability are
critical in support of spill preparedness and response decision-making. A well-structured response regime,
an established spill contingency plan and existence of active programs for training, drills and education
are critical for a country or region to effectively combat oil spills and minimize the damage to the
environment and danger to human health.

Responses to an oil spill can be highly dynamic, time-sensitive, multifaceted and complex due to
uncertain or changing information (location, oil properties, weather, currents, etc.) and response resources
(devices, manpower, money, etc.). A successful response operation usually starts with an accurate
detection of an oil spill through imaging analysis and early warning. The selection of BATs and the
precise prediction of oil trajectory to combat the oil spill also play important roles in terms of saving cost
and time. NEBA and risk assessment tools usually offer the decision-makers quantitative information to
evaluate the damage and to prioritize the response technologies. Cleanup processes should be simulated
and optimized simultaneously in order to improve the response effectiveness within a limited time
window. With the aid of advanced techniques, such as geomatic analysis, AI and computer visualization,
oil spill response decision support systems should be capable of not only integrating spill and cleanup
modeling and risk assessment but recommending, and graphically presenting in the most simplistic
format, optimal decisions and corresponding impacts under different scenarios to responders and
stakeholder before or during a spill. In addition, post-event assessment is recommended to deal with
changing conditions, such as weather and subsequent work activities. It should be noted that the efforts of
integrated oil spill response decision-making have been limited due to the complex, dynamic and
uncertain nature of oil spills. More efforts toward the improvement of oil spill preparedness and decision-
making are highly desirable, particularly for some emerging challenges, such as dilbit spills, derailments
and harsh environments.

7.4 Research Needs and Recommendations

7.4.1 High Priority Research Needs

 Research is needed to develop modeling methods to simulate and optimize individual and collective
cleanup processes (e.g., booming, in situ burning, skimming, dispersion and bioremediation)
individually and collectively and estimate their effects for supporting response decision-making.
 Research is urgently required on development and demonstration of oil spill response decision
support systems, which can dynamically and interactively integrate monitoring and early warning,
spill modeling, vulnerability/risk analysis, response process simulation/control, system optimization
and visualization.
 Research investment is needed on trial tests and field validation of new prevention and decision-
making methods to demonstrate feasibility, increase confidence for implementation and improve
response capabilities.
 Research is needed to better quantify modeling uncertainties, evaluate their propagation and mitigate
their impacts on spill response decision-making.
 Further research and development are desired on environmental forensics, remote sensing and in situ
measurement, early warning and diagnosis, and biological monitoring to improve spill prevention and
decision-making.
 Special attention of the above research should be given to harsh and Arctic environments and some
other emerging concerns (e.g., diluted bitumen, aging/subsea pipelines and railcars) to enhance
effectiveness and confidence of prevention and response strategies and decisions.

7.4.2 Medium and low priority research needs

 Innovative research should be strongly encouraged to develop more cost-efficient and eco-friendly
response technologies. Meanwhile, there is a need to update regulations to govern the use of response

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 274
technologies (e.g., dispersants) based on improved knowledge and better understanding of their long-
term impacts, especially for Arctic and subsea applications.
 Oil leak detection systems need to be further developed to detect spills rapidly after an incident. More
comprehensive and reliable in situ and remote detection methods are needed. Research and
implementation of autonomous, remotely operated detection devices (e.g., AUVs and ROVs) should
be encouraged. Advance of early warning methods and integration with monitoring/detection systems
are necessary to identify, diagnose and react promptly to minimize oil discharge into the environment
at the early and critical stage of an emergency.
 Analysis of cumulative impacts remains a challenge for NEBA of oil spill response options. Influence
of human judgement and socioeconomic factors needs to be more effectively reflected in the NEBA
process. There is also a need to assess NEBA’s effectiveness with emerging issues of concerns (e.g.,
Dilbit spills), countermeasures (e.g., biodispersants) and special environmental conditions (e.g., ice
coverage, permafrost and peat lands) to identify knowledge gaps for decision-making.
 Minor oil spills, such as those caused by leakage from routine operations or treatment/discharges of
oily wastewater (e.g., produced water and bilge water), should be given more attention in terms of
cumulative effects and in situ treatment and recovery methods.

7.4.3 Additional recommendations for response operations and preparedness

 A national guidance program for post-spill monitoring is needed to collect reliable, adequate, credible
and consistent information on the fate and effects of oil in the environment.
 It is important to develop long-term effective methods to support the monitoring of oil spill impacts
and the fate of released oil. The establishment of a defensible baseline dataset for shipment corridors
within high-risk areas is recommended. One possible place to start might be a national directory of
monitoring information with links to databases, GIS maps and reports.
 It is critical to properly assess the real significance of spills and subsequent response activities and to
develop lessons learned in preparation for future events. The requirements for improved onsite and
post-incident monitoring and assessment should be better defined. Adaptive management, as a
systematic process, is encouraged to continually improve management policies and practices by
adjusting subsequent actions based on learnings from past events and new knowledge development.
 A comprehensive national framework is needed for training and drilling exercises for ship-source oil
spill preparedness and response that involves all key stakeholders and unannounced spill response
drills to test the operator’s response capacity to contact, move and deploy the whole response team.
 Education through social media is being used and should be further encouraged in order to improve
the way the community is informed and engaged during oil spills about the situation and how
decisions are made. Educational outreach to enhance literacy and promote preparedness and
responses should be a key part of the training.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 275
CHAPTER 8: RISKS FROM OIL SPILLS

Abstract

This chapter provides a review of the important factors affecting the consequences of selected oil spill
case studies and presents lessons learned from each spill. Uncertainties and research needs revealed by
each spill case are presented. The strengths and weaknesses of relative risk assessments are discussed and
recommendations are made to improve the current practice of risk assessment in Canada.

A review of selected oil spill cases illustrated that each case had a unique combination of site-specific
factors. There were different combinations of factors that either increased or decreased the overall
impacts of each spill. The conclusions arising out of examination of the case studies are listed below.

 Delayed response was a critical factor affecting the consequences of all the oil spill case studies.
Notwithstanding the importance of weather, remote locations and technological challenges,
human error (at an individual and organizational level) was a dominant factor across all case
studies. Absent or inadequate planning, insufficient integration, inadequate training, poor
communication, insufficient capacity (personnel and equipment), poor or no information sharing,
and lapses in regulatory oversight were noted for most, if not all, spill case studies.
 The consequences of oil spills cannot be predicted simply on the basis of the type of crude oil,
although in general, spills of light crude had greater acute effects. Long-term effects were
dependent upon a combination of the type of crude oil, characteristics of the receiving
environment (including season) and effectiveness of cleanup.
 The lack of pre-spill baseline data seriously limits the ability to predict or monitor long-term
effects of crude oil spills on populations and communities of aquatic organisms.
 Major advances in the development and validation of spill response technologies are seriously
limited by the inability to conduct controlled field experiments with oil.
 Standard accepted methods have not been established for post-spill monitoring and validating the
efficacy of oil spill response protocols and operational endpoints (How clean is clean?).
 Effects of oil spill cleanup on aquatic ecosystems can be significant.
 Aquatic ecosystem resilience to shocks from oil spills is the key to the maintenance of ecosystem
services. The presence of long-term stresses may tax the capacity of ecosystems to be resilient to
shorter-term shocks.

A review of relative risk assessments for coastal areas of Canada and Alaska and risk assessments done in
support of Environmental Impact Assessments (EIAs) generated the following conclusions:

 The data needed for input to risk assessments in Canada are often either absent or widely
scattered among government agency, industry and academic sources;
 Information needed for reliable environmental sensitivity indices is very limited for large portions
of Canada;
 Input from Indigenous peoples and interested parties is required to produce reliable and credible
information to be used for environmental sensitivity indices;
 Based on case study results, the assessment of risks from pipeline spills have been largely
underestimated due to the assumption that it would take only a few minutes to recognize a
pipeline breach and to deploy shut-off valves; and
 The level of sophistication of risk assessments conducted in support of EIAs varied substantially
between regions.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 276
High-priority recommendations arising out of these conclusions are:

 Collect and evaluate baseline information from high-risk, poorly-characterized areas in Arctic
coastal areas and in freshwater systems (e.g. inland rivers with multiple pipeline crossings, the
Great Lakes, the Gulf of St. Lawrence);
 Conduct controlled field experiments on oil spills at a variety of sites for the development and
verification of current and emerging spill response countermeasures;
 Conduct long-term research into effects of different crude oil types on populations of aquatic
biota, including follow-up research at the site of old spills;
 Conduct a program of research on resilience of aquatic ecosystems;
 Develop national guidance for monitoring of oil spills to ensure that information gathered is
reliable, adequate, credible and consistent. The guidance should include provisions for adjustment
in response to specific characteristics of the receiving environment and should provide details on
operational endpoints which identify when cleanup and remediation can cease;
 Include studies of the relative effectiveness of response measures in all investigations of
significant oil spills in Canada;
 Pre-approved funding should be put aside for research and monitoring of ‘spills of opportunity’;
 To support the development of improved spill response guidelines, build a comprehensive
national database for the fate, behaviour and effects of various types of oil spilled and the efficacy
of current and emerging oil spill countermeasures over a range of environmental conditions;
 Investigate the socioeconomic impacts of oil spills in support of assessment of effects on
ecosystem services;
 Conduct regional risk assessments in areas of concern to provide further guidance for policy and
planning;
 Establish a ‘standard of practice’ regarding the assessment of risk of oil spills in support of EIA
in Canada which includes updated and refined assessment methods and a re-examination of
appropriate spill scenarios; and
 Integrate traditional knowledge and information about traditional uses of resources into
development of sensitivity indices used for risk assessment.

Introduction

To cover the range of potential spill scenarios that Canada may encounter in the future, this chapter
examines the risks arising from oil spills into aqueous environments in two main subsections: oil spill
case studies and reviews of risk assessments of hypothetical spills.

The case studies clearly illustrate that each spill incident is a unique combination of site-specific factors.
These factors include: the type of oil; the mechanism of release; the volume of the spill; the
weather/hydrology conditions at the time of the spill; the vulnerability of the habitats; the sensitivity of
the biota; and the effectiveness of applicable spill response options. The impact of a spill is often
influenced by the degree to which the oil reaches shorelines and sediments, as well as the physical-
chemical characteristics of the oil (e.g., tar balls, liquids trapped under cobbles, oil mixed with sand,
‘pavement’, etc.). The interactions of various physical-chemical forms of oil with shorelines and
sediments influence environmental persistence, bioavailability, toxicity and net environmental benefit of
countermeasure strategies applied. In each of the case studies and risk assessments reviewed in this
chapter, different combinations of factors either increased or decreased the overall impacts.

The case studies were selected to present combinations of geography, oil type and environmental factors
that influenced the response, outcome and knowledge gained:

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 277
 Marine environments:
 The Arrow tanker spill of heavy fuel oil in Chedabucto Bay, NS.
 The Baffin Island Oil Spill (BIOS) Project involving deliberate release of crude oil into
the Arctic environment.1
 The Exxon Valdez tanker oil spill (EVOS) of medium crude oil into cold, pristine
conditions with intensive shoreline intervention, in Prince William Sound, AK.
 The Deepwater Horizon (DWH) spill of light crude into the deep subsurface and
shoreline environments of the chronically oil-impacted Gulf of Mexico, with
unprecedented subsurface application of dispersant.
 Freshwater environments:
 A pipeline spill of sour light crude into the Pine River, BC.
 A derailment spill of a type of Bunker C oil (HFO 7102) into Wabamun Lake, AB.
 A pipeline spill of dilbit into Talmadge Creek and the Kalamazoo River, MI.
 Recent Canadian spills in 2015:
 The Marathassa Intermediate Fuel Oil (IFO) spill in English Bay, Vancouver, BC.
 Nexen Long Lake pipeline bitumen emulsion spill near Ft. McMurray, AB.

For each case study, the important factors affecting the consequences are identified. Lessons learned are
then listed. Finally, a list of research needs illustrated by the case study are identified and presented in
bold.

The risk subsection reviews: 1) predictive relative risk assessments conducted in support of policy and
planning (Transport Canada’s review of Canada’s preparedness and response for ship-source spills); 2)
risk assessment in support of EIAs prepared as part of the regulatory approval process for proposed
projects (e.g. the Northern Gateway Pipeline); and 3) risks from crude-by-rail transport. Reviews included
evaluation of the risk assessment methods used, the key factors contributing to risk, and the primary
uncertainties associated with the prediction of risk from oil spills. Recommendations for the improvement
of risk assessment based on the Panel’s findings are presented in bold.

8.1 The Arrow Spill

On February 4, 1970, the aging tanker Arrow encountered severe weather as it entered Chedabucto Bay,
NS. She ran aground on Cerberus Rock, a well-known navigational hazard. Severe weather hampered
attempts to off-load her cargo of Bunker C crude oil (Table 2.2) and on February 8, the tanker split into
two releasing about two-thirds of her cargo (10,000 tonnes of oil) into the icy waters of Chedabucto Bay,
much of which took the form of an oil-in-water emulsion that affected 300 km of coastline. Retrieval of
the remaining oil from the wreck was completed on April 11, 2015, using pioneered techniques
subsequently used in other tanker accidents (Maritime Museum of the Atlantic 2015).

The Arrow spill provides a long-term case study of reliance on natural attenuation (defined in Chapter 3).
Only 48 km of the approximately 305 km of oiled shoreline were cleaned (Owens 2010). Cleanup
primarily involved manual and mechanical removal of oiled coarse sediments, typically contaminated to a
depth of 25 - 100 cm and occasionally as much as 200 cm (Figure 8.1). The average volume of oiled
sediment generated by manual cleanup was approximately an order of magnitude less than mechanical
removal using bulldozers and front-end loaders; this had implications for waste management. In terms of
long-term impacts, mechanical removal of oiled beach material resulted in retreat of the beach crest at
some locations for periods up to 10 years (Owens 2010).

1
In addition to the information summarized in this report, the Panel notes a recent report by the Arctic Oil Spill Response Tech-
nology Joint Industry Programme (AOSRT 2014), which provides a comprehensive literature review and recommendations re-
garding the assessment of environmental impacts of Arctic oil spills.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 278
Figure 8.1. Testing the use of lime during shoreline cleanup of the Arrow spill. Image courtesy of Ed Owens.

Time-series measurements of petroleum-derived hydrocarbons in the water column of Chedabucto Bay


demonstrated that concentrations returned to pre-spill levels within a year (Levy 1972). Time-series
measurements in 1970, 1973 and 1976 showed a dramatic reduction in shoreline oiling, with 75% of the
heavily oiled shoreline cleansed by 1973, even in sheltered areas with low wave energy (Owens 2010).
By 1992, approximately 250 km of oiled shoreline were cleaned naturally, of which about 70 km were
sheltered environments where physical processes such as wave action and abrasion would have played
only a minor role in weathering and degradation (Owens et al. 1993; Owens 2010). The coastal waters of
northern Chedabucto Bay are rich in clays and these clays can interact, albeit slowly, with shoreline oil
suggesting that the formation of clay-oil emulsions (also referred to as clay-oil flocculation, oil-mineral
aggregates or OMA (Chapters 2 and 6) played a significant role in the natural removal of shoreline oil and
subsequent biodegradation (Owens 2010).

Twenty years after the spill, Vandermeulen and Singh (1994) studied two beach sites in Chedabucto Bay
and found that both contained a spectrum of degraded petroleum residues, from seemingly unaltered to
severely weathered residues. The authors noted that the relatively permeable nature of the Chedabucto
Bay sediments permitted stranded oil to become sequestered and protected within shore sediments
(Figure 8.2). The authors concluded that persistence was a direct function of beach sediment
permeability, sediment grain size and the depth to which entrapped tar residues penetrate. Persistence was
an inverse function of the depth and frequency in which sediments were reworked during tidal incursions.
The authors noted that the oil residues trapped within the spaces of protected sand-gravel lagoons and
deep cobble/boulder beaches would be the primary long-term source of hydrocarbon movement to pore
water and surface water. The chemistry of some of the sequestered residues still closely resembled the
original spilled oil (Vandermeulen and Singh 1994). In contrast, the surface-stranded Bunker C residues
progressively weathered to tar via exposure to oxidation, photooxidation (Chapters 2 and 6) and physical
weathering by wave action, sediment scouring and tidal washing.

Thirty-five years later, most of Chedabucto Bay was oil-free, but the areas with surface oil distribution
had changed little since mapping in 1992, and some areas of Black Duck Cove (a low-wave energy site)
still had residues on coarse-sediment beaches and within fine-grained sediments at an adjacent protected
lagoon (Figure 8.2, Owens et al. 2008). Exposed highly biodegraded and photooxidized surface asphalt
pavements observed in the upper intertidal and supratidal zones in 1992 were being eroded by wave
action at slow but observable rates. Subsurface oil that filled the pores below the boulder surface layer
was associated with the presence of finer sediments, limiting the downward migration of the oil. Analysis
of samples from these environments showed that the oil close to the surface was extensively degraded
relative to subsurface oil recovered at depths beneath the surface layer of boulders, cobble and fine
sediment where the composition of residual oil was similar to the original spilled material. The authors

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 279
suggested that the more deeply buried residues in the
cobble-boulder areas would likely remain until the High wave energy may strand oil in the
sediment was disturbed by major storms as a result of supratidal zone but also removes the oil
landward barrier migration. These observations on oil on shores by erosion. A hurricane
persistence supported those from a previous study in the removed a sandy beach in Nova Scotia
1990s (Vandermeulen and Singh 1994). Owens et al. being studied by researchers, and winter
(2008) noted that residual oil in the fine-grained sediment storms may have been the most effective
of the lagoon area of Black Duck Cove was extensively mechanism for the removal of stranded oil
biodegraded. The authors suggested that this indicated on rocky shorelines following the EVOS
that if unweathered oil is released from the cobble areas (K.Lee, pers. comm.)
into the lagoon, it would be rapidly degraded.

Figure 8.2 Comparison of oil on the shoreline immediately after the Arrow oil spill in Chedabucto Bay, NS and
thirty-five years later. Left: Cobble/boulder beach in 1970 immediately after the spill. Image courtesy of DFO.
Right: Interstitial oil residue in Black Duck Cove in 2005. Image from Owens et al. (2008)

The shorter-term effects of the Arrow spill were acute fauna mortality from exposure to the more toxic
(and shorter-lived) components of the oil, as well as physical smothering of waterfowl. Certain deposit-
feeding polychaetes had a high tolerance for living in sediments contaminated with the Bunker C oil, and
their feeding activity accelerated the weathering rate of oil through various processes, including
bioturbation and digestion (Gordon et al. 1978a). Other intertidal organisms appeared to be more
susceptible. Thomas (1978) observed that within the first six years after the spill, effects included
extensive mortalities of the attached algae Fucus on rocky shores and mortality of clams and marsh
grasses in lagoons. Species diversity was generally lower at oiled sites.

The longer-term ecological risks of the Arrow were largely a function of the weathering and release of
hydrocarbons from the oil sequestered within beach materials. Longer-term effects were investigated by
Lee et al. (1999) in 1993 and 1997, when oil was evident as asphalt pavements, stain on rocks, surface
sheen and liquids trapped within sediments at some low-energy sites, such as the lagoon in Black Duck
Cove.

Lee et al. (1999) conducted tests at a representative site in Black Duck Cove to determine the toxicity of
residual oil to fish (winter flounder), invertebrates (amphipods, sea urchins and grass shrimp) and
bacteria. They found that while much oil remained in the sediment, it was of low toxicity in most cases
(except for significant mortality in an amphipod survival test) as it had undergone significant
biodegradation (Lee et al.1999) and photooxidation processes (Prince et al. 2003). Habitat recovery was
indicated by field observations of diverse benthic invertebrate communities within the sampling area (Lee
et al. 2003), including green crabs (Carinus maenus), soft-shelled clams (Mya arenaria), common
periwinkles (Littorina littorea), amphipods (e.g. Gammarus oceanicus) and worms (e.g. Nereis sp.,

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 280
Arenicola sp.). Exposure to elevated concentrations of bulk remaining oil within the sediment at the site
to the flora and fauna has been reduced by burial at depth and dilution by storm events over the years.
Tissue hydrocarbon concentrations in Mya had declined to levels where residents were harvesting clams
from the study area, part of which had become a public park.

Wang et al. (1994) noted that after 22 years of weathering in the environment, most saturated
hydrocarbons and polycyclic aromatic hydrocarbons (PAHs) in Arrow oil samples had been lost. The
analysis of biomarker compounds by GC/MS were required to link residual oil samples to the Arrow
cargo oil as traditional fingerprint methods could not provide detailed information of the source,
characteristics and fate of the spilled oil.

In conclusion, the Arrow spill provides a valuable example of the fate and effects of spilled oil and the
relative merits of cleanup versus natural attenuation for heavily oiled coarse-grained beaches in mid-
latitude environments. Natural processes, including the interaction of oil and fine mineral particles that
enhanced the physical dispersion processes, were effective in significantly reducing shoreline oiling such
that residual risks appear to have been minimal.

8.1.1 Important factors affecting the consequences of the Arrow spill

8.1.1.1 Oil properties and behaviour

 Severe weather and turbulence resulted in both rapid dispersion of the oil and deposition onto
shorelines and lagoon areas.
 Formation of OMAs contributed to relatively rapid natural attenuation of shoreline oiling.

8.1.1.2 Effect of the environment on fate and behaviour

 The presence of cobble/boulder beaches and lagoons with fine sediments provided areas where
oil was sequestered and persisted.
 Storms buried and then re-exposed some of the spilled oil.
 Oil under aerobic surface conditions biodegraded at a much faster rate than oil sequestered within
sediments with limited oxygen availability and water/nutrient exchange.
 Bioturbation by benthic organisms enhanced oil biodegradation rates.

8.1.1.3 Oil toxicity

 Acute mortality was associated with exposure to the more toxic (and shorter-lived) components
of the oil, as well as physical smothering.
 Reductions in chronic effects were correlated to the weathering of the residual oil (including
biodegradation) and the episodic release of buried oil within beaches as the result of storm events.

8.1.1.4 Spill response

 The application of active spill response operations was limited by the lack of availability of spill
response measures and logistic constraints linked to inclement weather conditions.
 Mechanical removal of oiled sediments generated up to 10 times more solid waste than manual
removal.
 Removal of large volumes of coarse-grained sediment without replacement resulted in beach
retreat.
 Oil was naturally removed in low-wave-energy environments where fine particles created OMAs
that enhanced physical dispersion and biodegradation of the oil.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 281
 The persistence of residual oil was generally inversely correlated to shoreline energy level (e.g.
from waves, tides and ice) with oil remaining longer in low energy environments.

8.1.2 Lessons learned from the Arrow spill

 Delayed response due to adverse weather and oil spill response preparedness increased the
volume (due to water-in-oil emulsification processes) and extent of spilled oil.
 Over time, natural attenuation can effectively mitigate the environmental impacts caused by oil
spills in the marine environment.
 Socioeconomic effects can be both short- and long-term; short-term effects include loss of
income among fishermen. Long-term effects are largely aesthetic, and these can cause loss of
amenities (e.g., beach use) and tourist income.

8.1.3 Research needs related to the Arrow spill

Recommendation: Further assessment of cumulative effects of the residual oil in Chedabucto Bay
should be conducted. The assessment should evaluate species and indicators relevant to all stressors
of concern (from all sources) within a defined study area appropriate to the current and predicted
future location of stressor sources. The assessment should establish clear thresholds of effect for
each indicator such that the incremental contribution of the residual oil to overall cumulative risk
can be better understood.

Recommendation: An assessment of the residual effects of the Arrow spill on ecosystem services
would provide useful information for incorporation into net environmental benefit analysis
(NEBA), as well as input to future risk assessments.

8.2 The Baffin Island Oil Spill Project (BIOS)

The BIOS project was conducted from May 1980 to August 1983 in the eastern Arctic at Cape Hatt, on
the northern end of Baffin Island (Sergy and Blackall 1987). Information from the study and others, such
as the Balaena Bay study2 were intended for planning, approval and operational phases of hydrocarbon
development in the Arctic (Lee et al. 2011).

The BIOS Project consisted of a shoreline study and a near-shore study that involved controlled releases
of crude oil. Its primary objective was to determine if the use of dispersants in the Arctic near-shore
would reduce or increase the environmental effects of spilled oil and to determine the relative
effectiveness of other shoreline protection and cleanup techniques. The secondary objective was to
determine the chemical and physical fate of oil in Arctic near-shore and shoreline areas (Sergy and
Blackall 1987).

The shoreline of the study area consisted of steep, rocky promontories separating coarse sand, gravel and
cobble beaches. Open water occurred for about 65 days per year. Tides ranged from 1 to 2 m and the mid-
summer mean maximum air temperature was about 7 oC (Sergy and Blackall 1987). Pre-spill analysis of
water, sediment and benthic fauna confirmed that there were very low background petrogenic
hydrocarbon concentrations (Cretney et al. 1987a, b, c).

Three bays of similar coastal morphology and sedimentology were selected as test areas. Oil fate was
monitored in the intertidal and shallow subtidal zone in the water column, intertidal beach sediments,
subtidal sediments and the tissue of selected benthic invertebrates (Boehm et al. 1987; Humphrey et al.
1987a, b; Owens et al. 1987). The measurement of biological effects focused on the small and less mobile

2
Conducted in 1974-1975 with the principal objectives of assessing the impact of an offshore oil well blowout on the thermal
regime of the Beaufort Sea and developing potential countermeasure techniques (NORCOR Engineering Research Ltd. 1975).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 282
subtidal benthic flora and fauna (Sergy and Blackall 1987). The experimental approach was not suitable
for the direct measurements of effects on fish, sea birds and marine mammals.

8.2.1 The near-shore study

The near-shore study compared the consequences of dispersing an oil slick close to shore with the
consequences of allowing the oil to beach and leaving it to natural attenuation. Two experimental
discharges of 15 m3 of weathered (8% by weight) sweet medium crude oil (Venezuela Lagomedio) were
released at each site over a period of six hours or half a tide cycle. Corexit 9527 was the dispersant used
in a volumetric dispersant/oil ratio (DOR) of 1:10 (Lee et al. 2011).

The surface slick release began at high tide and under the influence of an onshore wind. The oil was
carried up and deposited on the beach by wind and wave action as the tide receded. The oil remaining on
the water surface after completion of the full tidal cycle was collected with skimmers. Booms were
deployed to prevent cross-contamination between test and reference areas and were left in place for
several weeks to contain oil sheen (Sergy and Blackall 1987). The dispersed oil release was achieved
using a subsea diffuser pipe placed on the bottom just outside the study area. The currents provided the
mixing and movement patterns required to distribute the oil throughout the study area (Sergy and Blackall
1987).

The untreated surface slick was stranded in the intertidal zone in a manner representative of an accidental
spill of moderate severity (Sergy and Blackall 1987). Relatively stable conditions were reached after
about 48 hours, when oil on the beach was no longer being refloated in large quantities. During this 48-
hour period, about one-third of the oil was recovered from the water surface by skimming, one-third
remained stranded on the beach and one-third was evaporated or dissolved in sea water (Sergy and
Blackall 1987). Over the next two years, the original amount of stranded oil was reduced by about 70%.
Sergy and Blackall (1987) commented that this rate of natural shoreline cleaning was surprisingly high,
given the short period during which it could occur and the protected nature of the shoreline.
Biodegradation was considered to have removed only minor quantities of stranded oil; the majority of oil
removal was attributed to physical processes. Oil
residues were highly visible on the beach after two
A pycnocline is a boundary layer separating
years; however, the distribution of oil was very
two layers of different densities. A large
patchy. The majority of the oil remaining on the
density difference between surface waters and
beach was incorporated into an asphalt pavement.
deep ocean water prevents vertical mixing.
Some of the oil from the beach was transported by
Formation of pycnoclines may result from
runoff and wave action to the adjacent subtidal
changes in salinity or temperature. Because the
sediments; however concentrations were still
pycnocline zone is stable, it acts as a barrier for
relatively low after two years (Boehm et al. 1987).
surface processes.
In situ biodegradation could not be confirmed;
therefore it was considered a negligible factor in the
removal of sedimented subtidal oil at the low concentrations encountered (Sergy and Blackall 1987).
Concentrations of oil in the water column were low beneath the initial slick and only trace amounts were
measured over the subsequent two years.

The subsea discharge of oil/seawater/dispersant was swept by currents through the designated study areas.
Oil contacted bottom sediments and organisms from the shoreline to a depth of 15 m. Within a few days
the oil was distributed throughout the surface waters of Ragged Channel, including the reference site. The
subtidal benthos received average exposures of 300 mg/kg, 30 mg/kg and 3 mg/kg in the two study Bays
and the reference Bay, respectively (Boehm et al. 1987). Concentrations of oil in the water column were
diluted to near background levels within a period of days. There was minimal recoalescence of oil on the
water surface and negligible oiling of intertidal sediments.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 283
Conditions in the immediate vicinity of the subsea release caused benthic organisms to be exposed to high
concentrations of toxic aromatic hydrocarbons, which would otherwise have been rapidly lost to surface
evaporation (Sergy and Blackall 1987). The dispersed oil was introduced at greater water depths than
would likely be the case of a surface oil slick. Sergy and Blackall (1987) noted that in many Arctic coastal
areas, the often present and well-pronounced pycnocline would suppress the downward mixing of oil to
productive benthic depth zones. The high, short-term exposures to the chemically-dispersed oil produced
acute behavioural and physiological effects in a wide variety of animals and, in a few species, a short-
term reduction in abundance (Sergy and Blackall 1987). Responses included the emergence from the
sediment and/or immobilization of infaunal and epibenthic invertebrates. Recovery and reburial occurred
in the two weeks following the spills. No behavioural changes were observed at the other study sites,
which were exposed to very low levels of dispersed oil.

Heterotrophic bacteria in the water column were not affected by the surface oil slick and only temporarily
affected by the chemically-dispersed oil. Effects on bacteria living in subtidal sediments were also minor
(Bunch 1987).

There was no major large-scale mortality of benthic infauna to either oil release. After the two-year post-
spill monitoring period, there was no evidence of large-scale mortality of subtidal benthic biota
attributable to either the chemically-dispersed oil or the oil-contaminated beach. There were few changes
in the populations or community structure of infauna, epifauna or macroalgae (Cross and Thomson 1987;
Cross et al. 1987a; Cross et al. 1987b). There were no measured biological effects in the intertidal zone
due to the absence of Arctic intertidal life.

There were indications that exposure to the persistent oil residues in subtidal sediments (from both spills)
was responsible for medium term (1-2 year) sublethal effects, but only in a few species. For example,
there were effects on condition in the bivalve Macoma calcaerea and decreases in density of the
polychaete Spio spp. (Cross and Martin 1983).

Exposure to even low concentrations of chemically-dispersed oil resulted in a rapid and significant uptake
by subtidal benthic fauna, particularly the filter-feeding bivalves (Humphrey et al. 1987a).
Bioaccumulation was observed over a relatively large area due to the extensive vertical and horizontal
movement of the oil cloud. The water column was only a short-term source of oil exposure. Body burdens
of filter-feeding bivalves were reduced within two weeks and those of all organisms were reduced
considerably within the first year. On a longer term but very localized basis, deposit-feeding benthos
living in contaminated sediments had elevated body burdens two years post-spill. In some cases an
uptake-depuration balance appeared to exist.

Indirect effects of bioaccumulation were not part of the scope of the BIOS Project. Sergy and Blackall
(1987) suggested that tainting may occur in shellfish harvested for human consumption. They also
suggested that bioaccumulation may be of ecological significance in concentrated feeding areas for birds
and mammals that use benthic biota as a food source. They stated that critical habitat for resource species
(e.g. fish, birds and mammals) should be protected from oil of any type, and particularly from floating
and beached oil slicks.

8.2.2 The shoreline study

The shoreline study evaluated options for cleanup of oiled shorelines by comparing cleanup methods to
natural self-cleaning. Small plots (20-40 m2) were used in the intertidal and backshore zones of the test
beaches (Sergy and Blackall 1987). The same Lagomedio crude was used as for the near-shore study. Oil
was deposited mechanically. Both emulsified and non-emulsified oil were used in paired plots (Figure
8.3). After application of oil, the plots were left for 24 hours prior to initiation of cleanup tests. This was
assumed to represent minimal response time.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 284
Figure 8.3. Shoreline study test plots during the BIOS project. Image courtesy of Ed Owens.

Wave energy was the dominant factor in the removal of oil from exposed beaches. On the partially
exposed beach, the majority of oil was removed within the first open water season and by the end of the
second season less than 0.03% remained.

Rising tides removed large quantities of oil on sheltered beaches over the first two days, after which
relatively stable conditions prevailed. Oil was less persistent in the fine-grained sediments than on the
pebble-cobble beach, where oil residues were still visible four years post-spill (Sergy and Blackall 1987).
Climate and soil conditions played major roles in the fate of oil in the backshore plots. Significant
amounts of oil remained in surface and subsurface sediments after three years; however there was
considerable change in the composition of the oil due to weathering and biodegradation (Sergy and
Blackall 1987).

Cleanup techniques evaluated were based on the operational realities of the eastern Canadian Arctic (Lee
et al. 2011). Small labour force and the impracticability of disposing of large volumes of contaminated
materials were the primary limiting factors; therefore, the emphasis was placed on the selection of
techniques that would either have low labour or simple waste disposal requirements. The methods chosen
were: in situ combustion; mechanical mixing of contaminated sediments; application of chemical
surfactants to disperse stranded oil (BP1100X and Corexit 7664); and application of a solidifying agent to
the stranded oil.

Chemical solidification was effective in stabilizing the oil but very labour intensive. Low-pressure
flushing by sea water of oiled fine-grained beach sediments did not reduce oil concentrations. Burning
also proved ineffective as even high-temperature igniters failed to sustain combustion on oiled cobble
beaches. Application of commercial fertilizers to the plots increased bacterial numbers and degradation of
oil on fine-grained backshore sediments but not on those of coarse material (Eimhjellen and Josefson
1984). Based on the experimental design (i.e., pseudoreplication), methodologies available (e.g., lack of
genomic analysis to study whole microbial community responses) and the type/amount of data collected
that limited statistical rigour, Sergy and Blackall (1987) could only suggest that biodegradation, enhanced
or natural, was unlikely to be a factor of quantifiable significance except over a long period of time.

Wave energy was insufficient to mix surfactant and oil on very sheltered beaches; therefore, this
treatment was ineffective in these areas. The use of surfactants quickly reduced the amount of oil present
on the surface of cobble beaches partially exposed to waves. Sergy and Blackall (1987) concluded that
surfactants would be most effective on small sections of coast where stranded oil might otherwise have a
severe short-term impact or where long-term persistence was not desirable. Furthermore, consideration of
the effects of flushing the oil from the beach into adjacent near-shore waters would be required.

Mechanical mixing of oiled intertidal and backshore areas reduced the total hydrocarbon concentrations
in surface sediments, but in many cases at the expense of increasing subsurface concentrations (Sergy and

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 285
Blackall 1987). The technique could be applied where the objective is to reduce contamination of surface
traffic, to prevent or reverse asphalt pavement formation, or, in the backshore, to increase rates of
weathering and to enhance biodegradation.

8.2.3 Important factors affecting the consequences of the experimental BIOS spills

8.2.3.1 Oil properties and behaviour

 One-third of the near-shore undispersed oil spill evaporated.


 Wave action was the primary mechanism for removal of oil from beaches.
 The majority of the residual oil on the beaches was incorporated into asphalt pavements.

8.2.3.2 Effect of the environment on fate and behaviour

 Rising tides removed large quantities of oil on sheltered beaches over the first two days, after
which relatively stable conditions prevailed.
 Oil was less persistent in the fine-grained sediments than on the pebble-cobble beach, where oil
residues were still visible four years post-spill.
 Biodegradation was considered a minor factor, but the evidence was inconclusive, especially in
light of findings from recent studies using ‘genomic’ techniques.
 Oil persisted for long periods of time on low-energy beaches and backshore areas and in nearby
seabed sediments.
 Climate and soil conditions played major roles in the fate of oil in the backshore plots.

8.2.3.3 Oil toxicity

 Biological impacts from both near-shore spills (dispersed and undispersed) were relatively minor
and effects were short-term with no evidence of large-scale mortality of subtidal benthic biota and
few changes in community structure of infauna, epifauna and macroalgae.
 Exposure to the persistent oil residues in subtidal sediments (from both dispersed and udispersed
near-shore spills) was responsible for medium-term (1-2 year) sublethal effects, but only in a few
species.
 The subsurface injection of dispersed oil caused benthic organisms to be exposed to high
concentrations of toxic aromatic hydrocarbons that would otherwise have been rapidly lost to
surface evaporation; responses included the emergence from the sediment and/or immobilization
of infaunal and epibenthic invertebrates. Recovery and reburial occurred in the two weeks
following the spills.

8.2.3.4 Spill response

 Skimming removed approximately one-third of the undispersed surface oil spill.


 Chemical solidification was effective in stabilizing the oil but very labour intensive.
 Low-pressure flushing by seawater of oiled fine-grained beach sediments did not reduce oil
concentrations.
 Burning was ineffective as even high-temperature igniters failed to sustain combustion on oiled
cobble beaches.
 Application of commercial fertilizers to the plots increased bacterial numbers and degradation of
oil on fine-grained backshore sediments but not on those of coarse material.
 Wave energy was insufficient to mix surfactant and oil on very sheltered beaches; therefore, this
treatment was ineffective in these areas.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 286
 The use of surfactants quickly reduced the amount of oil present on the surface of cobble beaches
partially exposed to waves.
 Mechanical mixing of oiled intertidal and backshore areas reduced the total hydrocarbon
concentrations in surface sediments, but in many cases at the expense of increasing subsurface
concentrations.

8.2.4 Lessons learned from the BIOS project

 According to Sergy and Blackall (1987), the BIOS Project provided no major ecological reasons
to prohibit the use of chemical dispersants on oil slicks in near-shore areas similar to the
experimental sites. Despite the high exposure of benthic organisms to chemically-dispersed oil,
there were no major population- or community-level consequences. However, for use in areas
where benthic organisms (such as filter-feeding bivalves) are consumed by wildlife or humans,
the authors raised concerns over bioaccumulation.
 Sergy and Blackall (1987) stated that chemical dispersion may be the only alternative in
situations where the immediate protection of shoreline and near-shore habitats is of primary
importance or where a shoreline cleanup operation is environmentally less desirable. They further
suggested that where practical and effective application methods and dispersant formulations are
available, it would seem appropriate to give pre-spill approval for dispersant use along sections of
Arctic coastline with ecosystems typified by the Cape Hatt site.
 Exposure to low-level oil residues in the sediments did not cause significant changes in the
subtidal benthic populations over a two-year period following the spill.
 Oil residues persisted for long periods of time on low-energy beaches and backshore areas and in
nearby seabed sediments; therefore, cleanup decisions should consider the implications of long-
term oil residues to shoreline users and the potential for chronic bioaccumulation or sublethal
effects in subtidal benthos.
 The results confirmed that Arctic beaches can be cleaned of oil by natural processes, despite the
short open water season, and that this can occur very quickly on beaches exposed to moderate
wave action.
 Sergy and Blackall (1987) stated that the BIOS findings show that natural cleaning of oil-
contaminated beaches can be an environmentally acceptable option for low-priority shorelines
similar to those at Cape Hatt. Response efforts would be more effectively directed toward areas of
greater importance and sensitivity, such as shores adjacent to communities, wildlife breeding and
staging areas and traditional hunting and fishing camp locations.
 Of the cleanup methods studied, two—surfactant flushing and mechanical mixing—produced an
immediate reduction in the quantity of oil on the beach surface. These might be desirable where
the objective is to reduce contact between oil and wildlife that frequent the shoreline.
 Washing with surfactant could also be considered as a means to prevent oil-sediment
consolidation (asphalt pavement) in the intertidal zone.

8.2.5 Research needs based on the BIOS project

Recommendation: The fate and behaviour of oil (dispersed and undispersed) released due to a
subsea blowout in the Arctic requires much more study. Models to predict the trajectory of
dispersed oil in open water and under ice are needed. The models will require real-time high
resolution metadata of oceanographic properties to provide emergency response support (Lee et al.
2011).

Recommendation: The relative risks of bioaccumulation across trophic levels in Arctic subtidal
communities should be assessed and the results of the assessment used to set research priorities.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 287
Risks should be estimated for organisms which consume subtidal benthic species, as well as the
benthic species themselves.

Recommendation: The comparative risks of undispersed and dispersed oil spills under Arctic
conditions to fish, seabirds and marine mammals should be assessed. The risk scenarios should
include subsea use of dispersants.

Recommendation: Future experimental spill in any environment must use statistically rigorous
procedures involving randomized and replicated plots.

8.3 Other Arctic Oil Spill Field Trials

8.3.1 Review of conclusions arising from other experimental spills in ice covered waters

Dickins (2011) reviewed research into the behaviour of oil spills in ice covered waters. The Dickins
review is comprehensive and provides a base upon which to develop research programs in the Canadian
Arctic. At the time of the review, Dickins noted that the SINTEF Oil in Ice Joint Industry Partnership
Program conducted between 2006 and 2009 in the Norwegian Barents Sea was the most comprehensive
Arctic oil spill research program completed. The field trials component of this program was between 2008
and 2009. The SINTEF projects have investigated: oil distribution and bioavailability; fate and behaviour
of oil spills in ice; in situ burning (ISB) of oil spills in ice; mechanical recovery of oil spills in ice; use of
dispersants; and chemical herders on oil spills in ice. Field trials have demonstrated that oil spilled in ice
conditions was significantly thicker than in open water, and the final area of spilled oil was much wider in
open water (Lee et al. 2011). Thus, oil contained in ice and snow would be thicker and more easily burned
or otherwise recovered. Furthermore, containment in ice reduces wave action, and relatively slow
weathering in the Arctic environment can provide time for spill response, extending the window of
opportunity for mobilizing response activities and enhancing the effectiveness of certain response
measures such as burning and skimming (Dickins 2011).

Lee et al (2011) provided a summary of conclusions drawn from experimental oil spills in the Arctic:

 Low water and air temperatures in addition to ice-covered waters generally result in greater oil
equilibrium thickness due to smaller contaminated areas and reduced spreading rates;
 Ignitability is enhanced in cold temperatures and ice due to the greater persistence of lighter and
more volatile components of petroleum hydrocarbons;
 Ice has a dampening effect on waves; therefore, the sea conditions in areas of the Arctic may be
significantly less severe than most open ocean areas, thus allowing for easier marine operations;
 Ice can naturally contain spilled oil and act as a barrier to spreading. The natural herding
properties of oil then enhance the effectiveness of ISB by thickening the slick;
 High ice concentrations (7/10 or more) tend to immobilize and encapsulate most spilled oil quite
rapidly, particularly from a subsea blowout;
 Ice encapsulated oil is effectively isolated from weathering and allows for a longer window of
opportunity to carry out cleanup activities. Effective combustion at a later date is possible
because the oil remains unweathered; and
 Most of the Arctic shoreline has a seasonal fringe of fast ice that acts as an effective natural
barrier against oil contamination on the coastline in winter.

Lee et al. (2011) also provided a summary of some of the challenges and limitations associated with
response to oil spills in the Arctic:

 Oil trapped on or under ice in moving pack ice is difficult to measure because crews cannot
maintain continuous operations with immediate means of evacuation;

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 288
 Skimming operations can be hindered due to the slow spreading and flow of oil in leads and
openings in pack ice;
 Moving vessels during response operations can cause rapid spreading of oil, creating a thinner,
less easily recoverable or ignitable slick; and
 Crude oils with pour points of 0 °C or less tend to gel.

8.3.2 Research needs regarding spills in Arctic conditions

According to Dickins (2011), future advances in the ability to respond to spills in ice will require a new
approach to permitting experimental spills in the Canadian Arctic. Lee et al. (2011) stated that Arctic field
trials could be used to address research questions, such as:

 Remote sensing systems to detect, monitor and map the transport and spreading of oil in ice and
below ice;
 Weathering of oil in cold water conditions;
 Development and verification of oil-in-ice drift and fate models;
 Improved mechanical response, such as skimmers, pumping systems for viscous oil and the
removal of oil from ice and water;
 ISB in broken ice;
 Oil dispersion enhancement by dispersants and enhanced OMA formation;
 Bioremediation of oil stranded on shorelines;
 Characterization of water-soluble components and biological effects on Arctic species; and
 Development of operational endpoints for spill cleanup operations.

Two major initiatives have been recently funded to address some of these oil spill response knowledge
gaps in the Arctic:

The Arctic Oil Spill Response Technology Joint Industry Programme (JIP) was initiated in 2012 and is
currently ongoing. It represents a collaboration of ten international oil and gas companies to enhance
industry knowledge and capabilities in the area of Arctic spill response, as well as to increase
understanding of potential impacts of oil on the Arctic marine environment under the management of the
International Association of Oil and Gas Producers. The JIP (described at:
www.arcticresponsetechnology.org) has several specific projects each focusing on a different key area of
oil spill response:

 Project 1 - Fate of Dispersed Oil under Ice: The project will provide important information for
dispersants use in ice-covered marine environments and develop a tool to support contingency
planning.
 Project 2 - Dispersant Testing under Realistic Conditions: The project will define the operational
criteria for use of dispersant and mineral fines in Arctic marine waters with respect to oil type, oil
viscosity, ice cover (type and concentration), air temperatures and mixing energy (natural, water
jet and propeller wash). Another objective is to identify the regulatory requirements and
permitting process for dispersant and mineral fines use for each Arctic nation/region.
 Project 3 - Environmental Impacts from Arctic Oil Spills and Oil Spill Response Technologies:
The project will improve the knowledge base for using "Net Environmental Benefit Analysis"
(NEBA) for response decision-making and ultimately facilitate stakeholder acceptance of the role
of EIA in oil spill response plans and operations.
 Project 4 - Oil Spill Trajectory Modeling in Ice: The project will advance the oil spill modeling
for oil spills in ice-affected waters by evaluating ice trajectory modeling approaches and
integrating the results into established industry oil spill trajectory models.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 289
 Project 5 - Oil Spill Detection and Mapping in Low Visibility and Ice: The project will expand
remote sensing and monitoring capabilities in darkness and low visibility, in pack ice and under
ice. This project is split into two elements: surface remote sensing (i.e. satellite-borne, airborne,
ship-borne and on-ice detection technologies) and subsea remote sensing (i.e. mobile-ROV or
AUV based and fixed detection technologies).
 Project 6 - Mechanical Recovery of Oil in Ice: The project will evaluate novel ideas for
improving efficiency of mechanical recovery equipment in Arctic conditions.
 Project 7 - In Situ Burning of Oil in Ice-Affected Waters. State of Knowledge: The project aims
to raise the awareness of industry, regulators and external stakeholders to the significant body of
knowledge that currently exists ISB.
 Project 8 - Aerial Ignition Systems for In Situ Burning: The project will develop improved
ignition systems to facilitate the use of ISB in offshore Arctic environments, including ice when
the presence of sea ice restricts use of vessels as a platform for this response option.
 Project 9 - Chemical Herders and In Situ Burning: The project will advance the knowledge of
chemical herder fate, effects and performance to expand the operational utility of ISB in open
water and in ice-affected waters.
 Project 10 – Field Research: Results from previous research projects show that many of the
advances in our state of knowledge about Arctic response technology were gained through
controlled field experiments with oil. This project will pursue opportunities for large scale field
releases for validation of response technologies and strategies.

Within Canada, the Churchill Marine Observatory was established in July 2015, with research
infrastructure funding is provided through the Canada Foundation for Innovation’s Innovation Fund,
Aboriginal Affairs and Northern Development, the Province of Manitoba and collaboration with the
universities of Calgary and Victoria. This facility will be a dedicated multi-disciplinary research facility
for study on the impact of oil spills in sea ice, as well as the investigation issues facing arctic marine
transportation.

Recommendation: The effects of climate change on ice cover and ice behaviour in the Canadian
Arctic is an important issue requiring investigation.

Recommendation: Explicit research questions regarding effects on marine biota require further
study and validation both for spills under ice-covered conditions and in open-water conditions.

Recommendation: Effects of cleanup activities on Arctic biota require study in order to support
comprehensive and reliable NEBA.

Recommendation: If future multi-disciplinary field trials are allowed in the Canadian Arctic, they
should be collaborative and incorporate the concerns and knowledge of Indigenous peoples.

8.4 The Exxon Valdez Oil Spill (EVOS)

On March 24, 1989, the Exxon Valdez altered its course from shipping lanes to avoid floating ice and
struck Bligh Reef in Alaska’s Prince William Sound. The grounding was caused by a series of human
errors committed by its crew members, including the master; however broader safety issues associated
with inadequate policies and procedures within the Exxon Shipping Company, as well as the U.S. Coast
Guard were identified by the NTSB (1990). Specific policy and procedure issues included: failure of the
Exxon Shipping Company to provide a fit master and a rested and sufficient crew; the lack of an effective
Vessel Traffic Service; inadequate personnel training; deficient management oversight; and the lack of
effective pilotage services (NTSB 1990).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 290
The remote location, accessible only by helicopter, plane or boat, made government and industry response
efforts difficult, as did severe weather with high winds, which came about two and half days later. About
42,000 tonnes of Prudhoe Bay crude oil (Alaska North Slope medium crude oil; Table 2.2) was released
into the ocean affecting an area of about 28,000 km2. The dispersant Corexit 9580 was applied on the day
of the spill, but there was not enough wave action to mix the dispersant with the oil in the water (Chapter
6). Some of the surface oil was burned, reducing 113 m3 of surface oil to a removable residue; however,
unfavourable weather prevented further burning. Booms and skimmers (Chapter 6) were deployed for
mechanical recovery, but skimmers were not readily available during the first 24 hours, and thick oil and
kelp tended to clog the equipment. A major storm 2.5 days after the spill led to extensive mousse
formation. About 20-30% of the spilled oil evaporated, was diluted or dispersed into the water column or
photooxidized (Chapter 2) (Wolfe et al. 1994). Within three weeks, less than 25% of the oil remained on
the sea surface.

Immediate effects of the spill included large-scale seabird mortality (estimates vary but average about
250,000 comprising over 90 species), as well as the deaths of about 1,000 to 5,500 sea otters (Platt and
Ford 1996; Ballachey et al. 2014). Twenty-two killer whales died. One pod of highly-exposed killer
whales lost seven members within a week of the spill, including three adult females, and an additional
seven or eight members of this pod died over the next two years (Spies et al. 1996). Other species affected
directly by mortality included river otters, harbour seals and bald eagles, plus unknown numbers of
herring, salmon and other fish species (Ballachey et al. 2014).

About half of the oil was distributed along the shoreline and inter- and subtidal areas as far as 970 km
from the spill site (Owens 1991; Wolfe et al. 1994). About 782 km of Prince William Sound (about 16%
of shoreline) and 1,315 km of Gulf of Alaska (about 14% of shoreline) were oiled to some degree (Owens
1991; Neff et al. 1995).

Shoreline cleanup included manual removal and the use of surfactants and high-pressure hot water
(Chapter 6). The cleanup focused on the intertidal zone, including removal of tar mats, mousse and
subsurface deposits, as well as sediment relocation and biostimulation to enhance microbial degradation.
An assessment in 1990 showed that most remaining deposits were isolated from the biological
environment (with the exception of exposure pathways leading from mussels to sea otter (Chapter 4);
therefore, only a few larger deposits of subsurface oil were relocated to the middle and upper intertidal
zones where it was subject to wave and tidal action and natural cleaning. Less intrusive cleanup efforts
continued through the summer of 1991 involving mostly manual removal and bioremediation. Although
biostimulation on an unprecedented scale (48,600 kg N in various nutrient formulations applied at 1,400
different sites) initially accelerated oil removal from shoreline surfaces (Atlas and Hazen 2011), after a
few years even the deliberately untreated (‘set-aside’ or control) sites were equivalent in terms of residual
oil load (Pritchard et al. 1992).

The aggressive cleanup methods had significant effects. For example, high-pressure washing (Figure 8.4)
removed all of the macroalgal and mussel communities in some areas, extending the time to recovery of
intertidal communities (State of Alaska 1993). Furthermore, the high-pressure washing, while removing
oil from the upper and mid-intertidal zone where its effects were somewhat restricted to relatively tolerant
organisms, such as barnacles, rockweed and mussels, transported the remobilized oil into the lower
intertidal and shallow subtidal zones where the oil was placed into contact with relatively more sensitive
organisms, such as hard shelled clams and crustaceans (State of Alaska 1993). The physical features of
the shoreline were also affected; e.g., silty sediments were washed out of the beach areas and into the
water. Most of the animals that normally lived in these beach areas required a certain mix of fine-grained
sediments; therefore, many would not return until the beach sediments had stabilized (Shigenaka 2014).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 291
Figure 8.4 Cleanup workers spray oil-covered rocks on Prince William Sound with high-pressure hoses. (Photo
Courtesy of Exxon Valdez Spill Trustee Council)

Eight years after the spill, removal of oil was attempted on armoured portions of beaches in Prince
William Sound using high-pressure injection of a surfactant (Brodersen et al. 1997). The cleaning was
conducted at the request of the local subsistence-use community. Cleaning resulted in a mean reduction of
62% at treated areas with further reduction occurring in the following year. However, newly visibly oiled
sites were exposed by winter storms the year after the cleanup.

Shigenaka (2014) stated that despite the unprecedented scale, duration and cost of the response, modeling
of the fate of the spilled oil estimated that the cleanup itself removed only a small portion (a little more
than 10%) of the spilled oil from the environment. By far, the largest part of the total was naturally
weathered or degraded (Figure 8.5).

Figure 8.5. Modeled fate of the spilled Exxon Valdez oil with time. The portion recovered by the cleanup is in
green (after 1,000 days). Reprinted with permission from Wolfe, D.A., Hameedi, M.J., Galt, J.A., Watabaytashi, G.,
Short, J., O’Clair, C., Rice, S., Michel, J., Payne, J.R., Braddock, J., Hanna, S. and D. Sale. 1994. The fate of the oil
spilled from the Exxon Valdez. Environmental Science and Technology 28: 561–568. Copyright (1994) American
Chemical Society.

The Exxon Valdez Oil Spill Trustee Council (EVOSTC) has adopted an official list of resources and
services injured by the spill as part of its Restoration Plan. This list includes fish and wildlife resources
that “experienced population-level or chronic injury from the spill” (EVOSTC 2014). The list is divided

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 292
into six categories (Table 8.1). The EVOSTC Restoration Plan was first adopted in 1994 when it was
created as guidance for the expenditure of public funds.
Table 8.1 Exxon Valdez Oil Spill Trustee Council List of Resources and Services injured by the Spill
Recovering Recovered Not Recovering Recovery Human Services
Unknown
 Designated  Archaeological  Marbled  Kittitz’s  Commercial
wilderness resources murrelets murrelets fishing
areas  Bald eagles  Pacific  Passive use
 Intertidal  Barrow’s herring  Recreation and
communities goldeneyes  Pigeon tourism
 Killer whales –  Clams Guillemots  Subsistence
AB Pod  Common loons  Killer
 Sediments  Common Murres whales –
 Cormorants AT1
 Dolly Varden population
 Harbour seals
 Harlequin ducks
 Mussels
 Ink salmon
 River otter
 Sea otters
 Sockeye salmon

Some authors maintained that the persistent, residual oil in unconsolidated sediments of the intertidal
zone continues to contaminate invertebrates to the extent that some vertebrate consumers (including fish,
otters and seabirds) are still being exposed to toxic oil concentrations (Peterson et al. 2003; Ballachey et
al. 2014). In contrast, other authors have concluded that the Prince William Sound ecosystem has
effectively recovered from the EVOS (Harwell and Gentile 2006).

There appears to be a consensus that continued contamination of subtidal areas is not a concern
(Ballachey et al. 2014). The authors noted that some of the estimated 55,000 kg of oil remaining in mid-
and upper-intertidal habitats may persist at some sites for several decades and that some of the residual oil
is largely unweathered because it exists in anaerobic patches where biodegradation is slow (Figure 2.3;
Chapter 3). This estimate is for subsurface oil in the upper two-thirds of the intertidal zone. Inclusion of
the lower one-third and surface oil would substantially increase this estimate (Short et al. 2004).

Potential chronic effects to the pink salmon population have been linked to direct embryo mortality in the
first five years after the spill and continuing exposure to residual oil. Ballachey et al. (2014) summarized
the results of several field and laboratory studies of pink salmon, and concluded that these studies
“provided compelling evidence of chronic impacts to pink salmon from oil persisting in intertidal
habitats”. They cited evidence including: 1) observations of elevated embryo mortality in the five years
following the spill; 2) identification of the exposure mechanism from contaminated beaches to spawning
gravels; and 3) measured effects on fitness following embryonic exposures under laboratory conditions.
They concluded that the group of pink salmon studies they reviewed provided “unprecedented evidence
that exposure of embryonic life stage (sic) to low level PAHs (in ppb) from persistent oil can have a
population level effect”. Ballachey et al. (2014) also concluded that chronic exposure to oil and possibly
latent effects of acute exposure appeared to have decreased survival and constrained recovery of the sea
otter population in Prince William Sound for more than two decades.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 293
Not all authors agree that population-level effects on pink salmon persisted. In a detailed review of
studies of effects on pink salmon in Prince William Sound, Brannon et al. (2012) concluded that problems
with experimental design and methods called into question the potential for population-level effects.
Brannon et al. (2012) stated that there was no evidence that weathered oil increased toxicity, either in the
form of total PAH or as HPAH (high molecular weight PAH) leachate. They went on to state that field
evidence suggested that “well-weathered oil deposits in the actual areas impacted by the EVOS do not
represent a continuing and increasing threat of interstitial toxic water to pink salmon eggs incubating in
adjacent streams”. Further, the authors examined studies of growth rate of juvenile pink salmon in Prince
William Sound and concluded that the assumptions necessary to discern growth effects either were not
met or contained uncertainties about their validity. As discussed in Chapter 4, the tributaries to Prince
William Sound produce so many juvenile fish that any deficit in production in tributaries affected by the
spill did not appear to affect total abundance.

The effect of spatial scale may underlie much of the debate regarding effects on pink salmon. Hatchery
production of pink salmon in Prince William Sound is comparable in numbers with wild stock
production. Hatchery production was not affected by the EVOS, and effects on wild stock production
were negligible at the scale of Prince William Sound as a whole. However, pink salmon spawning streams
that bisected beaches that were heavily oiled by the EVOS were subject to mortality and sublethal effects
on embryos, as stated by Ballachey et al. (2014).

Herring have been the focus of extensive studies since the EVOS because they were commercially
harvested in Prince William Sound and their numbers showed large declines within a few years of the
spill. The herring population was still very low in 2014 with no commercial fishery present; however, the
role of the spill in the decline and lack of recovery of herring, relative to other factors, including disease,
predation and recruitment, is unclear and, according to Ballachey et al. (2014), is unlikely that it will ever
be completely understood.

Peterson et al. (2003) emphasized that there can be delayed population reductions and cascades of indirect
effects that postpone recovery. For example, harlequin ducks appeared to be affected via the energetic
costs of metabolizing PAHs, leading to lower body mass and elevated overwintering mortality. According
to a study by Iverson and Esler (2010 cited by Ballachey et al. 2014), full recovery of the harlequin duck
population would require from 16 - 32 years under best-case and worst-case scenarios, respectively.
Peterson et al. (2003) pointed out that there were
also cascade effects from the cleanup actions. For
“If the Exxon Valdez experience has taught example, the removal of Fucus stands with their
us anything, it has emphasized the associated community of grazers, such as limpets
importance of variability as both a key and periwinkles, led to initial blooms of green algae
feature of biological communities and a and opportunistic barnacle growth with declines in
critical consideration to integrate into invertebrates associated with the Fucus canopy.
assessment of disturbance and recovery. As After regrowth of the Fucus, there was another mass
we inevitably consider oil spill scenarios for mortality in 1994, probably caused by simultaneous
the Arctic, they are framed against the senility of a single-aged stand. All of this extended
background of change that is occurring at the recovery process for a decade or more.
unprecedented rates” (Shigenaka 2014).
The Shigenaka (2014) analysis concluded that the
intertidal biota of Prince William Sound had
recovered, at least in terms of the metrics used in the analysis (attainment of parallel temporal trends in
abundance), after an initial six-year period. This conclusion applied to infauna, algae and invertebrates
living on the surface of rocky substrates. The authors emphasized that their conclusions were for intertidal
populations only.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 294
The Exxon Valdez Oil Spill Trustee Council (2015), Harwell and Gentile (2006) and Shigenaka (2015) all
note that while the spill and cleanup clearly caused significant ecological effects for months to a few
years post-spill, natural variability and the occurrence of multiple anthropogenic stressors not associated
with the spill are now making the detection of any potential residual effects of the spill very difficult. In
addition, current natural and human stressors may be hindering recovery of some resources initially
injured by the spill. The passage of time and the evolution of science have shifted the purpose and utility
of the Injured Resources and Species List. “The Council recognizes that the complexities and the
difficulties in measuring the continuing impacts from the spill result in some inherent uncertainty in
defining the status of a resource or service through a specific list and the Council’s focus has accordingly
expanded to a more ecosystem approach” (EVOSTC 2015).

The inherent natural variability of biological communities complicated interpretation of monitoring data
following the spill (Shigenaka 2014). A new statistical approach to analysis was designed, which
acknowledged that: unoiled reference sites may be biologically different from oiled sites at the time of
impact, thus rendering absolute convergence of conditions in the recovery phase less useful; and
conditions at both oiled and unoiled locations are likely influenced by a host of factors not related to the
spill disturbance (Shigenaka 2014). The new approach analysed long-term datasets for patterns of
abundance. The underlying assumption was that in the absence of spill effects, sites would respond
similarly to climate, ocean conditions or other determinants of biological communities in a defined study
region.

The risk of oil spills must be assessed against the backdrop of responses to large-scale phenomena. The
Shigenaka (2014) analysis noted that results across three independent experimental studies revealed a
strong correlation of biological metrics in Prince William Sound with cycles of the Pacific Decadal
Oscillation (PDO). For example, mussels and molluscs appear to respond positively to warmer cycles of
the PDO whereas rockweed shows greater abundance during cool phases. The author stated that in recent
years, linkages between changes in Alaskan biological communities and conditions and large-scale
atmospheric, oceanic and climatic shifts have grown more numerous.

The use or interpretation of data in support of specific values or policies is called ‘normative science’,
and, according to Landis (2007), normative science can explain at least some of the differences in
interpretation of data related to long-term effects of the EVOS. According to Landis, ecosystem health,
ecosystem integrity, ecological significance and recovery are constructs that incorporate values and
policies. “Separation of science from policy or at a minimum a transparent acknowledgement of the
science-policy interaction is clearly necessary to obtain a clear picture of the ecological system under
investigation” (Landis 2007). Landis suggested two alternatives for dealing with normative science: 1) to
stop using terms such as ecological significance, integrity and recovery (in the Clementsian context)3; or
2) to understand that ecological policy is a complex and multi-component decision-making process that
cannot be summarized in metaphorical clichés, such as ecosystem health. If a policy goal is being
discussed or supported, it should be clearly defined and obvious to the reader. Landis (2007) asked
whether controversies such as those surrounding long-term effects of the EVOS are really debates about
policy rather than scientific merit.

8.4.1 Important factors affecting the consequences of the Exxon Valdez spill

8.4.1.1 Oil properties and behaviour

 Evaporation occurred at lower rates due to low temperature.


 Natural dispersion was wind driven and very widespread.

3
Clementsian ecology is based on assumptions of steady-state equilibrium and development of predictable and specific climax
communities, which have fallen out of favour given the complex and dynamic properties of ecosystems.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 295
 Stranding was widespread on shorelines.
 Remobilization of oil deposited on shorelines occurred by tidal and wave action.

8.4.1.2 Effect of the environment on fate and behaviour

 Biodegradation was slow but significant at the ambient temperatures present in Prince William
Sound.
 Oil persisted in unconsolidated sediments of the intertidal zone.
 Some persistent oil remained in a relatively unweathered state due to low oxygen conditions in
sediments.
 Wave exposure enhanced the weathering of non-sequestered residual oil.

8.4.1.3 Oil toxicity

 Acute toxic effects of floating oil from the spill included large-scale seabird and sea otter
mortality.
 Chronic toxicity caused mortality-related damage to the social structure of a killer whale pod.
Indirect effects may have postponed recovery of some species (e.g., energetic costs of
metabolizing PAHs caused lower body weight and higher winter mortalities in harlequin ducks).
Long-term effects of residual oil on pink salmon were caused by exposure of early life-stages and
may have been limited to sub-populations using spawning streams that bisected beaches that were
heavily oiled.
 Intertidal biological communities recovered relatively rapidly from the combination of oil toxicity
and the effects of cleanup activities, although there was a mass mortality of Fucus in 1994
(associated with simultaneous senility of a single-aged stand), which extended recovery of areas
affected by the mortality by a decade or more.

8.4.1.4 Spill response

 Limited regional response capability and the remoteness of the location with no ground access
delayed the response.
 Severe weather following the spill limited some response options, such as ISB and use of
dispersants, so that the spread of the surface oil was increased and more oil was entrained into
intertidal sediments on affected shorelines.
 Aggressive shoreline cleanup methods, such as high temperature and pressure washing, impaired
site recovery rates due to mortality and removal of indigenous species and habitat destruction.

8.4.2 Lessons learned from the Exxon Valdez spill

 Challenges inherent in oil spill response in northern, remote areas prone to adverse weather
increase the probability that oil released from near-shore tanker accidents will rapidly disperse
and reach shorelines over a wide area.
 Some response measures can be difficult or impossible to implement because of weather and sea
conditions (e.g., the use of dispersants or ISB).
 Aggressive shoreline cleanup methods, such as high-pressure washing, can impart substantial
negative effects on biological communities, including indirect and cascade effects, and can delay
recovery.
 Cleanup removed a small proportion of the spilled oil (about 10%).
 Weathering and biodegradation were of primary importance with respect to the nature and extent
of residual oil.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 296
 Altering the physical features of a beach or shoreline can significantly affect the recovery of
impacted plants or animals. Physical recovery and stabilization of a site are necessary for
biological recovery.
 Long-term, population-level effects are difficult to study and the results can be controversial,
which emphasizes the critical need for baseline data on natural resources, as well as consensus-
based definitions of monitoring endpoints.
 The inherently high degree of natural variability found in systems such as Prince William Sound
can limit or preclude the use of standard or traditional statistical methods. Risks of oil spills in the
Arctic must be framed against the background of large-scale oceanic and climatic phenomena and
changes that are occurring at unprecedented rates (Shigenaka 2014).
 NOAA found that so-called ’set-aside sites’ that were oiled but intentionally left uncleaned have
been critical to the NOAA monitoring program's ability to determine impacts due to oiling alone
and those due to cleanup, and to enable scientifically rigorous interpretation of data.
 Normative science may have played a role in the debate regarding long-term, population-level
effects of the spill. This is not surprising given the high social, economic and cultural value of the
receiving environment. The consequences of being wrong about the interpretation of data can be
ecological and societal. Therefore, both consequences should be identified and considered in a
transparent manner.

8.4.3 Research needs related to the Exxon Valdez spill, in the Canadian context

Recommendation: The most commonly-used suites of biological metrics for Canadian receiving
environments (e.g. intertidal or subtidal community metrics) should be examined across broad
gradients of natural factors and over longer time-frames to establish the relative roles of local,
regional and global factors. Research should be conducted on a range of species and/or
communities with various levels of resilience and should focus on identified high risk marine
environments (note: WSP and SL Ross, 2014a,b).

The relative role of local, regional and global factors in determining the natural variability of biological
metrics is poorly understood, particularly at the population and community level. Factors, such as climate
and oceanographic variability, overexploitation and invasive species, may overwhelm any signal of
subtle, long-term effects of residual oils. However, the incremental stress produced by exposure to oil
may contribute to ’tipping points’ for species whose resilience has already been challenged by other
stressors. Furthermore, future field studies should incorporate multiple study areas (both reference and
treated) to enable determination of experimental error, which might mitigate the effect of natural
variability affecting scientific conclusions.

8.5 Deepwater Horizon (DWH) Blowout

The DWH offshore oil drilling rig was situated in the Macondo oil prospect in the Mississippi Canyon, a
valley in the continental shelf of the Gulf of Mexico. The well over which it was positioned was 1,522 m
below the surface and extended 5,486 m into the geological formation under the seabed. On the night of
April 20, 2010, a surge of natural gas blasted through a concrete core recently installed to seal the well.
The natural gas travelled up the rig’s riser to the platform where it ignited, killing 11 workers and injuring
17. The rig capsized and sank on the morning of April 22, rupturing the riser, through which drilling mud
had been injected to counteract the upward pressure of oil and natural gas. Without any opposing force,
Macondo light crude oil (Table 2.2) began to discharge into the Gulf (Britannica Online 2015).

Several unsuccessful attempts were made to slow or stop the flow of oil, a technically challenging task
given the extreme depth of the blowout. A cap was successfully placed in June that greatly reduced the
flow, but didn’t eliminate it. The well was permanently sealed with the successful completion of a relief
well on September 19, 2010. The spill is considered the largest accidental marine oil spill in the history of

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 297
the petroleum industry. The estimated volume released varied from about 430,000 to 500,000 tonnes
(McNutt et al. 2012; Fingas 2013; Chapter 6). According to satellite images, the spill directly affected
180,000 km2 of ocean (Figure 8.6). Response measures included over 1,300 km of containment booms to
either corral the surface oil or function as barriers to protect coastal areas, shrimp/crab/oyster ranches or
other ecologically sensitive areas. About 7,000 m3 of two different Corexit chemical dispersants were
used, including about 2,920 m3 injected into the subsurface oil plume at the wellhead during the spill
(Chapter 6). Oil was removed from the water surface via ISB, in addition to physical recovery (e.g.,
booming and skimming).

Figure 8.6 Oil from the Deepwater Horizon oil spill approaches the coast of Mobile, Alabama, May 6, 2010.
Image from US Navy.

According to the Federal Interagency Solutions Group (2010), the expected mass balance was: 17% of the
oil was collected as it was released from the well; 5% was burned; and 3% was skimmed from the surface
for a total of 25% (Figure 8.7). The fate of the 75% of the oil not collected, burned or skimmed was
estimated as follows:

 13% dispersed naturally;


 16% dispersed with Corexit;
 23% evaporated or dissolved; and
 23% other mechanisms such as stranded on shorelines or sinking to the seafloor.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 298
Figure 8.7. Oil budget estimates expressed as percentages of the cumulative volume of oil discharged through
July 14, 2010. These estimates served as a guide for the response to the Deepwater Horizon blowout. Image from
Federal Interagency Solutions Group (2010).

A more recent review of the mass balance concluded that of the total amount of oil released, about 6 to
7% of the oil was burned in situ, about 12 to 15% of the oil was skimmed, and about 6 to 26% (best
estimate about 10%) of the oil arrived onshore (Fingas 2013). Fingas estimated that 50 to 55% of the oil
remained in the water column.

Despite the magnitude of the DWH spill, the rate and extent of natural oil dispersion and weathering
processes was highly significant. Following the release of the oil and its exposure to various
countermeasure treatments (including dispersant applications) the presence of oil components was
confirmed by detection in separate subsurface intrusion layers at depths ranging from 800 to 1400 m and
surface waters (Camilli, et al. 2010; Diercks et al. 2010; Hazen et al. 2010; Valentine et al. 2010;
Kujawinski et al. 2011; OSAT 2011; Socolosfsky et al. 2011; Spier et al. 2013). In a comprehensive study
on the spatial and temporal distributions of hydrocarbon data within the Gulf of Mexico, Wade et al.
(2015) compared pre-spill background concentrations of total petroleum hydrocarbons (TPH) and PAH in
waters samples with samples (over 20,000) collected during and after the DWH incident (13,000
stations). Samples in the database were collected by multiple response agencies, trustees and BP and
reported in the Gulf Science Data (file W-01v02-01.csv available at gulfsciencedata.bp.com) which had
undergone an extensive quality assurance/quality control validation process prior to release. Locations of
the sample collection ranged from a few meters to over 800 km in all directions from the wellhead.
During the incident, samples with the highest concentrations of hydrocarbons were collected within 25
km of the wellhead or in samples collected from surface slicks and dispersant use. Of the 13,172 water
sample TPH concentrations reported, 84% were below 1 μg/L (background). Of the 16,557 water sample
PAH concentrations reported, 79% were below 0.056 μg/L (the median field blank, background). The
percentage of samples below background increased rapidly after the well was capped.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 299
Subsequent results have shown that sinking played a more important role than previously thought.
Valentine et al. (2014) suggest that about 318,000 m3 of submerged oil from the subsurface plume may
have been trapped in deep ocean intrusion layers at depths of about 1,000 - 1,300 m. Based on spatial,
chemical, oceanographic and mass balance considerations, they calculated that between 4 - 31% of oil
sequestered in the deep ocean was deposited in patches to deep sea sediments. The pattern of
contamination points to deep ocean intrusion layers as the source and was most consistent with dual
modes of deposition: 1) a ‘bathtub ring’ formed from an oil-rich layer of water impinging laterally upon
the continental slope (at a depth of ∼900 - 1,300 m); and 2) a higher-flux ‘fallout plume’ where
suspended oil particles sank to underlying sediment (at a depth of ∼1,300 - 1,700 m). Independently,
Chanton et al. (2015) also calculated that a large proportion of the ‘missing’ Macondo oil has been
partially buried in deep sea sediments in the Gulf of Mexico. The precision of these estimates is still
debated due to sample heterogeneity. Furthermore, intermittent natural oil seeps, with a fingerprint similar
to the Maconda oil, also contributed to uncertainty within the oil budget. In terms of shoreline impact
assessments, it was difficult to distinguish tar mats or tar balls originating from the spill from those
originating from seeps.

8.5.1 Dispersant use

During the spill response, approximately 6.9 million litres of Corexit were applied, 4 million litres of
Corexit 9527 and 9500A at the surface and 2.9 million litres of Corexit 9500 via subsurface injection
(Federal Interagency Solutions Group 2010).

One of the controversial aspects arising from the DWH oil spill response operation was the injection of
dispersants at the wellhead located at 1,500 m depth. Monitoring programs identified the presence of a
plume of oil at a depth between 1,100 and 1,200 m below the surface (Camilli et al. 2010; Reddy et al.
2011). While the plume was initially thought to be direct evidence of dispersant effectiveness, a number
of studies suggested that the subsurface dispersants had little impact in dispersing the spill or preventing
oil from reaching the surface (e.g., Paris et al. 2012; Peterson et al. 2012). They argued that the turbulence
in the jet of gas and oil from the well head was sufficient to induce massive dispersion. Both processes
likely occurred. In addition to changes in oil droplet size (Johansen et al. 2013; Brandvik et al. 2013),
evidence supporting the benefits of the subsurface dispersant application included aerial photographs
taken during the DWH oil spill that showed a loss and gain in the magnitude of the surface slick
following subsurface dispersant injection and its shutdown.

The trade-offs regarding the use of dispersants are the subject of debate since the dispersants may also
enhance the bioavailability of the spilled oil or be toxic on their own. Furthermore, laboratory studies
have recently suggested that dispersant additions may initially suppress the activity of natural oil-
degrading microorganisms by altering microbial community composition towards dispersant-degrading
Colwellia sp. (Kleindienst et al. 2015). However, there is also strong evidence that the formation of small
oil droplets with a larger oil:water interface and higher dissolution rates facilitated by dispersant
applications would favour enhanced biodegradation rates. In a recent review of the pros and cons of
dispersants, Prince (2015) concluded that in most cases, the potential environmental costs of dispersant
use are likely outweighed by the much shorter residence time of dispersed oil in the environment.

The fate and effects of the dispersants with associated dispersed oil within the water column of the Gulf
of Mexico is the subject of considerable uncertainty. Kujawinski et al. (2011) found that the concentration
of dioctyl-sodium sulfosuccinate or DOSS (a key ingredient of these dispersants) was sequestered in
deepwater hydrocarbon plumes at 1,000 - 1,200 m water depth and did not intermingle with surface
dispersant applications. They also found that the concentration distribution was consistent with
conservative transport and dilution at depth and it persisted up to 300 km from the well, 64 days after
deepwater dispersant applications ceased. Thus, the surfactant does not appear to have been rapidly

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 300
biodegraded. This finding agrees well with the laboratory results reported by Campo et al. (2013) , who
showed that at 5 °C DOSS degraded slowly by only 61% in triplicate microcosms by day 42.

While Corexit 9500 and Corexit 9527 are ranked within the low to moderate toxicity level for aquatic
species (George-Ares and Clark 2000), a level of sensitivity remains over their application. Evidence
from studies with invertebrates and fish have shown that oil dispersed as small droplets in the water
column was more bioavailable, and therefore more toxic, than the oil alone (Chapter 4). However, while
the results of laboratory studies have raised much media attention over the high risks of dispersant use,
ecological relevance must be considered. Following its dilution in an ‘open sea’ environment, it is
unlikely that the dispersants would be present in concentrations toxic to pelagic organisms (Kujawinski et
al. 2011; Lee et al. 2013). Nevertheless, caution must still be taken as most toxicological studies have
been based on short exposure periods (e.g. 96 h LC50 tests with standard ‘regulatory’ reference test
organisms). Many of the long-term Natural Resource Damage Assessment (NRDA) studies on the long-
term effects of oil and chemically-dispersed oil on trophic level dynamics in the Gulf of Mexico have not
yet been released, and no community/population level studies have been conducted on chronic effects of
the chemically-dispersed oil on deepwater organisms.

Microbial communities in the Gulf of Mexico rapidly responded to the spill, including in deepwater
plumes (Atlas and Hazen 2011). Redmond and Valentine (2012) showed that the deepwater microbial
community was very different from the surface microbial community, reflecting the colder temperatures
at the plume depth, as well as exposure to natural oil seeps in the Gulf and possible leakage from other oil
wells in the area. The authors suggested that the high natural gas content of the spill may have provided
an advantage to specific taxa, notably Colwellia and to a lesser extent Oceanospirillales (Chapter 3).

8.5.2 Deposition and effects on shorelines

A review by Michel et al. (2013) of the extent and degree of shoreline oiling stated that about 1,770 km of
shoreline was affected consisting of 51% beaches, 45% marshes and 4% other shoreline types.
Simulations of oil transport from the footprint of the spill on the water surface showed that the mass of oil
that reached the shorelines was between 9,000 and 27,200 tonnes, with an expected value of 19,960
tonnes (Boufadel et al. 2014). It should be noted that this predicted mass is much lower than estimated by
Fingas (2013). Shoreline cleanup was authorized on 550 km of shoreline. Two years after the spill, oil
remained on 687 km of the shoreline but to a much lower degree (e.g., the authors reported that the
heavily oiled category declined from 360 km to 6.4 km using the Shoreline Cleanup and Assessment
Technique (SCAT)). The bulk of the oil stranded during a three-month period when many of the beaches
were in an erosional state that led to burial of the oil. In addition, oil was stranded high in the supratidal
zone due to high water levels and wave activity. The oil was buried, exposed and remobilized multiple
times in some areas. Removal of deeply buried oil required extensive mechanical and manual excavation
and sieving. In the lowest intertidal/near-shore subtidal zones, some of the oil/sand mixture accumulated
in the near-shore subtidal zone forming extensive submerged oil residue mats. Elsewhere, the oil/sand
residues adhered to relict4 marsh platforms composed of clay and peat at the toe of sand beaches—these
mats that were exposed only during the lowest of tides and/or buried by beach accretion were difficult to
remove. In marshes, the oil tended to strand along the marsh edge and spread no more than about 10 - 15
m inland. Some of the most heavily oiled marshes were cleaned using: intensive manual and mechanical
raking and cutting to remove the oiled vegetation mats and wrack (Chapter 6); careful removal or
reduction of the thick oil layers on the substrate; and limited application of loose organic sorbents.

Microbial communities on oiled shorelines showed a distinct response to the contamination, both in terms
of overall bacterial numbers and on the abundance of known oil degraders. Kostka et al. (2011) showed

4
A group of animals, plants or objects that exists as a remnant of a formerly widely distributed group in an environment different
from that in which it originated.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 301
that bacteria in Pensacola Beach sands were on average two to four orders of magnitude more abundant in
the presence of oil contamination, and high cultivatable counts as well as nucleic acid-based analyses
supported the premise that the majority of bacteria in the oiled sands were active. The authors noted that
the native microbial communities responded fairly quickly to oil contamination and suggested that
conditions in subtropical sands (temperature, oxygen supply and nutrients) appear to favour a broader
diversity of hydrocarbon degraders that might render biostimulation (via nutrient addition) unnecessary.

The impacts of the oil spill on shoreline vegetation were variable. Along some heavily oiled shorelines,
there was nearly complete flora mortality. Moderate oiling had no significant effect on some species (e.g.
Spartina; cord-grass) but significantly lowered live above-ground biomass and stem density of others
(Juncus; rushes) (Mendelssohn et al. 2012). Since the spill, some recovery has been noted (Mendelssohn
et al. 2012). However, there were concerns about whether some shorelines would revegetate naturally
before shoreline erosion occurred. Biber et al. (2014) found that there was more rapid removal or
degradation of oil along coastlines in Mississippi exposed to higher energy levels. Plant recovery was
more rapid in these locations. In contrast, at low-energy locations, oil was still detected in sediments and
on plants one year post-spill and plants in these locations exhibited chronic stress, which depressed
photosynthesis. Silliman et al. (2012) reported that while rapid salt marsh recovery was observed, there
were also permanent marsh area losses. They also observed thresholds of oil coverage that were
associated with severity of salt marsh damage. Plant death of marsh edges more than doubled rates of
shoreline erosion, further driving marsh platform loss that is likely to be permanent. However, in non-
eroded areas, marsh grasses had largely recovered. The authors noted that heavy oil coverage on
shorelines that were already experiencing elevated erosion because of intense human activities induced a
geomorphic feedback that amplified erosion and thus limited the recovery of otherwise resilient
vegetation.

8.5.3 Effects on fisheries

Near-shore fisheries were vulnerable to the spill both in the spawning grounds in the Gulf of Mexico and
in estuarine nursery areas. The spill overlapped with peak spawning periods for several important species,
including brown shrimp (Farfantepenaeus aztecus), white shrimp (Litopenaeus setiferus), blue crab
(Callinectes sapidus) and spotted seatrout (Cynoscion nebulosus) (Mendelssohn et al. 2012). Although the
location of the spill was in deep water, currents carried oil into the shallow spawning areas of these
species. The short- and long-term effects of this oil (and/or dispersants) on eggs and larvae are still
uncertain, but Fodrie and Heck (2011 cited by Mendelssohn et al. 2012) did not find short-term negative
effects on juvenile fish associated with inshore seagrass beds. Some evidence of low impact to fish
populations was noted at sites distant from the heavily oiled Louisiana coast (Anderson 2014). Although
coastal fishes likely have adapted for shifting habitat availability in Louisiana, it is uncertain whether
wetland losses due to the spill would have a negative effect on fish production (Anderson 2014). Fishery
responses to the spill are difficult to tease out because of high annual variability and the effects of many
other anthropogenic stressors (Mendelssohn et al. 2012). Anderson (2014) emphasized the importance of
continued monitoring of sediments, plants and animals in oiled areas.

A study of the effects of the spill on spotted seatrout by Brown-Peterson et al. (2014) showed a
substantial, but short-term effect on reproductive parameters after the spill compared with historical, pre-
spill data. The authors noted that the availability of pre-spill (baseline) reproductive data allowed direct
comparisons from the same sites before and after the spill. Important environmental variables, such as
temperature and salinity, were similar during most of the months of the reproductive season pre- and post-
spill, and the two significant differences in temperature and salinity pre- and post-spill were deemed to be
unimportant biologically. The authors acknowledged that other impacts on the spotted seatrout
populations in the years between the two sampling events, such as heavy fishing pressure, habitat loss and
changes in salinity due to drought and flood conditions, may also contribute to differences in reproductive
parameters, “although impacts from DWH seem to be the most parsimonious explanation for the observed

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 302
differences” (Brown-Peterson et al. 2014). Uncertainty regarding the interpretation of results is increased
because measured PAH concentrations in water taken from the sampling sites in 2011 were generally
below the 1.03 µg/L detection limit and were not significantly different from pre-spill concentrations. The
authors suggest that larval spotted seatrout may have been exposed to oil during the summer of 2010,
leading to altered reproductive dynamics in 2011 as the fish reached sexual maturity. Reproductive
parameters were returning to pre-spill levels by July, but the effective reproductive season was reduced
from the normal six months to three months.

Remedial operations may have caused significant secondary effects. The creation of a 35-km oil
prevention berm off the Chandeleur Islands, a unique habitat for neonatal and juvenile lemon sharks
(Negaprion brevirostris), may have reduced available habitat for the sharks and may also have inhibited
sharks pupped on the windward side of the island from reaching the protected sea grass beds on the
leeward side (McKenzie et al. 2014). A lack of suitable protected nursery habitat may have, in turn,
resulted in a further reduction of lemon shark numbers around the islands. The authors noted that it will
be important to continue monitoring the population numbers and the occurrence patterns to determine if
the effects are long-term.

8.5.4 Effects on invertebrates

Studies on the impacts of the DWH oil spill on blue crab (Fulford et al. 2014) and oysters (Le Peyre et al.
2014) revealed that effects, if any, were indirect (for blue crab) or the data indicated no significant
differences in biomarkers of exposure to PAHs (for oysters). Fulford et al. (2014) found that megalopal
(final crab larval stage) settlement patterns were more likely a reflection of climatic conditions and
pointed out that cyclic uncertainty in recruitment of decades or longer is important and should be
accounted for in interpreting data time series. They also found a low level of larval mortality due to PAH
exposure in 2010 in the wild, which was supported by laboratory assays that showed a mortality effect
only at PAH concentrations of 1 mg/L or higher. They suggested that direct larval exposure to oil in the
pelagic zone may not be the most important point of vulnerability for the crab population. Rather, effects
might be indirect, operating at the larval source and, in turn, affecting larval delivery to the near-shore
habitat.

The nature of the spill in terms of its magnitude and release at depth raised concerns regarding effects on
deepwater coral communities. White et al. (2012) studied coral communities at 11 sites three to four
months after the well was capped. Healthy coral communities were observed at all sites greater than 20
km from the well. However, at one site 11 km southwest of the well, coral colonies showed several signs
of stress. Of the corals examined at the affected site, 46% exhibited evidence of impact on more than half
of the colony, and nearly a quarter showed impacts to more than 90% of the colony. Floc deposited on the
corals was traced to the Macondo well via analysis of hopanoid biomarkers (White et al. 2012). The
significance of particulate deposition (i.e., mortality caused by smothering) from the failed ‘top kill’
operation and/or the episodic release oil (with similar biomarker ratios) from natural seepage (which has
been reported to occur within the region) has not been fully addressed.

8.5.5 Effects on trophic level dynamics

There is currently much interest within the NRDA process to identify potential community and
population level changes, as well as alterations in trophic level dynamics. Graham et al. (2010) found that
δ13C depletion, used as a tracer of oil-derived carbon in mesozooplankton and suspended particulate
samples in near-surface and bottom waters from four stations in 8 to 33 m water depth in the northern
Gulf of Mexico, corresponded with the arrival of surface slicks from the DWH oil spill and demonstrated
that carbon from the spill was incorporated into both trophic levels of the planktonic food web.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 303
Results of in situ microcosm experiments performed in North Inlet Estuary, SC, with DWH and a Texas
crude oil showed a decrease in chlorophyll a in phytoplankton as crude oil concentration increased from
10 to 100 µL/L (Gilde and Pinckney 2012). This change was interpreted as a decrease in biomass rather
than chlorophyll due to the short duration of the experiments. Diatom, cyanobacteria, euglenophyte and
chlorophyte abundances were unaffected or increased with increased oiling, whereas cryptophyte
abundance decreased. The authors suggested that oiling could result in changes in phytoplankton

The ecosystem services approach provides a useful framework regarding how to manage human
activities (and the consequences of those activities) in order to sustain the ecological structure and
function necessary to provide essential services. The approach can be used to inform the public and
decision-makers about the connections between human activities and the effects of those activities on
ecological services. However, in order to advance the approach beyond generalizations, data are
required which describe explicit connections between ecosystems and benefits (economic, cultural
and spiritual). These data are fragmentary and scattered among many disciplines.

community composition within salt marsh estuaries impacted by oil from the DWH, thereby affecting
higher trophic levels, such as zooplankton, which selectively feed on phytoplankton that might be killed
by the oil (Gilde and Pinckney 2012).

Tarnecki and Patterson (2015) recently reported a potential shift in the diet (prey including fish, decapods,
cephalopods, stomatopods, gastropods, zooplankton and other invertebrates) and trophic position
(determined from stable isotope ratio-mass spectrometry analysis of δ13C, δ15N, and δ34S) for red snapper
(Lutjanus campechanus) from the north-central Gulf of Mexico following the DWH oil spill. Stable
isotope data indicated a post-spill increase in red snapper trophic position (15N enrichment) and an
increase in benthic versus pelagic prey (34S depletion), that was consistent with observed dietary shifts,
likely linked to relative abundance of prey resources.

Indirect effects on the menhaden (Brevoortia tyrannus) population in the Gulf occurred due to trophic
cascade effects (Jeffrey Short, pers.comm. 2015). The effects began with the death of close to 1 million
seabirds due to physical contact with the oil. The mortality of birds increased the abundance of juvenile
menhaden—a primary food source for the birds. The menhaden population increased to unprecedented
abundance and biomass levels in 2011-2012. However, this was accompanied by a decline in the
condition of the fish (and thus their oil content) as they exhausted their food supply. These effects
occurred on a much larger spatial and temporal scale than direct toxicity from the spill. As of 2015, the
menhaden population appeared to be recovering.

8.5.6 Effects on ecosystem services

An ecosystem services approach can supplement traditional methods of assessing or valuing damage to
natural resources by estimating flows of goods and services before and after an event (NRC 2013;
reviewed in Chapter 4). The approach focuses not on the natural resources themselves, but on the goods
and services these resources supply to people.

Ecosystems are subject to natural disturbances, such as floods, droughts and disease outbreaks, as well as
human-caused disturbances, including oil spills. Ecosystems are also subject to slowly changing long-
term stresses, such as nutrient enrichment and changes in the sediment supply, as observed in the Gulf of
Mexico. These long-term stresses can affect the ability of the system to respond to a shock, such as the
DWH spill (NRC 2013).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 304
Ecosystem resilience to shocks from oil spills in the context of long-term stresses is key to the
maintenance of ecosystem services. In some cases, systems will undergo fundamental shifts in structure
and function following disturbances (e.g., changes in trophic structure caused by invasive species in the
Great Lakes). Although resilience to preserve ecosystem services is an important concept, the
understanding of complex and highly variable systems, such as the Gulf of Mexico (or the Gulf of St.
Lawrence) is insufficient to support specific, resilience-based recommendations regarding response to or
monitoring of oil spills. If policies and decisions are to be based upon the ecosystem services concept, a
large, coordinated, long-term, multi-disciplinary effort will be required.

NRC (2013) summarized the effects of the DWH spill on four key components of the Gulf of Mexico that
provide ecosystem services. Their conclusions are presented below.

Wetlands

Acute effects on marshes, where the biota are not expected to recover, appear to be confined to the
edges of bays, canals and creeks in a limited subset of the oiled wetlands.

 Where the vegetation has died and root systems have been lost in heavily oiled areas, the
erosion of sediment is leading to the conversion of once-productive marshland to open water;
 Subsequent tropical storm activity resulted in additional erosion of oiled marshes; and
 Based on numerous studies that document a rapid recovery from oiling and a relatively low
sensitivity of perennial marsh vegetation to hydrocarbons, marsh vegetation can be expected
to suffer little or no long-term impairment in areas where roots and rhizomes survived the initial
impact of oil fouling. If roots and rhizomes do not survive, then an area will likely not recover on its
own due to the loss of habitat by erosion.

These impacts need to be viewed in the context of significant and continuing losses of wetlands in the
Gulf of Mexico due to many other stressors, including subsidence, canal dredging, salt intrusion and
sediment starvation.

Fisheries

Despite long-term studies and ongoing development of models, the ability to detect spatial and temporal
differences in fishery productivity in the Gulf of Mexico is limited. Recent developments in fishery data
collection, such as the introduction of vessel monitoring systems in the reef fish fishery, could improve
estimates of abundance. However, any mortality or reduction in individual fitness caused by the spill
directly or indirectly may take years or, for some species, decades to transfer through the ecosystem and
be observed.

The direct value of commercial fisheries to the fisherman is calculated using the dockside value of the
catch minus any expenses incurred to capture those fish. The method for evaluating the economic effects
of an oil spill on commercial fisheries is derived from either reduced production (due to mortality or
fishery closures) or by reduced consumer demand (due to the perception of reduced fish quality or safety).

The immediate economic impact of the DWH spill was a 20% decrease in commercial fish catch for
2010.

Marine Mammals

Bottlenose dolphins were the representative marine mammal examined by the NRC because of the role
dolphins play in three ecosystem services—regulating, supporting and cultural. As apex predators 5 ,

5
An apex predator is a predator residing at the top of a food chain on which no other creatures prey upon.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 305
dolphin populations serve as important indicators. They are the most studied and among the most popular
and charismatic marine mammals. The stranding of hundreds of dolphins in the Gulf of Mexico before,
during, and after the spill stimulated considerable public concern. NRC (2013) suggested that if the post-
spill mortality event is determined to be linked to the spill, then an opportunity may exist to establish a
plan to protect and restore the dolphin habitat and to reduce dolphin mortality due to human activities.
Venn-Watson et al. (2015) studied unusual mortality events of bottlenose dolphins (Tursiops truncates) in
Louisiana, Mississippi and Alabama from 2010 to 2014 and concluded that the rare, life-threatening and
chronic adrenal gland and lung diseases identified in stranded dolphins were consistent with exposure to
petroleum compounds. The authors suggested that the DWH spill was a contributor to increased dolphin
deaths.

Deep Gulf of Mexico

The deep sea area of the Gulf is so vast and sampling is so sparse that gaps in knowledge inhibit the
ability to apply an ecosystem services approach in a quantitative way. The NRC (2013) assumed that the
primary ecosystem services of the deep Gulf are supporting, e.g., resupply of nutrients. In addition, the
NRC (2013) stated that the deep sea area of the Gulf provides the regulatory service of pollution
attenuation, e.g. via bacterial degradation. The release of crude oil from the Macondo well at depth
created a unique opportunity to study deep sea oil biodegradation.

8.5.7 Important factors affecting the consequences of the Deepwater Horizon Blowout

The huge volume of spilled oil and the length of time it took to stop the release were primary factors in
determining consequences. The presence of significant anthropogenic stressors prior to the DWH spill,
including degraded or lost shoreline habitats, overexploitation of some fishery resources, salt intrusion,
sediment starvation and nutrient influxes leading to algal blooms and oxygen depletion, added
confounding factors to the interpretation of the effects of the spill. In addition, the Gulf of Mexico is a
highly complex ecosystem subject to major natural cycles.

8.5.7.1 Oil properties and behaviour

 Dissolution of BTEX and naphthalenes into the water column occurred as the oil moved to the
surface from the well.
 There was rapid evaporation of lighter fractions in the floating oil.
 Substantial spreading of surface oil occurred driven by wind and currents.
 Biodegradation along shorelines and within dispersed oil was significant, with a fairly quick
response of native microbial degraders to the oil.
 Sinking of physically- and chemically-dispersed oil droplets from mid-column occurred.

Natural dispersion and oil weathering processes rapidly reduced the concentrations of TPH and PAH
within surface and subsurface waters after the well was capped. In analysis of over 20,000 water samples
from over 13,000 stations, 84% of TPH and 79% of PAH values were below estimated background levels.

8.5.7.2 Effect of the environment on fate and behaviour

 High microbial degradation rates occurred along shorelines and in shallower waters of the Gulf
because of subtropical temperatures (Chapter 3).
 Rapid penetration of oil occurred along sand shorelines due to high porosity.
 Repeated burial and exposure of oil occurred (and continues to occur) along shorelines.
 Wave action led to stranding of oil in the high intertidal zone.
 Heavy coating of vegetation along the margins of wetlands and estuarine shorelines including the
formation of ‘oiled’ subsurface mats.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 306
8.5.7.3 Oil toxicity

 Mass acute mortality occurred due to exposure to floating oil.


 Death of vegetation occurred along heavily oiled margins of marshes.
 The extent of chronic toxicity is difficult to determine given multiple confounding factors.
 There may have been trophic level effects, including a shift in the diet of red snapper and a
temporary increase in menhaden abundance originating with the mass mortality of predatory
seabirds.
 Chemical and biological markers of oil exposure and toxicity were not often included in case
studies of oil effects on aquatic biota.
 Extrapolation of laboratory and mesocosm study results to the Gulf of Mexico ecosystem is very
difficult.

8.5.7.4 Spill response

 Novel application of dispersants in the deep subsurface was associated with the formation of a
deepwater plume.
 Surface application of dispersants had limited effect in reducing oil reaching shorelines.
 Deployment of oil booms, berms and other barriers protected specific shoreline or island areas.
 ISB removed 6 - 7% of the oil.

8.5.8 Lessons learned from the Deepwater Horizon Blowout

8.5.8.1 Lessons learned according to the USGS (2015)

 Oil was consumed by bacteria, seafood was not contaminated by hydrocarbons or dispersants and
the oil budget was by and large accurate. There was consensus that most of the oil was
biodegraded. The only part of the oil budget that was later found to be inaccurate was the fraction
of oil that was chemically-dispersed versus naturally-dispersed. That information had no impact
on public safety, seafood safety or the response effort, but understanding the amount of oil that
was dispersed chemically versus naturally is important for future such efforts.
 The scale and complexity of DWH taxed the agencies involved.
 Future oil spill response preparedness should include the following actions:
 Gathering adequate environmental baselines for all regions at risk;
 Developing new technologies for rapid precise reconnaissance and sampling to support a
timely and robust response effort;
 Filling large information gaps regarding biological effects of oil, changing climate and
other simultaneous drivers of variability in coastal and aquatic ecosystems;
 Requiring that future oil extraction permits be conditional on having mechanisms in place
to rapidly assess flow rate; and
 Conducting research on the impacts of dispersants and dispersants-plus-oil on a wide
range of species and life stages.
 The scale of the spill required unprecedented collaboration among government, academic and
industry scientists and engineers. Scientific and engineering information was crucial to guide
decision-making for questions never before encountered.
 The lack of peer-reviewed scientific publications from prior marine well blowouts was a
significant drawback in addressing many of the issues.
 The event also showed the value of federal partnerships with academic institutions.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 307
8.5.8.2 Other lessons learned (from the RSC Panel’s review of the literature)

 Considering the size of the spill, relatively little oil made it to shore and tar balls and tar mats
were distributed heterogeneously.
 Receiving environments that are already degraded by human activities (e.g. coastal marshes) have
enhanced vulnerability to oil spills.
 Fate and transport of submerged, physically- and chemically-dispersed oil is poorly understood.
 The consequences of deposition of oil to deepwater sediments require investigation.
 The ability to determine effects of oil spills (even those as huge as the DWH) depends upon the
availability of pre-spill baseline information, an understanding of natural variability, and an
understanding of the effects of multiple anthropogenic stressors.
 Barriers to prevent oil incursion can have unintended consequences (e.g., reduced available
habitat for lemon sharks and inhibition of sharks from reaching the protected sea grass beds on
the leeward side).
 Natural biodegradation on beaches in warm climates can be rapid and may not require nutrient
supplementation, especially if available nutrients are already plentiful enough to support growth
on the oil incursion. However, biodegradation will not remove all of the oil components.
 Effects on trophic dynamics via changes in biomass of primary producers and shifts in diet can
occur. However, the spatial and temporal scales at which this could occur suggest that effects
may not be widespread and/or long-term.
 Oil spill response plans should accommodate and exploit scientific opportunities and oil spill
response should incorporate these opportunities.
 Experience of people involved in the response to the spill showed a pressing need for improved
analytical procedures for the detection and characterization of oil spill constituents, including
standard sample tracking and reporting procedures (Laboratory Information Management linked
to sample collection) (Ken Lee, pers. comm.).
 Effects on ecosystem services can be very difficult to establish and quantify, particularly for
resources where the ability to detect spatial and temporal differences is limited by lack of data or
a lack of fundamental understanding of population or community structure, function and
dynamics, or for ecosystem components that have had very little study (e.g., deep sea areas).
 Even if effects on ecosystem services can be quantified, a comprehensive model is needed that
incorporates biophysical, social and economic data for the Gulf of Mexico for the long-term
(NRC 2013).

8.5.9 Research Needs Related to the Deepwater Horizon Blowout, in the Canadian Context

Recommendation: Research into the consequences of deposition of oil to deep sea sediments in the
Arctic, as well as deep sea sediments south of the 60th parallel in Canada would provide part of the
required knowledge base for support of decisions regarding offshore oil and gas exploration and oil
transport. This is particularly important in light of frontier offshore oil and gas operations moving
into deeper waters and emerging interest in the use of dispersants by subsea injection in the case of
well blowouts. Offshore drilling in the Arctic and on Canada’s east coast will create the potential
for impacts on benthic ecosystems and commercial fisheries that have not yet been studied.

Recommendation: In the Canadian context, the current level of understanding of trophic dynamics
in identified high-risk offshore and inshore marine habitats requires evaluation in order that
critical gaps in understanding are identified.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 308
8.6 Pine River, British Columbia pipeline break

On the night of August 1, 2000, a break in an oil pipeline released 952 m3 of sour crude oil (BC-Light;
Table 2.2) into the Pine River in northeastern British Columbia. The Pine River is a major tributary of the
Peace River with mean flows in August of about 200 m3/s (Water Survey of Canada 2015). About half of
the oil entered the river and was dispersed downstream. The other half remained on land, contaminating
surficial soils and groundwater. The Pine River is a source of drinking water for the town of Chetwynd.

The spill occurred in the vicinity of a previous spill of 30 m3 of gasoline and 24 m3 of diesel fuel on
August 18, 1994, that caused the mortality of about 150 mature and 1,000 juvenile fish (Goldberg 2011).

8.6.1 Immediate effects, response and cleanup

As a result of the spill, Chetwynd closed its water intake and started exploring alternate water sources. A
total of 1,637 dead fish were found 2-50 km downstream, comprised primarily of larger, more visible fish.
Most were bottom-feeding species, consisting of 64% mountain whitefish (Prosopium williamsoni), 16%
slimy sculpins (Cottus cognatus) and 5% burbot (Lota lota). The remaining fish were surface feeders,
consisting of 5% bull trout (Salvelinus confluentus), 2% rainbow trout (Oncorhynchus mykiss) and 1%
Arctic grayling (Thymallus Arcticus) (Alpine Environmental and EBA Engineering 2001). It was
estimated that 15,000 – 27,900 fish mortalities occurred, based on numbers of fish per kilometer and
assuming complete fish-kill in the first 30 km.

The recovery of oil from the river occurred over a period of about two months. According to Alpine
Environmental and EBA Engineering (2001), about 91% of the oil was recovered or accounted for as
follows:

 447.5 m3 entered the river;


 358 m3 recovered as liquid oil from the river and spill site prior to soil excavation;
 89.5 m3 lost to volatilization (calculated);
 5 m3 recovered from the river as liquid oil using absorbents;
 21.4 m3 recovered as liquid oil during soil excavation;
 416 m3 recovered from the spill site as part of soil excavation; and
 83.2 m3 not accounted for (included dissolved, adsorbed in soils along bank and in sediments,
trapped in backwaters and eddies, trapped in logjams).

A number of emergency response operations were deemed to have caused detrimental effects on fish and
wildlife habitat. These included:

 The alteration of an 115 m-long backchannel from the break site to the river (to stop oil flow into
the river) that resulted in the cut off of an oxbow channel that provided abundant fish food and
the isolation of a back channel from fish access;
 Bank armouring to prevent erosion and the removal of logjams and riparian vegetation coated
with oil removed floodplain habitat available to fish during high flows (Sumners 2001). In this
case the contractor also failed to implement the consultant’s recommended fish habitat features
when installing the armouring;
 The removal of logjams and woody debris structures from the river channel without consultation
and approval of regulatory agencies (Department of Fisheries and Oceans-Canada [DFO] and BC
Ministry of Environment);
 Rerouting of the river near a tributary by the removal of a logjam diverted flow from a 3.4 km
stretch of creek, the Lemoray Meander Loop. The new 1.6 km long channel that formed was
unstable, having a steep gradient that resulted in mass erosion with vegetation, soils, gravels and
sands washing downstream into spawning habitat—however there was no investigation to access

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 309
the level of potential impact. While the logjam was eventually rebuilt, it was overwhelmed by the
spring freshet in 2001 a 1-in2- or 1-in-3-year event). Sumners (2001) deemed this failure as
“catastrophic to the stream channel and the habitat therein”. Consulting engineers involved in the
project concluded that the magnitude of the changes in the channel prohibited the effectiveness of
further attempts at logjam construction; and
 Habitat damage was also caused by the actions of the regulatory agencies which allowed large
machinery in the river channel because of the urgency to finish the cleanup work before winter.

Sumners (2001) noted that poor records were kept during the cleanup. In terms of remediating the damage
to habitat caused by the initial emergency response operations, at least 13 logjams were rebuilt on advice
from a consultant, primarily to maintain the stability of the river. Sumners (2001) expected that where
jams were not rebuilt, others would likely form over time; however, the author concluded a net loss of
habitat resulted.

8.6.2 Water quality

Visual and olfactory evidence of oil contaminants were evident over an 80-km section of the river
following the spill. After three weeks, while concentrations of extractable petroleum hydrocarbons were
still detected in waters downstream of the spill site, they were below water quality criteria (Alpine
Environmental and EBA Engineering 2001), and few samples contained the more volatile fractions
(Amec 2001a). By the end of August (one month later), petroleum hydrocarbons in the water had declined
to below detection limits (Amec 2001a) and the presence of sheen was largely restricted to back eddies
and other calm water areas along shorelines.

8.6.3 Sediments

Petroleum hydrocarbons accumulated in depositional environments within the river, in areas with soft,
muddy sediments along the banks, in back eddies and other calm-water locations (Amec 2001b). There
were also high concentrations of organic debris (e.g., branches, leaves and algae) that had accumulated
along the shoreline, in front of logjams and attached to sweeper logs (overhanding trees with some limbs
and branches submerged during high flows). Six PAH compounds and cadmium were detected in these
materials at concentrations sufficient to be a concern for the health of fish and other aquatic biota (Amec
2001b). Heavy rainfall in late August and September mobilized previously stranded oil and oil-
contaminated debris and sediments into relatively uncontaminated areas. Higher water levels also scoured
some of the more heavily-contaminated depositional areas. The concentration of detectable oil
concentrations declined by an average of 71% by October (Alpine Environmental and EBA Engineering
2001).

8.6.4 Periphyton and benthic invertebrates

Effects were observed on periphyton and benthic invertebrate communities immediately after the spill. A
substantial increase in algal biomass that exceeded the apparent seasonal effects was observed at one
sampling station (Alpine Environmental and EBA Engineering 2001). Total abundance of benthic
invertebrates declined in August, with substantial recovery by November (Alpine Environmental and
EBA Engineering 2001). The benthic invertebrate community structure was also affected. The effects on
benthic invertebrates were of concern with respect to a decreased food supply for fish, especially young
fish, during the recovery period between August and November—a critical period leading up to over-
wintering (see discussion below).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 310
8.6.5 Fish

Chemical analysis of sportfish sampled above and below the spill site did not indicate any trends in
concentrations of total extractable hydrocarbons or PAH in liver and muscle tissue (Amec 2001c). Amec
concluded that consumption of the Pine River fish caused no significant unacceptable risk to humans.

The initial spill of crude oil into the Pine River produced sufficient toxic material (especially that which
dissolved to produce a water-soluble fraction likely a concentration of higher than 3 mg/L, Chapter 4) in
the river that killed fish over a relatively short period of time (hours) (Birtwell 2003). Unfortunately, the
actual concentrations in river water were not determined as samples were not collected at the time when
the fresh crude oil was entering the river (Birtwell 2003).

Results of analysis of fish tissue and stomach contents indicated that longer-term risks to fish can be
driven by feeding behaviour combined with oil dispersion and stranding on large woody debris. Analysis
of fish bile showed no benzo(a)pyrene metabolites but some phenanthrene metabolites, indicating at least
some PAH exposure (Alpine Environmental and EBA Engineering 2001). Although concentrations in
whole fish were at levels of detection for total PAHs in surface feeders, predators and bottom feeders, the
stomach contents of rainbow trout showed large quantities of unmetabolized PAHs. These fish were
sampled at a logjam where a considerable amount of free oil was collected, suggesting that the fish were
feeding within the pooled oil on drifting invertebrates. The hydrocarbons in the stomach contents
mirrored the oil characterization results (Alpine Environmental and EBA Engineering 2001).

Exposure of fish to hydrocarbons can also be driven by temperature-related behaviour. Fish often conceal
themselves in the substrate at temperatures less than 9 °C, particularly bull trout (Alpine Environmental
and EBA Engineering 2001). If the concealment areas coincide with areas of oil deposition, the fish will
be exposed. No consideration of hyporheic flows and effects on spawning gravels was apparent in the
published reports about the spill.

The annual fidelity of fish to feeding areas could result in reduced food intake if prey species were less
abundant or unavailable due to the impact of spilled oil (Birtwell 2003). Fish must obtain enough food
prior to ice cover to survive the cold northern winter. However, if this is not accomplished and/or the
metabolism of the fish is elevated due to exposure to contaminants and/or other stressful circumstances,
survival is jeopardized (Lemly 1993 cited by Birtwell 2003). Because exposure to oil can elevate
metabolism in fish, their survival under the colder winter conditions was a concern, especially at a time
when metabolic activity and food intake usually decreases (Birtwell 2003).

The concerns about sublethal effects on fish led to an assessment of snorkel surveys conducted in the Pine
River in 1993 (pre-spill), 1994 (after the 1994 spill), 2000 (about two months after the pipeline rupture),
2005, 2006 and 2007 (Goldberg 2011). The results suggested that the sportfish species composition
(mountain whitefish, Arctic grayling, bull trout and rainbow trout) and abundance observed in 2005, 2006
and 2007 were similar to pre-spill observations (1993) in the river sections surveyed.

In summary, the 2000 spill caused mortality of several thousand fish, but there were insufficient studies to
establish, with confidence, the nature and extent of sublethal effects on reproduction and recruitment. The
concentrations of PAH compounds in fish stomach contents confirmed PAH exposure via the food chain.
There was a potential for concentrations of PAH in spawning habitat to have exceeded thresholds for
sublethal effects, particularly for fall spawning species such as bull trout in 2000. However, there were no
coordinated efforts to measure PAH concentrations and fish egg/embryo survival in spawning areas
downstream of the spill.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 311
8.6.6 Important factors affecting the consequences of the Pine River spill

8.6.6.1 Oil properties and behaviour

 Evaporation of light oil plus dissolution and natural dispersion were important factors leading to
hydrocarbon concentrations within the water column of the river declining to detection limits
within one month.
 Physical interactions with shorelines, sediments and woody debris led to substantial stranding of
oil.

8.6.6.2 Effect of the environment on fate and behaviour

 Water flow led to rapid dispersion of acutely lethal concentrations of the residual oil up to 80 km
downstream.
 Temperature declines in the fall triggered fish behaviour that increased exposure to oil in
sediments.
 Substrate particle size was an important factor leading to retention of oil in fine sediments in
slow-water areas.
 River channel and shoreline characteristics were also important with respect to creating
conditions that allowed oil to become associated with the river bottom (backwater and side
channel areas).

8.6.6.3 Oil toxicity

 Substantial acute lethality to fish and benthic invertebrates occurred immediately after the spill.
 Most fish deaths were of bottom-feeding species. Exposure may have been via both water
and food organisms based on stomach content analysis; and
 Mortality of benthic invertebrates was observed in the first month but rapid recovery was
observed.
 There was possible increased overwintering mortality or increased susceptibility to predation
(because of decreased fitness) caused by exposure to PAHs in overwintering habitats and possible
effects on embryo survival and recruitment due to oil deposition in spawning areas.
 There was no demonstrated avoidance of oiled areas by fish (anecdotal evidence only);
 Fidelity to feeding and overwintering areas may have exposed the fish to oil in sediments
and food organisms, or reduction in food abundance may have reduced fitness going into
the winter season; and
 Metabolizing the PAHs may have created an energy cost that affected fitness.
 Population-level effects.
 Similar abundance and species composition of fish in snorkel surveys conducted in 1993
and 2005, 2006 and 2007. However, the natural variability of abundance and the
influence of migration from un-surveyed areas is unknown, and there were no measures
of reproduction or recruitment; and
 No definitive statements can be made regarding the presence or absence of population-
level effects.

8.6.6.4 Spill response

 Active recovery of liquid oil and oil-plus-soil from the river and spill site removed about 84% of
the total spilled oil; 91% of the oil was either removed or accounted for.
 Physical removal from the river channel, shorelines and woody debris.
 Direct damage to habitat from heavy machinery within the river channel;

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 312
 Changes in channel morphology caused by bank armouring cut off fish access to an
important back-channel habitat area; and
 Removal of logjams eliminated important fish habitat and altered flow dynamics, in some
cases to such an extent that significant areas of side channel habitat were lost, and
subsequent seasonal increases in flow restructured channels and degraded fish habitat.

8.6.7 Lessons learned from the Pine River spill

 Fish kills in moderate-sized rivers can extend far downstream if the spill is not contained
immediately.
 The extent of impacts on fish can be obscured if there is a slow response to the spill (e.g., remote
spill site) and dead fish are consumed by scavengers before being identified.
 Benthic invertebrate communities can recover quite quickly after spills that occur in summer
months, provided the initial scale of the impact is sufficiently small.
 The effects of cleanup activities were substantial, decreasing or eliminating important fish
habitats.
 Data were insufficient to support any definitive statements regarding long-term effects on fish
abundance and species composition.
 Opportunities for more detailed and sophisticated investigation of the effects of the spill on were
not pursued, perhaps due to logistic and financial constraints.

8.6.8 Research needs related to the Pine River spill

Recommendation: Standard guidance is required for measurement of oil contamination of


sediments, pore water and biota within flowing waters of Canada. This guidance should consider
the range of lotic receiving environments and logistic constraints common to remote locations. The
guidance should include the de minimis level of investigation required for decision-making
regarding cleanup requirements and techniques, as well as for determination of acceptable residual
levels of oil.

Recommendation: The Pine River spill offered the opportunity to track longer-term response to
cleanup activities, as well as to residual oil. Such opportunities should not be lost if such spills occur
again, because the longer-term trade-offs involved in selection of cleanup methods remain
uncertain.

Recommendation: Standard guidance for monitoring the fate of spilled oil in all major
environmental compartments of lotic systems is required.

Recommendation: Research is needed on the relative role of hyporheic flows in contributing to


exposure of fish to hydrocarbons after a spill. A range of experimental conditions using relevant
concentrations and both weathered and unweathered oil should be tested on salmonid species (and
demersal spawners in general) under conditions that represent actual field conditions in terms of
variables such as flow, temperature, substrate and dissolved oxygen.

Recommendation: The Pine River spill case study illustrates the importance of baseline information
with which to compare post-spill monitoring data. The Panel strongly recommends the
establishment of a national baseline database for freshwater systems adjacent to pipeline corridors
(e.g., data collected for EIAs, environmental effects monitoring (EEM) programs, compliance
monitoring [e.g. for municipal or industrial discharges], research programs, etc). The Panel is
aware that some databases are already established (e.g. for EEM data); however, there is no readily
accessible, national portal to metadata on freshwater systems.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 313
Recommendation: The Pine River spill could have yielded highly valuable data regarding effects of
spills in riverine systems in the short, medium and long-term. The Panel recommends that there
should be national guidance regarding ‘spills of opportunity’, by implementing communication and
coordination protocols among agencies and industry to optimize the collection of scientific
knowledge and lessons learned from incidents.

8.7 Wabamun Lake, Alberta Train Derailment

On August 3rd, 2005, 43 Canadian National Railway Company (CN) rail cars derailed immediately
adjacent to Wabamun Lake (Figure 8.6), approximately 65 km west of Edmonton, AB. The lake (area =
82 km2; mean depth = 6.3 m; maximum depth = 11 m) is moderately to highly enriched with nutrients and
is generally well mixed and thus generally well oxygenated throughout the water column during the open-
water period (Prepas and Mitchell 1990; Hollebone 2008).

Of the 43 derailed train cars, 11 containing heavy fuel oil (HFO 7102, a type of Bunker C oil; Table 2.2)
ruptured, spilling 712 m3. A single car carrying Imperial Pole Treating Oil6 (PTO) also ruptured spilling
about 88 m3 on to the ground at the derailment site. A total of about 149 m3 of heavy fuel oil entered the
lake (Birtwell 2008). The oils ran onto the lawns of cottages about 100 m from the lakeshore. HFO
entered the lake less than 1.5 hours after the derailment along a broad front of about 0.5 km (Figure 8.8).
The flow was aided by the fact that the HFO 7102 had been loaded a few hours before and was still warm
and relatively less viscous than it would be at ambient temperature (Hollebone 2008).

Figure 8.8 Wabamun Lake Train Derailment. Image from: Dangerous Goods Newsletter, Spring 2006. Transport
Dangerous Goods Directorate, Transport Canada (https://www.tc.gc.ca/eng/tdg/newsletter-spring2006-323.htm)

8.7.1 Early behaviour and effects of the spill

Initially, all of the oil appeared to be floating on the surface, rapidly spreading in warm, calm weather
conditions, and escaping boomed areas once the wind increased. Over 1.1 million m2 of lake surface was
visibly oiled, of which 63% was heavily coated (Birtwell 2008). A few days later, strong westerly winds
and waves concentrated the oil along the north, east and south shorelines (Anderson 2005). Langmuir
circulation patterns during the windy conditions contributed to the rapid horizontal movement of the oil,
as well as vertical movement of oil droplets into the water column. Much of the shoreline was boomed off
and oil was trapped with varying degrees of efficiency in the littoral zone (Anderson 2005; Hollebone et
al. 2011). Measurement of hydrocarbon concentrations within the boomed area showed concentrations
well in excess of guidelines for the protection of aquatic life. Concentrations of BTEX and PAHs in open-

6
A hydrocarbon-rich liquid used for preserving wooden poles, containing naphthalene and other PAHs
(http://www.enr.gov.nt.ca/sites/default/files/pahs.pdf)

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 314
water areas of the lake outside of the boomed areas were well below water quality guidelines (Anderson
2005).

Tar balls quickly formed (from < 1 to 10 cm in diameter) in near-shore areas. The rapid formation of tar
balls was attributed to the loss of volatile components from the spilled oil which also picked up fine
mineral and organic particles along its overland flow-path to alter its density (Parker-Hall and Owens
2006). Within hours, some of the tar balls were showing neutrally-buoyant behaviour, and were seen
riding up and down in the water column. Some would rise to the surface, others would be seen sinking to
the bottom. In addition to tar balls, tar logs were observed, up to 30 cm in diameter and 5 m in length,
consisting of a mixture of organic debris and oil (Hollebone 2008). Some near-shore areas had extensive
tar mats.

Dead fish were observed in oiled areas immediately following the spill and for about two months
afterward. A total of about 100 dead fish were observed by cleanup crews. Live fish captured two weeks
after the spill showed biochemical evidence of exposure and had elevated levels of hydrocarbons in
tissues (Hodson et al. 2007; Birtwell 2008).

People in the Whitewood Sands cottage community (where the derailment took place) were evacuated
within 15-20 minutes of the derailment and allowed back to their properties later that evening (Lake
Wabamun Residents Committee 2007). The health authority advised residents on the day of the spill to
avoid using the lake until further notice. In August, 2008, three years later, advisories were still being
issued against eating certain wildfowl and the fishery was catch-and-release only.

8.7.2 Shorelines

During the Shoreline Cleanup and Assessment Technique (SCAT) survey in 2005, the majority of tar
balls or tar mats were observed near the shores in water depths from 0.1 to 1.5 m, with their frequency
decreasing with increasing depth (Hollebone 2008). After shoreline treatment a high proportion of oil
remained. The coverage of oil varied but was highest in treated reed beds (Figure 8.9). In the reed bed
areas on warm days, tar balls would rise to the surface, shed oil from several points around their
circumference and create a sheen (Hollebone 2008). Beach re-oiling continued until freeze-up during the
first winter.

Figure 8.9. Near-shore reed bed with tar balls in Wabamun Lake. Image from Hollebone (2008).

By the spring of 2007, a SCAT survey showed almost all oiling to be in the Very Light or Trace
categories found in marsh, peat-soil or vegetated bank shoreline types.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 315
8.7.3 Weathering

Four months after the spill, about 79% of the HFO 7102 and 47% of the PTO had been recovered from
the lake spill site, and flow path to the lake and water concentrations had decreased to below detection
limits (Birtwell 2008). According to Birtwell (2008), acutely toxic PAH concentrations were likely
limited to the reed beds, where high concentrations of PAHs in the water would be promoted because the
water circulation would be poor and the surface area of the oil large.

While on the surface of the lake, the oil weathered considerably with a loss of most of the lighter
components. Oil coating the reeds was heavily weathered 40 days after the derailment. The weathering of
tar balls varied, with some tar ball samples collected about 80 days after release showing only moderate
weathering. Some free-floating oil was able to survive until near freeze-up in a largely fresh, lightly-
weathered state (Hollebone 2008).

A year and a half after the derailment in February 2007, oil was found in the lake in primarily two forms:

 Large (> 5 cm) flat conglomerations on the lake bottom, often tangled into vegetation and highly
weathered, with high sediment loading, water content and viscosity; and
 Small (< 5 cm), less chemically-weathered spherical balls of soft, fluid oil containing a
significant fraction of the original aromatic content of the fuel oil, surrounded by a tough, more
weathered layer of oil, which were easily stirred up from the bottom and readily moved with
currents and wind due to lower density and viscosity (Hollebone 2008).

Virtually all of the alkanes and ‘Priority Pollutant’ PAHs had attenuated by the spring of 2007 (Parker-
Hall and Owens 2007). Tar balls, particulates and the oil coating the vegetation contained primarily
asphaltenes (which are unregulated chemicals) and low concentrations of non-‘Priority Pollutant’ PAHs
(Parker-Hall and Owens 2007). Foght (2006) stated that while competent hydrocarbon-degrading
microbes existed in lake sediment, the mass of oil they were likely to be able to degrade was small due to
the recalcitrant nature of the oil in tar balls and coating the vegetation. Foght also pointed out that the tar
balls would limit microbial access to the biodegradable components of the spilled oil and concluded that
the prognosis for extensive natural biodegradation over 5-10 years was poor.

Given the amount and chemical nature of residual oil in the spring of 2007, Parker-Hall and Owens
(2007) concluded that further treatment or recovery of sunken oil beyond that planned for May and June
of that year was not practical or feasible and would pose a risk of further damage to lake bottom and
near-shore communities. Short (2008) estimated that cleanup activities recovered more than 95% of the
oil components that were not lost through natural weathering.

8.7.4 Observed effects on aquatic biota

The chemistry of the heavy fuel oil spilled in Wabamun Lake differed in specific ways from typical
Bunker C fuel oil and may have contributed to the observed toxicity, especially to fish embryos. The HFO
7102 product was primarily composed of saturated hydrocarbons and also contained a high content of
aromatic hydrocarbons (48%) and high concentrations of PAHs (60,400 µg/g oil) (Hollebone 2008). In
comparison, a typical Bunker C fuel has a total aromatic content of 29% and a total PAH content of
29,000 µg/g oil (Hollebone 2008). The PAHs of the HFO 7102 were predominantly alkylated
naphthalenes (Table 2.1) with smaller amounts of alkylated 3-ring phenanthrene and members of the
fluorene homologous series (Hollebone 2008). No BTEX compounds were detected except in the PTO.

8.7.4.1 Algae and macrophytes

Field observations of phytoplankton community characteristics and macrophyte survival and growth did
not indicate any effects of oil exposure overall (Golder Associates 2007 cited by Birtwell 2008). The spill

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 316
had little direct effect on the abundance and productivity of softstem bulrush (Schoenoplectus
tabernaemontani) (Thormann and Bayley 2008; Wernick et al. 2009) (see Chapter 4).

8.7.4.2 Zooplankton

Standard acute toxicity tests using the water flea species Daphnia magna and chronic tests using
Ceriodaphnia dubia showed no acute lethality but reduced growth near the spill site. The reduced growth
was considered to be a result of factors other than oil by Golder Associates (cited by Birtwell 2008b).
However, Birtwell (2008) argued that the significant results of tests between reference and oiled locations
in the lakes had validity even if background ambient conditions differed from those in the laboratory
setting. The characteristics of the zooplankton community did not change when compared with pre-spill
data (Golder Associates 2007 cited by Birtwell 2008).

8.7.4.3 Benthic invertebrates

No data existed on immediate effects of the spill on benthic invertebrates but physical smothering effects
were considered likely (Birtwell 2008). Laboratory toxicity tests showed significant toxicity to two test
species Chironomus tentans (a midge) and Hyalella azteca (an amphipod crustacean) from sediments
collected near the spill site. However, the proportion of sand in the sediment was an important
confounder, with higher mortality in high-sand substrates. Both test species showed significant reductions
in growth in both high- and low-sand sediments. PAH concentration was correlated with toxicity to the
nematode Lumbriculus variegatus, with significant reductions in growth in both high- and low-sand
sediments.

8.7.4.4 Fish

Risk to fish was due to direct toxicity immediately after the spill, and longer-term direct and indirect risks
associated with the near-shore areas, which were important for at least one life stage for the eight fish
species inhabiting the lake. The near-shore areas were primary nursery, rearing and food supply areas, and
fish did not avoid oiled areas. The summer timing of the spill increased the risk because it coincided with
the presence of numerous life stages of each of the eight fish species present in the highly exposed areas
(Birtwell 2008). Risk was also related to destruction of habitat during cleanup and survey activities.

Hodson (2008) noted that the exposure of adult fish to the spilled oil corresponded to the discovery of
significant numbers of dead fish in the oiled part of the lake. However, Hodson (2008) commented that
heavy oils with low concentrations of LMW compounds would be less likely to cause acute lethality. The
area closest to the oil spill was also affected by non-spill stressors, including: a thermal plume from a
power plant; physical removal of macrophytes (to address concerns of cottagers regarding access to the
lake for swimming and boating); and high fishing pressure (Schindler et al. 2004). Thus, fish inhabiting
the area nearest to the spill were already exposed to other stressors and may have been more susceptible
to the additional stress of the spill.

Within two weeks of the spill, a survey of adult fish demonstrated that exposure to PAHs from the spilled
oil was widespread; this exposure was relevant with respect to sublethal effects (Hodson 2008). A second
survey, about 10 weeks after the spill, showed continuing exposure of adult fish, although at a lower level
than in August (Hodson et al. 2007; Hodson 2008). One year after the spill, detectable PAH
concentrations were found in fish flesh. However, PAHs in fish decreased over time due to metabolic
processes in the fish plus weathering of the hydrocarbons in water and sediment (Birtwell 2008).

Embryo-toxic levels of PAHs were found in August 2005 through to spring of 2007. Toxicity studies
conducted in situ by Golder Associates (2007 cited by Birtwell 2008) showed embryo deformities in
caged eggs of lake whitefish (Coregonus clupeaformis) at spawning shoals in the winter and spring
following the spill and on northern pike (Esox lucius) eggs caged in the spring of 2007. Effects were most

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 317
pronounced in whitefish embryos with significant moderate to severe deformities correlated with PAHs in
lake water months after the spill. Pike suffered > 50% increased frequency of moderate to severe
deformities relative to reference areas but few deformities overall (Golder Associates 2007 cited by
Birtwell 2008; deBruyn et al. 2007; Hodson 2008). Uptake of PAHs into semi-permeable membrane
devices deployed at the egg incubation site confirmed a correlation between PAHs and egg deformities.

Although some embryos survived exposure, the marginal state of a once-thriving whitefish population in
Wabamun Lake raises concerns for cumulative impacts of oil and other anthropogenic stressors
(Schindler et al. 2004; Donahue et al. 2006) on recruitment and production of whitefish. The whitefish
fishery and all other sport fish in Wabamun Lake are catch and release only (Government of Alberta
2015).

8.7.4.5 Effects of the cleanup

Effects on fish habitat occurred due to the requirement to remove oil from Wabamun Lake (Evans 2008).
These actions resulted in loss of structural habitat and alteration and disruption of lake substrates.
Additional impacts occurred due to increased suspended sediment and turbidity during habitat removal
and cleaning, increased erosion along shorelines that became unprotected from wave action, the
remobilization of oil, and physical disturbance by trampling and similar activities.

Although the oil spill in Lake Wabamun did not have direct toxic effects on the reed beds, the ’treatment’
impacts in the deeper reed beds reduced plant and rhizome density, and this may have had long-term
consequences, remaining open or sparsely vegetated for several years and thereby providing lower quality
habitat for fish.

Fish were killed when aquatic vegetation was cut. The presence of fish within harvested vegetation was
probably less an issue of entrapment and an inability to escape than it was an example of their
vulnerability due to habitat fidelity (the need for cover, etc.).

Benthic habitat and the associated communities were impacted due to the removal of oil by the use of
numerous techniques, as well as the cutting of macrophytes and removal of whole plants. The techniques
included low-pressure flushing, vacuuming and dragging over substrates. Trampling in shallow waters for
survey work or for cleanup caused further damage. The impacts of the cleanup on the quality of habitat
persisted for several years (Hollebone 2008) (see Chapter 4).

A survey of riparian habitat along the shoreline of Wabamun Lake in 2015 showed that approximately
57% of the lake’s Riparian Management Area was in healthy condition, 9% was moderately impaired and
34% was highly impaired (North Saskatchewan Watershed Alliance 2015). Residential development was
cited as the major cause of riparian disturbance.

8.7.5 Analysis of the response to the spill

In a review of the response to the spill, McCleneghan (2008) noted that containment of much of the oil
within the first few hours of the spill would have been possible if the railway company had taken steps in
the years preceding the event to prepare for an on-water spill from its operations.

CN estimated that it spent $28 million on cleanup and $7.5 million in compensation to property owners
(Lilwal and Fitzpatrick 2015). The company was fined $1.4 million for its part in the disaster (Lilwal and
Fitzpatrick 2015). The Wabamun spill spurred the establishment of a new agency in Alberta to coordinate
disaster response. The Alberta Emergency Management Agency was created in 2006 (CBC 2006).
According to the Agency website, it leads the coordination, collaboration and cooperation of all
organizations involved in the prevention, preparedness and response to disasters and emergencies
(http://www.aema.alberta.ca).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 318
8.7.6 Public consultation and involvement

The spill generated outrage and protests, including a blockade over the CN tracks, with people accusing
CN of being more concerned with restoring rail service than stopping the spread of the oil. The blockade
ended after the company promised to meet with the public (Lilwal and Fitzpatrick 2015).

Ten years after the spill, some members of the public stated that the lake still suffered some lingering
damage, with reports of tar balls and oil buried in the sand. However, a representative of the cottage
community stated that CN had fairly compensated cottage owners and applauded the establishment of the
Alberta Emergency Management Agency and stronger regulations for railways (Lilwal and Fitzpatrick
2015).

Public consultation was conducted by Alberta Environment, including workshops to discuss the
establishment and operation of a Lake Wabamun Citizen’s Panel that became the Wabamun Watershed
Management Council, which provides input to watershed planning.

8.7.7 Aboriginal community concerns

Interaction between the Paul Band First Nation and provincial and/or federal government agencies was
hampered due to the lack of notification, which resulted in challenges with respect to communication and
collaboration. The First Nation filed a lawsuit over damage to the band’s land and water. A settlement
was reached with CN in 2008, resulting in a payment of $10 million (CBC 2008).

8.7.8 Important factors affecting the consequences of the Wabamun Lake spill

8.7.8.1 Oil properties and behaviour

 Wind dispersed the surface oil over a wide area and onto shorelines. Langmuir circulation may
have increased horizontal and vertical dispersion.
 The unique circumstances of the spill (hot oil flowing over land and picking up sediment before
entering the lake) meant that tar balls were formed with accumulated sediment and sank.
 The heavy fuel oil contained about six times more total PAHs than most crude oils and also
contained a higher proportion of alkyl PAHs than most crude oils (alkyl PAHs pose greater risk
of chronic toxicity).
 The tough, weathered coating on the tar balls sequestered the less weathered oil within the balls,
greatly limiting access to biodegradation and creating a longer-term source for release of less
weathered (and more toxic) oil in near-shore environments as the tar balls broke apart.

8.7.8.2 Effect of the environment on fate and behaviour

 Lake morphometry and mixing characteristics.


 The lake is shallow, wind-swept and well-mixed, which created conditions conducive to
rapid surface dispersion; and
 Shoreline characteristics that include small bays, creek mouths and marshy areas
increased the areas with a high potential for accumulation of oil.
 Seasonality.
 Winter conditions extended the time required for biodegradation and interrupted cleanup
activities;
 Less mixing energy under winter ice may have contributed to oil-related effects on fish
already stressed by natural ecological factors; and
 The behaviour of the residual oil in its various forms (e.g., tar balls) was influenced by
temperature.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 319
 Presence of existing anthropogenic stressors.
 The area closest to the oil spill was also affected by a thermal plume from a power plant,
physical removal of macrophytes (for aesthetic reasons) and high fishing pressure
(Schindler et al. 2004). Thus, fish inhabiting the area nearest to the spill were already
exposed to other stressors and may have been more susceptible to the additional stress of
the spill.

8.7.8.3 Oil toxicity

 Relatively few fish deaths were observed, commensurate with the temperature of the oil when it
entered the lake which was at summer temperatures, leading to rapid loss of low molecular
weight components.
 There were possible effects on benthic invertebrate community structure but confounding factors,
such as water depth and substrate particle size, may have been at least partially responsible for
observed results.
 Sublethal toxicity:
 Embryo deformities were observed in whitefish and pike associated with exposure to
PAHs (the heavy fuel oil had a higher relative concentration of alkylated PAHs than
typical Bunker C);
 There was no avoidance of near-shore oiled areas by fish. Fidelity to nursery, rearing and
feeding areas may have exposed the fish to oil in sediments and food organisms or
reduction in food abundance may have reduced fitness going into the winter season; and
 Metabolizing the PAHs may have created an energy cost that affected fitness and
sufficient concentrations of PAH metabolites and reactive oxygen species to have caused
oxidative stress (see Chapter 4).

8.7.8.4 Spill response

 Physical removal of oil caused direct damage to habitat from trampling, low-pressure flushing,
vacuuming and dragging over substrates and direct fish mortality due to being removed together
with cut vegetation.
 Cleanup activities recovered more than 95% of the oil components that were not lost through
natural weathering. The effects of physical removal were balanced against the degree of cleanup
required to re-establish human uses of the lake.
 Delays in response were related to lack of preparedness and response capabilities.

8.7.9 Lessons learned from the Wabamun Lake spill

 The consequences of spills of oil adjacent to lakes can be greatly reduced if spill response is rapid
and effective in preventing the oil from reaching the lake in the first place.
 Formal notification and involvement of Indigenous communities directly affected by oil spills is
necessary for provision of ground support and to avoid adversarial relationships during the
response.
 The exposure of spilled oil to soils and organic matter along the spill path increase the potential
for production of tar balls.
 Deposition of tar balls and tar mats into heavily vegetated shorelines greatly increases the
difficulty of cleanup and prolongs exposure of aquatic biota to physical smothering effects of oil,
as well as to toxic effects of the oil.
 Cleanup of shoreline areas resulted in reduced density of vegetation and a decrease in habitat
quality for several years.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 320
 An advance plan should be developed for decisions regarding when the effects of cleanup
activities outweigh the benefits.
 Wabamun Lake was one of the most well-studied lakes in Alberta, allowing a more confident
assessment of the effects of the spill relative to baseline conditions (including effects of existing
anthropogenic stressors).
 Crude oil spills into freshwater lakes can cause sublethal effects on fish which are observable in
the field.
 There was a significant impediment to an assessment of impacts following the spill due to the
priority given to enforcement activities and the control on collection and analysis of samples.

8.7.10 Research needs related to the Wabamun Lake spill

Recommendation: The Panel recommends a follow-up study of residual oils in reed beds and along
sandy shorelines of Wabamun Lake. Anecdotal reports in 2015 of tar balls and buried oil in sandy
beaches indicate the need for a study. This study would generate useful information, with respect to
the chemistry and persistence of the residual oil, the impact of cleanup operations and the longer-
term effects on ecosystem services, such as recreational use.

Recommendation: This spill represented a lost opportunity to learn about oil fate and effects in a
small ecosystem where there was a high potential to discriminate oil effects from other natural and
anthropogenic stressors. The Panel reiterates its earlier recommendation regarding identification of
‘spills of opportunity’.

8.8 Spill from a Ruptured Pipeline into Talmadge Creek and the Kalamazoo River, Michigan

On July 25, 2010, a rupture in a pipeline carrying diluted bitumen (dilbit) resulted in the release of ~3,200
m3 into Talmadge Creek and from there into the Kalamazoo River. It took over 17 hours for the rupture to
be confirmed and the flow of dilbit to be stopped. The spill consisted of 23% Western Canadian Select
and 77% Cold Lake Blend dilbit (Table 2.2). Western Canadian Select is a heavy blended unconventional
sour crude (CrudeMonitor.ca) composed of bitumen blended with sweet synthetic and condensate
diluents and 25 streams of conventional and unconventional Alberta heavy crude oils blended at the
Husky Terminal in Hardisty, AB. Cold Lake Blend is an asphaltic heavy crude blend of bitumen and
condensate.

The spill occurred during a period of high rainfall. Water-soaked soils allowed the dilbit to easily run
overland to Talmadge Creek and from there downstream to the Kalamazoo River. The flood flows caused
significant contamination of riparian lands. The flow conditions had an exceedance probability of 4%
with a mean velocity of about 1.1 m/s and a mean depth of 1.2 m (Fitzpatrick et al. 2015). About 3.2 km
of Talmadge Creek, 60 km of the Kalamazoo River and three impoundments involving medium to high
quality wetlands were affected (Noble 2012). The spilled dilbit ultimately reached a reservoir called
Morrow Lake about 65 km downstream. At that point, the high flows dropped, stranding oil on the
floodplain.

Local residents self-evacuated from 60 residences and about 320 people reported symptoms consistent
with crude oil exposure, including headache, nausea and respiratory symptoms (Stanbury et al. 2010;
EPA 2012; Fitzpatrick et al. 2015). Benzene (Table 2.1) was the primary public health concern for
residents and workers during the first 30 days.

About 2,900 m3 of the spilled oil was recovered during the first year and about 10% submerged (about
300 m3) (Fitzpatrick et al. 2015). The diluent of natural gas condensate volatilized. The weathered
bitumen was positively buoyant at room temperature but submerged when mixed with river sediment

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 321
under natural turbulent river conditions where it caused persistent globule and sheen releases (Fitzpatrick
2014).

8.8.1 Oil-particle interactions and submergence

The spill illustrated the importance of understanding the factors contributing to the formation and
submergence of oil particle aggregates (OPAs; Figure 2.7). Flood conditions at the time of the spill
increased turbulence and the presence of suspended particulate matter that, in turn, associated with oil
particles to form OPAs (Fitzpatrick et al. 2015). Additional mixing from flows over two dams may also
have played a role, although OPA and submerged oil accumulated in the first 5 km of river length,
between the spill source and the first dam.

Some additional features of the Kalamazoo River may have been important factors in OPA formation,
transport and deposition (Fitzpatrick et al. 2015). These features included:

 Abundant wetlands in the floodplain and flooding of riparian lands at the time of the spill. Thus,
suspended and bottom sediments had relatively high organic matter content;
 The river is wide and has an average gradient of 0.06% in the spill-affected reach;
 Channel margins, backwaters, side channels, oxbows and impoundments all provided
depositional areas; and
 Post-spill high-flow events caused re-suspension and resettling of OPA in downstream areas.

The USGS (Fitzpatrick et al. 2015) suggested that for impoundments with accumulations of thick fine-
grained sediment, the process of gas bubble formation (e.g., from microbial methane production in
anaerobic sediments) and release from sediments (ebullition) is likely to be an important mechanism for
re-suspending OPAs in the water column and releasing oil as sheen on the water surface. Spontaneous
releases of oil globules and floating OPAs have been observed regularly in the impounded sections of the
Kalamazoo River during 2011–14, resulting in oil sheens at the water surface.

Fitzpatrick et al (2015) noted that OPA formation is a natural process that enhances the physical
dispersion of oil and might result in enhanced biodegradation (Figure 2.3). The authors noted that the
formation of OPAs would reduce the bioavailability and toxicity of the residual oil to aquatic organisms.
They suggested that active enhancement of OPA production could be an alternative to the use of chemical
dispersants. However, the prescribed sinking of spilled oil or the use of sinking agents is currently
prohibited by the U.S. EPA because of the potential risks of acute and chronic toxic effects on benthic
organisms and possibly decreased biodegradation once the oil is deposited and buried in anaerobic
sediments (Fitzpatrick et al. 2015). These risks were illustrated in the case of the Wabamun Lake spill,
where the tar balls or tar mats formed from interaction with soil particles were resistant to biodegradation
and prone to releasing relatively unweathered oil from the interior of the tar balls. Tar balls and tar mats
also continue to come ashore in the Gulf of Mexico subsequent to the DWH spill.

Fitzpatrick et al. (2015) concluded that a major factor in ecological risk of OPAs is whether they are
physically diluted in suspension (less risk) or concentrated by deposition (more risk). The added context
of water depth and the geographic extent are also important contributors to risk.

8.8.2 Environmental effects

Limited acute toxicity testing was performed on oiled sediments from the Kalamazoo River. As part of a
NEBA, effects on aquatic organisms from weathered oil were assessed in laboratory acute toxicity studies
of seven sediment samples collected from oil-affected backwater habitats along the Kalamazoo River in
February 2012, about 19 months post-spill (Bejarano et al.2012 cited in Fitzpatrick et al. 2015). Ten-day
whole sediment toxicity tests used Chironomus dilutus and Hyalella azteca and included survival, growth

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 322
and biomass as the toxicity endpoints. Results from the toxicity tests indicated that C. dilutus were more
sensitive to oiled sediment (and presumably OPAs) than H. azteca but that all samples exceeded the
minimum survival (70%) and growth (0.48 mg ash-free dry weight at test termination) criteria for
acceptable controls for the C. dilutus tests.

Fitzpatrick et al. (2015) concluded that on the basis of the weight-of-evidence approach and additional
risk metrics, it is possible that residual oil at two heavily and one lightly oiled area may pose some risks
to benthic receptors. However, chronic toxicity of residual oil to sensitive life stages of benthic
invertebrate and fish species remained unknown at the time of writing (January 2015).

In the fall of 2010, all locations on Talmadge Creek and the Kalamazoo River showed signs of impact,
but it was unclear whether the impacts were from the spilled dilbit or from physical disturbance caused by
cleanup activities (see discussion of cleanup below). Benthic invertebrate abundance and diversity were
reduced in Talmadge Creek, and habitat disturbance was severe (particularly overbank areas) (Noble
2012). Abundance and diversity in the Kalamazoo River were reduced due to severe habitat disturbance,
sedimentation problems and bank erosion. By the summer of 2011 (13 months after the spill) benthic
invertebrate diversity increased in areas of Talmadge Creek altered by the cleanup. Sheen and oil were
present. The creek channel was more open, with more sunlight penetration due to extensive removal of
oiled vegetation. Abundance and diversity of benthic invertebrates had also improved in the Kalamazoo
River, but ongoing work continued to cause impacts. Increased sediment movement and deposition were
problematic, and sheen and oil were noted during surveys. In August 2012, a second round of response
work disturbed 3 km of the creek. However, according to Noble (2012) the restored creek provided much
better habitat compared to 2011.

No fish kills were observed although the Talmadge Creek fish community was reduced and habitat
greatly diminished in 2010 with some recovery in 2011 (Milsap et al. 2012). Some declines in fish
community diversity and abundance were observed at some sites in the Kalamazoo River.

Crushed and freshly dead mussels were found in the spill area but not in a reference area. Over 3,000
turtles, 170 birds and 38 mammals were brought to rehabilitation centres with survival rates of 97%, 84%
and 68%, respectively (Milsap et al. 2012). Turtles were affected the most because they burrow down into
sediments. The affected sections of the Kalamazoo River were closed to public access for nearly two
years.

8.8.3 Effects of cleanup activities

Talmadge Creek was reconstructed due to removal of bitumen deposits (Figure 8.10). Significant areas of
riparian vegetation were stripped from the banks of Talmadge Creek and riparian lands converted to dirt
roads.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 323
Figure 8.8 Response operations near the source of the spill on Talmadge Creek. Image from US EPA.

The persistent appearance of an oil sheen on the water was a cause for concern among the public, spurring
efforts to locate, dislodge and disperse submerged oil (the source of the sheens) (Figure 8.11; Chapter 6).
This activity employed long poles inserted into sediment to detect the submerged oil, followed by
agitation to release buoyant oil, which continued until there was no sheen (S. Hamilton pers. comm.).
Aggressive sediment agitation techniques (raking, flushing, aeration and skimming the river bottom
physically or with water jets) was conducted in 2011 to liberate submerged oil as recoverable sheen,
potentially contributing to further OPA formation and transport of OPAs to downstream reaches
(Fitzpatrick et al. 2015). This was illustrated by results before and after agitation at the Ceresco Dam
(Noble 2012).

Figure 8.11 Agitation of Kalamazoo River sediment with jets of water to flush submerged oil to the surface.
Image from US EPA.

The persistent residual submerged oil and oiled sediment in the Kalamazoo River resulted in a protracted
cleanup that ultimately required dredging and removal of several islands (because they had been over-
washed by oil) and has accounted for a major share of the cleanup costs, which to date have surpassed
$1.2 billion (Fitzpatrick et al. 2015).

An advisory group established to evaluate the risks and benefits of cleanup concluded that natural
biodegradation could not be relied upon in this case, primarily because heavy oil is chemically recalcitrant
to biodegradation and furthermore became buried in anaerobic sediment. The group noted that the
weathered submerged oil had low acute toxicity and suggested that the majority of the submerged oil
should be left alone. However, there were specific reservoir areas that were identified for further

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 324
dredging. The group noted that agitation should no longer be considered an option since less than 1% of
the oil was released and potentially further degraded the habitat (S. Hamilton, pers. comm.).

8.8.4 Evaluation of the spill response

The National Transportation Safety Board (NTSB) (2012) noted “pervasive organizational failures” in the
pipeline company that included:

 Deficient pipeline integrity management procedures, which allowed well-documented crack


defects in corroded areas to propagate until the pipeline experienced catastrophic failure;
 Inadequate training of control centre personnel, which allowed the rupture to remain undetected
for 17 hours and through two start-ups of the pipeline that increased pressures and the amount of
oil lost; and
 Insufficient public and agency awareness and education, which allowed the release to continue
for nearly 14 hours after the first notification of an odour to local emergency response agencies.

The NTSB (2012) stated that the following contributed to the severity of environmental consequences:

 The pipeline company’s failure to identify and ensure the availability of well-trained emergency
responders with sufficient response resources;
 The lack of regulatory guidance for pipeline facility response planning by the Pipeline and
Hazardous Materials Safety Administration (PHMSA); and
 PHMSA’s limited oversight of pipeline emergency preparedness that led to the approval of a
deficient facility response plan.

8.8.5 Important factors affecting the consequences of the Kalamazoo River spill

8.8.5.1 Oil properties and behaviour

 Rapid evaporation of the diluent occurred, creating a respiratory and combustion hazard.
 There was a consequent increase in density of the weathered dilbit.
 Biodegradation potential was limited due to the high proportions of HMW compounds, such as
resins and asphaltenes.
 Abundant wetlands with high organic matter contributed to OPA formation, followed by
submergence.

8.8.5.2 Effect of the environment on fate and behaviour

 River flow.
 Flood conditions at the time of the spill increased turbulence and the presence of
suspended particulate matter that, in turn, associated with oil particles to form OPAs, and
caused oiling of riparian lands and vegetation; and
 Post-spill high-flow events caused re-suspension and resettling of OPA in downstream
areas.
 River channel characteristics.
 Channel margins, backwaters, side channels, oxbows and impoundments provided
depositional areas.
 Sediment characteristics and processes.
 Gas bubble formation and release from fine-grained sediments in reservoirs (ebullition)
re-suspended OPAs, resulting in a sheen on the water surface; and
 Fine-grained sediment reduced oxygen replenishment and led to anaerobic conditions.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 325
8.8.5.3 Oil toxicity

 There were few apparent acute toxicity effects—no fish kills were noted immediately after the
spill and there was high survival among oiled turtles and birds.
 There were possible effects on benthic invertebrate community structure but effects of cleanup
activities may have played a greater role than oil toxicity; e.g. crushed and dead mussels.
 OPAs appeared to be low in toxicity (at least to two benthic invertebrates in laboratory toxicity
tests).
 There was no publicly available information on sublethal toxicity to fish.

8.8.5.4 Spill response

 Delays in response were related to lack of preparedness and response capabilities, as well as to
deficiencies in regulatory oversight.
 Physical removal of oil caused:
 Direct and major damage to habitat along Talmadge Creek. However, reclamation
appeared to have successfully produced viable habitat;
 Apparent direct mortality to mussels (via crushing) from cleanup activities; and
 Agitation to dislodge submerged oil and create recoverable sheen appeared to cause
greater impacts as it resulted in negligible oil recovery and most likely dispersed the oil
into deeper anaerobic layers of the sediment where it would become more persistent.

8.8.6 Lessons learned from the Kalamazoo Spill

 Spilled dilbit can form OPAs under conditions of turbulent flow with elevated suspended
sediments and abundant organic matter along the spill flow-path.
 Acute effects from spills of dilbit can be minimal given rapid evaporation of diluents.
 Chronic effects of residual bitumen can be difficult to distinguish from the effects of cleanup
activities; therefore, a careful evaluation of net environmental benefit is required.
 It may be best to rely on (slow) natural cleanup processes for submerged oil since the negative
effects of cleanup may be significant. However, the NEBA may also have to include net
sociological benefit because of concerns of the public.
 The costs of cleanup can be huge relative to demonstrated effects. Cleanup appeared to be driven
by concerns over submerged oil even though effects from the submerged oil were not particularly
apparent (apart from the aesthetic concerns raised by sheens).

8.8.7 Research needs related to the Kalamazoo spill

Recommendation: Investigation of OPA behaviour and effects on habitat quality and on benthic
invertebrates and fish would contribute valuable information to the overall evaluation of risks of
spills of dilbit. Various approaches could be used to represent Canadian receiving environments,
including a combination of laboratory and mesocosm experiments. Experimental conditions which
include ice are required for the Canadian context.

8.9 Other Recent Spills

8.9.1 Westridge 2007 Spill, Burnaby, British Columbia

On July 24, 2007, a backhoe operated by a third-party contractor accidentally ruptured the Trans
Mountain pipeline carrying crude oil to the Westridge Martine Terminal. Crude oil from the punctured
pipeline sprayed about 12 to 15 m into the air for about 25 minutes. Fifty homes and properties, as well as
a section of the Barnett Highway, were affected (TSB 2007). Approximately 234 m3 of heavy synthetic

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 326
crude oil blend was spilled. Approximately 40% of the oil entered the storm drain system and reached
Burrard Inlet through shoreline storm outfalls, a submerged storm outfall and Kask Creek. Once in
Burrard Inlet, the oil began to spread further into the inlet through wind and tide action (TSB 2007).

Western Canada Marine Response Corporation (WCMRC) responded to the spill within an hour and
began booming marine areas within 45 minutes of oil being discharged through the shoreline storm water
outfall (Stantec 2012).

Emergency phase cleanup techniques included:

 Booming to contain oil around the release points and also exclusion booming to protect sensitive
shorelines; and
 Skimmers and absorbent pads to remove oil.

According to Kinder Morgan (2014), approximately 95% (210 m3) of the released oil was recovered. An
estimated 5.5 m3 was not recovered and was considered to be released to the marine environment (Kinder
Morgan 2014).

SCAT was used to identify oiled shorelines, establish cleanup methods and set priorities. Approximately
1,200 m of shoreline were affected by the spill (TSB 2007). Recovery and rehabilitation of affected
wildlife was performed through daily surveys for six weeks following the spill (Stantec 2012). By 2011,
most of the recovery endpoints established for the spill had been met (water quality in 2007; intertidal
sediment quality, PAH levels in crab and intertidal community structure in 2011) (Stantec 2012). In 2012,
PAH levels in mussels had not yet met the endpoint. However, results were considered to be confounded
due to other sources of PAH in the area (e.g., urban runoff and vessel traffic) (Stantec 2012).

Favourable environmental factors at the time of the spill included:

 Sunny weather (no rainfall runoff to increase movement of oil in storm drains, good evaporation
conditions);
 Slack tide conditions which helped keep the oil near the shore while booms were placed;
 Timing which was of primary migration and overwintering period for birds and after the breeding
bird season; and
 Timing was prior to the main salmon migration period (Stantec 2012).

8.9.1.1 Lessons learned

The Westridge Pipeline spill was a case where spill response was appropriate and successful.

The TSB (2007) determined the factors contributing to the third-party pipeline breach. These factors are
listed below:

 The field location of the Westridge Pipeline was not accurately indicated on design drawings, and
the location was not verified, as required under the NEB Pipeline Crossing Regulations, Part I.
Therefore, the actual field location was not discovered before the start of construction;
 Inadequate communication within Kinder Morgan Canada, Inc., and between Kinder Morgan, the
consultant and the contractor;
 The conditions of the crossing agreement and the NEB Pipeline Crossing Regulations Parts I and
II respecting an onsite pre-construction meeting, locating the pipeline and supervision of
construction activities were not adhered to, thus compromising the safe operation of the pipeline;
and

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 327
 The initial decision by Kinder Morgan to shut down delivery pumps without isolating the gravity
feed from the terminal instead of continuing with drain-down of the Westridge Dock Transfer
Line increased the volume of crude oil released and was not in conformity with standard
emergency shutdown procedures (TSB 2007).

A review of the incident by Kinder Morgan led to the implementation of a Pipeline Protection Department
with responsibility for:

 Public awareness;
 Pipeline and associated facilities markings;
 Permits for safe work around pipeline and associated facilities;
 Aerial and ground patrols; and
 Response to emergency calls (Kinder Morgan 2014).

The WCMRC noted that there would be a clear benefit of having an on-water facility linked closely to
personnel access and stated that it is reviewing different options within Burrard Inlet to achieve that end
(Kinder Morgan 2014).

8.10 Release of Fuel Oil from the Marathassa Grain Carrier into English Bay, British Columbia

On April 8, 2015, about 2.7 m3 of intermediate fuel oil (suspected to be IFO 380) was leaked from the
Marathassa bulk grain carrier into English Bay, Vancouver (Figure 8.12) (CCG 2015a). This resulted in
highly visible consequences due to the multi-use nature of the English Bay area (anchorage for Port Metro
Vancouver tanker traffic, tourism, recreation). For example, the spill formed a visible sheen and fouled
beaches around English Bay, the North Shore, Stanley Park and up into Burrard Inlet, resulting in closing
of beaches by Vancouver Coastal Health. One week after the spill, the DFO closed recreational fisheries
in a section of the affected area after advice from Vancouver Coastal Health. The Musqueam First Nation
issued its own urgent notice one day after the spill, warning those who harvest crab and prawn in the area
to stop fishing. However, subsequent chemical analyses of water samples showed hydrocarbon levels
below laboratory detection limits, including in waters surrounding the ship, in line with this very small
and localized spill.

Figure 8.12 Booms placed around the Marathassa. Image from CCG (2015a) (http://www.ccg-
gcc.gc.ca/independent-review-Marathassa-oil-spill-ER-operation)

An operational update by the Canadian Coast Guard on April 23, 2015 (CCG 2015b) reported that
shoreline cleanup and assessment teams had completed work, with no observed oil reported. Beaches
were re-opened at all affected beaches, in conjunction with Vancouver Coastal Health. Sediment and
mussel sampling was reported to be ongoing and the precautionary recreational fisheries closure remained
in effect. An independent review of the spill response provided to the Canadian Coast Guard in July,
2015, noted that impact on the public was minimal from a health and safety perspective. However,

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 328
Environment Canada estimated that approximately 20 birds were affected. As of the date of this writing,
monitoring continues.

8.10.1 Lessons learned

An independent review of the spill response (CCG 2015a) noted a number of areas for improvement in
spill response:

 A need for improved communication with partners to ensure accuracy of communications;


 The Canadian Coast Guard did not have sufficient nearby initial capacity to respond due to
demobilizing from another pollution response. Therefore, the Canadian Coast Guard contracted
the WCMRC to initiate the on-water response and provide support to the Incident Command
Post;
 Information sharing on a common network was not optimal due to federal government electronic
policies and protocols;
 Canadian Coast Guard had not yet reached full operational capacity for its Incident Command
system. Therefore, it took several days for Unified Command to achieve an operational rhythm;
 Early alerting of municipalities, First Nations and stakeholders was delayed due to the low
classification of the incident in the provincial alerting system;
 Vancouver Area Emergency Response Planning timelines did not align with the immediate need
to engage partners in the development of an efficient and effective plan in Vancouver harbour;
 The lack of a physical presence of Environment Canada impacted the effectiveness and efficiency
of the Environmental Unit; and
 Public communications from Unified Command were challenging because energy was focused on
supporting government officials in media briefings, rather than ensuring key facts were being
shared with citizens and Unified Command partners.

The independent review presented 25 recommendations for consideration by the Canadian Coast Guard
and partners. The recommendations covered communication, training, standards, adequacy of staffing,
systems, protocols, tools, approval processes and need for involvement of other federal agencies, notably
Environment Canada (CCG 2015a).

In addition to the findings of the independent review, a representative of the Vancouver Aquarium (one of
the groups that assisted in the response) pointed out that:

 Gaps exist in research and preparedness because of the lack of cohesive long-term monitoring of
coastal ecosystems, creating a lack of baseline data which, in turn, makes it difficult to assess
spill impacts;
 There is no official clarity around who is to monitor the effects of a spill; and
 Coordination among agencies and organizations was lacking, resulting in duplication of effort
(e.g. at least three agencies collected water samples for testing) and/or gaps in information (Kane
2015).

Recommendation: The Panel trusts that the recommendations made to the Canadian Coast Guard
by the independent reviewer arising out of the Marathassa spill will be fully implemented. The
Panel notes that deficiencies in communication and coordination are common themes among the
Canadian case studies reviewed in this chapter.

Recommendation: The Panel reiterates its recommendation for standard guidance regarding
monitoring after oil spill events, as well as a targeted program of baseline data collection focused on
areas identified as high-risk.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 329
8.11 Release of Bitumen Emulsion from a Pipeline in Alberta

On July 15, 2015, Nexen Energy reported a pipeline failure at its Long Lake facility near Fort McMurray,
AB, that released an estimated 5,000 m3 of an oil emulsion consisting of bitumen, process water and sand.
The release affected an area approximately 40 m by 400 m and was confined primarily to the pipeline
right of way, which includes muskeg (AER 2015a). A water body near the release site had apparently not
been impacted. The spill created concern because of its size, and the fact that it involved the rupture of a
new, double-walled pipeline that had recently been inspected. There were also reports of at least one
waterfowl death.

Subsequent to the release and investigation by the Alberta Energy Regulator (AER), it issued a
Suspension Order to Nexen due to what was referred to as “noncompliant activities at Long Lake oil
sands operations pertaining to pipeline maintenance and monitoring” (AER 2015b). The order directed
the company to: immediately suspend 15 pipeline licenses, which required shutdown of 95 pipelines
carrying natural gas, crude oil, salt water, fresh water and emulsion; and provide sufficient documentation
to assure the AER that the pipelines can be operated safely. As of this writing (September 2015), the
suspension has been partially lifted. The cause of the failure is under investigation.

8.11.1 Lessons learned

This spill illustrates that the technology incorporated into newer pipelines is not sufficient to eliminate
failures, even of monitored double-walled pipelines. New pipelines must be managed, maintained and
monitored, and their performance must be reported rigorously and transparently.

This case also illustrates that the AER apparently was not aware of Nexen’s lack of compliance regarding
maintenance and monitoring until after the spill occurred. Regulatory oversight is required at all times,
not just after a spill event.

Recommendation: Systematic and efficient regulatory oversight of industry inspection, testing and
maintenance is required for all pipelines, old and new.

8.12 Overall Conclusions Arising from the Case Studies

Delayed response was a critical factor affecting the consequences of all of the oil spill case studies.
In some cases, adverse weather was a factor, which delayed initial response. Remote locations made
quick response difficult. Ineffective well-capping technology resulted in a flow of oil from the DWH
blowout, which lasted three months (although flow was greatly reduced two months after the blowout).
However, notwithstanding the importance of weather, remote locations and technological challenges,
human error (at an individual and organizational level) was a dominant factor across all case studies.
Absent or inadequate planning, insufficient integration, inadequate training, poor communication,
insufficient capacity (personnel and equipment), poor or no information sharing and lapses in regulatory
oversight were noted for most, if not all, spill case studies. Despite the improvement in engineering
design and monitoring technology, spills continue to go unnoticed and unreported for unacceptable
lengths of time (e.g., the recent Nexen pipeline spill).

The case studies illustrate that the consequences of oil spills cannot be predicted simply on the basis of
the type of crude oil. As discussed in Chapters 2, 3 and 4, the interactions among oil chemistry and the
receiving environment produce a wide range of potential exposures to aquatic organisms, with an
associated wide range of toxicity to individuals and effects on populations. Thus, the effects of diluted
bitumen spills will not always be more severe than spills of conventional crude oil. For example, the spill
of light crude oil into the Pine River, BC, had more severe consequences (particularly with respect to
acute mortality) than the spill of diluted bitumen into the Kalamazoo River.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 330
Studies of the medium and long-term effects of spills into fresh water are rare to non-existent. The
consequences of crude oil spills to freshwater systems in Canada can be substantial because of less
dilution potential and because of the tendency for oil to quickly become stranded along shorelines where
it can remain for long periods of time. It cannot be assumed that spills in fresh water cause fewer adverse
effects than spills in marine systems.

Relevant baseline data are often limited or completely lacking, thus hampering rapid decision-
making regarding the most appropriate spill response and cleanup measures and limiting the
ability to design an appropriate monitoring program with sufficient statistical rigour to reliably
indicate effects. Data should include key species at risk by region and type of watercourse (marine
offshore, marine inshore, estuaries, rivers, lakes, wetlands). Studies should include analysis oil
biodegradation potential to assist in the prediction of environmental persistence as well as the
development of remedial technologies.

There is often inadequate knowledge of the natural variability of indicators used in post-spill monitoring.
Therefore, achieving a rigorous statistical sampling design capable of establishing whether there are
significant effects from an oil spill is challenging. The effects of confounding anthropogenic factors are
also often not well understood (e.g., other point and non-point hydrocarbon sources, the presence of other
physical and chemical stressors, invasive species, commercial or recreational fisheries and habitat
degradation caused by human activities). The influences of large-scale phenomena, such as the Pacific
Decadal Oscillation and climate change, on biological indicators must also be considered.

SL Ross (2014) provided the following suggested requirements for baseline information in support of
monitoring effects on fish:

 Species of potential interest in the area threatened by the spill (e.g., sentinel species, most
abundant species, protected species, exploited species, i.e., subsistence, commercial,
recreational);
 Abundance and spatial distribution of key species within and beyond the areas threatened by oil;
 Location of nearby areas that could serve as uncontaminated control areas for comparative
purposes (e.g., similar habitat, same species, age distribution, conditions, physical conditions);
and
 Information concerning baseline levels of biomarkers in key species prior to the spill (e.g., tissue
levels of PAHs, EROD levels, information on spatial or seasonal variation in biomarker
parameters).

In addition to the above list, there is a need for baseline information on measurements that allow
assessment of effects on productivity in aquatic systems, particularly fish productivity. For example, data
on fish age and size distribution and recruitment of young would provide the basis for evaluating whether
a spill had the potential to alter the population via partial elimination of a year class.

Priorities for the acquisition of baseline data can be based upon the identification of locations that
are under the greatest risk of contamination from oil spills. The results of existing national relative
risk assessments (such as those reviewed in the following section) can be used for prioritization. A review
of locations with a combination of characteristics that increase risk should be conducted. These
characteristics could include increased likelihood of exposure to oil spills due to proximity to multiple
modes of transport of crude oil (e.g. rivers with pipeline crossings, railways and/or roads running parallel
to the river, and oil tanker traffic in estuarine areas). Other characteristics could include relative
sensitivity of the receiving environment determined by the presence of critical habitat and/or sensitive
species, and the presence of culturally important species. Professional judgement will also be needed for
this prioritization. Furthermore, it may be more appropriate to focus on indicators, such as benthic
invertebrates, for standard monitoring purposes.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 331
Aquatic ecosystem resilience to shocks from oil spills is key to the maintenance of ecosystem
services. The presence of long-term stresses may tax the capacity of ecosystems to be resilient to shorter-
term shocks. In some cases, ecosystems will undergo fundamental shifts in structure and function
following disturbances (e.g., changes in trophic structure caused by invasive species). If policies and
decisions are to be based upon resilience and the preservation of ecosystem services, multi-disciplinary
research effort on the resilience of key marine and freshwater ecosystems will be required.

Effects of oil spill cleanup on aquatic ecosystems can be significant. This is particularly true for
cleanup of stranded oil in both marine and freshwater environments, whether it is in the intertidal zone,
the shorelines of lakes and rivers, or on woody debris and man-made structures in river channels. A
careful, but rapid, NEBA is required tailored to each unique spill situation. Unfortunately, there is
insufficient knowledge of the long-term consequences of residual stranded oil to allow for confident
NEBA in many environments, including Arctic and sub-Arctic systems.

The potential for long-term effects on animal populations continues to generate much debate and
controversy. The Panel recommends that there be support for long-term research into effects of different
oil types on populations of aquatic biota, especially fish, marine mammals and waterfowl. This research
could be part of the suite of studies associated with significant oil spills.

There is an urgent need for the development and production of innovative, cost-effective and
readily available spill prevention and response measures for use in Canada, including the Arctic. As
already stated in this report, if spills do occur, maximum advantage should be derived from the
opportunity to study the fate, behaviour and effects of the spill in the short, medium and long-term.
Studies of the relative effectiveness of response measures should also be part of a suite of investigations
associated with all significant oil spills in Canada. Pre-approved funding should be put aside for ‘spills of
opportunity’ that would incorporate a combination of research and monitoring. Responsibility for each
component of spill studies must be clear and research and monitoring efforts must be well coordinated to
eliminate redundancy and to ensure that all required study components are implemented.

The consensus of scientists in oil spill countermeasure research is that major advances in the
development and validation of spill response technologies are being hampered by our inability to
conduct controlled field experiments with oil. In addition, validation is hampered by crude or no
methods for measuring the efficacy of response measures. While ‘spill of opportunity’ case studies will
provide lessons learned, they do not provide an optimal platform for the delivery of science due to
logistical constraints (e.g., lack of experimental replication due to site heterogeneity, etc.). To support
predictive numerical models and operational guidelines for spill response, there is a need for a rigorous
database for the fate, behaviour and effects of various types of oil spilled and the efficacy of current and
emerging oil spill countermeasures over a range of environmental conditions. This will require support
for the conduct of field trials designed with statistical rigour that incorporate controlled releases of oil.

National guidance for monitoring of oil spills is urgently required to ensure that information
gathered is reliable, adequate, credible and consistent. The guidance should include provisions for
adjustment in response to specific characteristics of the receiving environment. The guidance should also
include requirements for standard baseline datasets (tailored to specific receiving environments). The
guidance could be divided into two parts: 1) which information to collect without exception; and 2) which
information gathering can be deferred until the full scope of the spill and its potential effects are better
understood. The guidance should include information on data quality requirements, including
determination of minimum sample sizes, standard sampling protocols and laboratory quality
assurance/quality control.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 332
SL Ross (2014) also provided a list of recommendations for specific areas of research and development
that can improve the ability to mount monitoring efforts efficiently for spills in Canada. The list is as
follows:

 For potential Canadian biomonitoring species (cod, Arctic cod, snow crab, salmon, flatfish,
mussels, oysters) prepare an annotated bibliography of studies on effects and relationships among
exposure, bioaccumulation of and effects of PAHs (e.g., physiological biomarkers). The Panel
adds specific mention of freshwater species, such as whitefish, pike or small-bodied fish species,
used in programs such as the national Environmental Effects Monitoring program.
 For representative species used for monitoring, determine the exposure-response-recovery
relationships for biomarker activity and occurrence of histopathological changes. The Panel
suggests that the relevance of physiological biomarkers to exposure versus effects should be
examined (see Chapter 4).
 Identify and acquire or develop a model for estimating fate, behaviour and persistence of spills
into rivers. The Panel notes that several hydrodynamic models are already available that are in
common use in Canada. However, very few are three-dimensional and there appear to be no
widely accepted models to describe surface water-groundwater interactions across a wide array
of river types.

8.13 Prediction of Risk of Crude Oil Spills in Canada

8.13.1 Risk assessment for marine spills in Canadian waters south of 60th parallel

A risk assessment of marine spills along four sectors of the Canadian coastline - the Atlantic and Pacific
Coasts, the Estuary/Gulf of St. Lawrence and the Great Lakes/St. Lawrence Seaway System - was
conducted for Transport Canada and reported in 2014 (WSP and SL Ross 2014a), hereafter referred to as
the TCRA (Transport Canada Risk Assessment). A brief description of the approach and key findings of
this assessment is provided in Chapter 7. In this chapter, uncertainties associated with that risk assessment
are identified and discussed in detail because of its potential influence on future policy, spill preparedness
and response activities.

8.13.2 Risk assessment methods

8.13.2.1 Spill frequency

The TCRA considered production (offshore wells) and transportation sources and divided spills into four
sizes: 10 - 99.9 m3; 100 - 999.99 m3; 1000 - 9999.9 m3 and > 10,000 m3.

Each of the four sectors was divided into subsectors that were further subdivided into three zones
representative of near-shore, intermediate and deep sea environments. A total of 77 zones were allocated
a frequency of spill and an environmental sensitivity, which were then applied to generate a risk estimate
(WSP and SL Ross 2014a).

Estimated spill frequency was based on the past 10 years of spill data from Canadian and international
data sources. Spill frequency was estimated for crude, refined cargo and fuel according to the relative
volumes transported across each subsector. The Panel focused on the results for crude oil.

Uncertainty Regarding Spill Frequency

 The TCRA noted that while the variability in spill rates is not wide among different governance
structures around the world except perhaps at the highest end of spill size ranges, the use of
worldwide incident data may have overestimated the likelihood of spills in Canada given the
robust marine governance regime and the actual spill record. This assumption is subject to debate

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 333
based on evidence such as the global review of environmental damage liability regimes by
Goldsmith et al. (2014). Furthermore, despite advances in technology and improved safety
protocols, accidental releases, such as the recent Marathassa spill, will continue to occur (Section
8.12). The effects of increases in the volume of tanker traffic on spill rates in the various sectors
and subsectors were not discussed in the report.
 Probability estimates by the TCRA for the St. Lawrence Seaway were identified to be
conservative as they were not corrected to account for closure of the Seaway about 30% of the
year due to ice cover. However, their analysis did not include US traffic in the Great Lakes,
which would have provided a similar precautionary approach for evaluating the risk of spills in
the Great Lakes.

8.13.2.2 Environmental sensitivity

The TCRA estimated environmental sensitivity of each zone using three indicators: physical sensitivity
(PSI), biological resource (BRI), and human-use resource (HRI). The environmental sensitivity index
(ESI) was calculated as follows: ESI = 0.3(PSI) + 0.5(BRI) + 0.2(HRI). The weights applied to each
indicator were from a review of costs, such as the influence of physical factors in determining cleanup
costs.

The PSI rank was related to zone sensitivity, natural oil persistence and ease of cleanup.

 Exposed rocky shores were ranked least sensitive;


 Most sensitive were sheltered rocky shores, tidal flats, marshes, wetlands, eelgrass and ice
infested waters; and
 Gravel was ranked in the middle.

A 0.3 PSI for open sea was based on average cost of cleanup per tonne compared to near-shore
environments The BRI and HRI were based on a method developed for Australia (DNV 2011 cited in
WSP and SL Ross 2014a). It was assumed that the BRI of intermediate and deep sea zones that did not
include any of the identified biological components (e.g., species at risk, birds, mammals, reptiles, fish,
meroplankton and invertebrates) would be related to that of the nearest near-shore zone times a weighting
factor of 0.4x for intermediate and 0.1x for deep sea. Weights attributed to each biological component
included data from the Strategic Environmental Assessment for the Gulf of St Lawrence study
(GENIVAR 2013 cited in WSP and SL Ross 2014a) and from other marine environmental studies
(dredging, harbour, etc.).

The HRI was based on a combination of commercial fishing intensity, tourism employment intensity,
freshwater use intensity and freight tonnage index. Weights applied to each component were: 0.55 for
commercial fishing, 0.2 for tourism, 0.15 for water usage and 0.1 for port industry. Commercial fishing
intensity scores did not include recreational or traditional fishing due to absence of comparable data, but
the authors maintained that the index would provide a sufficient indicator of overall vulnerability.
Tourism Employment Intensity was based on the ratio of tourism industry employment versus total
employment. The Freight Tonnage Index was based upon total tonnage in each port. The use of fresh
water by humans (Freshwater Use Intensity) was based upon coastal population as a proxy because data
on water intakes were not available.

Recommendation: Future relative risk assessments should include explicit consideration of


intermediate and deep sea zones with greater sensitivity, particularly in zones with current or
projected offshore oil production.

The TCRA report noted that the environmental sensitivity values were generally lower for intermediate
and deep sea zones. They noted exceptions due to the availability of data for some particular, sensitive

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 334
intermediate and deep sea zones. However, there was no attempt to include these areas in the analysis (for
example, as exception layers). A discounting term may be appropriate in order to account for increased
capacity for dilution, dispersion and biodegradation in some of the zones or subzones.

Recommendation: Future relative risk assessments should build upon the TCRA results by
focussing on high-sensitivity areas within each of the assessment zones in order that preparedness
and response plans can include explicit plans for the areas with the highest potential consequences.

The TCRA report presented ESIs as average values for an entire zone. The authors noted the
shortcomings of this approach as the use of averaged numbers results in loss of detail—high sensitivity
areas within a zone might be concealed if surrounded by relatively low-sensitivity areas due to the
subsectors being so large.

Recommendation: Consideration should be given to the refinement of the PSI values based on
recent shoreline mapping data (an ongoing initiative of Environment Canada) and site-specific data
from past spills.

Recommendation: The methods for development of BRIs require review and refinement in order to
increase the reliability of the relative risk indices at the subzone level. These refinements should
reflect the nature of the receiving environment within subzones and the specific valued components
within the subzones.

Recommendation: The TCRA should be updated to include a fresh water-specific set of ESIs. Such
an assessment would be relevant not only to ship-source spills, but to spills from pipelines, crude-
by-rail and truck transport.

The reliability of the BRIs depended on the quantity and quality of data on the components (e.g., birds,
mammals, fish, etc.) in terms of where they are located within the zones plus the weighting for sensitivity.
Data for some components in some of the subzones were limited or non-existent. The information
presented in the TCRA on chronic effects provides evidence that raises questions about the weighting of
sensitivity; e.g., Table 4.1 in the TCRA. For example, fish are given a medium-low sensitivity despite the
sensitivity of early life stages to PAHs, as demonstrated in laboratory as well as field studies (Chapter 4).
However, evidence from spills around the world seems to indicate that fish populations are not as
sensitive as other marine biota. The cumulative impacts of stressors that can exacerbate the effects of
spills (e.g. increasing water temperature and/or nutrient loads) were not considered under the BRI. Within
the TCRA report there seems to be marine bias in the weighting approach used for sensitivity (probably
due to the larger database available); thus, the applicability to fresh water (i.e., the St. Lawrence River and
Great Lakes) is questionable.

Recommendation: Future relative risk assessments should include refined human use resource
indices, which include the uses of waterbodies for traditional purposes by Indigenous peoples, and
which include the presence of critical municipal, agricultural and industrial uses that may be
adversely affected by a spill. The Panel notes that the TCRA report includes a similar
recommendation.

The appropriateness and completeness of the HRI is subject to criticism since the inputs to the index, and
the weightings applied to those inputs, would be expected to vary significantly within zones and
subzones. For example, the presence of critical water intakes for agriculture in areas with lower human
population size along the St. Lawrence River might justify an increased weight for water use intensity in
those areas. In addition, the HRI did not incorporate the significance of traditional resource uses by
Indigenous peoples.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 335
8.13.3 Results of the risk assessment

The results of the TCRA are described in Chapter 7, whereas this chapter focuses on key parameters
determining the outcome of the risk assessment, followed by a review and discussion of uncertainties
associated with the TCRA results.

WSP and SL Ross (2014a) identified the key parameters influencing the relative risk outcomes as:

 Oil persistence, a consequence of chemical composition and microbiological sensitivity or


recalcitrance.
 Extent of the spill over time.
 Potential movement of the spill.
 Characteristics of the receiving environment.
 Shoreline habitat, sediment type, topography, currents, hydrology;
 Presence and concentrations of factors that would enhance natural attenuation or
biodegradation, such as nutrients, electron acceptors, etc.;
 Coastal resources in area of influence of spill;
 Physiological and behavioural characteristics of coastal resources; and
 Type and intensity of human activities.
 Type and volume of the spill.
 Toxicity, bioaccumulation rates;
 Volume and exposure concentrations in various media (water, sediments, soil, food, air);
 Direct versus indirect exposure via food;
 Exposure duration; and
 Time of year—organism life cycles plus weathering.
 Type and effectiveness of the cleanup response.
 Possibility of damage from cleanup activities.

The TCRA authors stated that effects on natural habitats in fresh water will resemble those of marine
spills. However, they also stated that spills in fresh water have a much greater potential for contaminating
potable water supplies, affecting areas of concentrated populations, damaging manmade structures and
disrupting other human activities such as recreation.

The TCRA authors also noted that, in most habitats, site recovery from an oil spill will occur within 2-10
years, and cautioned that perception of contamination can last longer than actual contamination, with
economic effects on fisheries and tourism sectors. The perennial and thorny question “How clean is
clean?” has been noted in Chapters 6 and 7.

It cannot be assumed that the TCRA risk assessments are uniformly precautionary. The TCRA assessment
is conservative in that it did not assume any marine oil spill prevention measures or responses, such as
mechanical recovery or dispersant application. However, the overall level of conservatism in the
assessment is difficult to determine because components within the assessment, such as environmental
sensitivity, may not have been uniformly conservative.

Recommendation: Preparedness planning should consider information from Indigenous,


provincial, territorial and municipal sources in order that smaller-scale, higher-risk locations can
be identified. For example, the Strait of Georgia and the Gulf of St. Lawrence have areas of
concentrated seabird and migratory bird use that would be a high priority for specific response
measures targeted at reducing impacts to birds. Many of these areas have already been mapped at
a finer scale than presented in the TCRA. Provincial laws and regulations must be considered in
preparedness to ensure harmonization and avoidance of redundancy.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 336
Not including data from provincial, territorial, and municipal sources is understandable for such a high-
level assessment; but it produced significant limitations in the risk analysis.

Except for the St Lawrence River, the assessment did not include rivers such as the Mackenzie River and
Fraser River, which have barge traffic and could have increased traffic in the future.

Recommendation: First Nation, Inuit, and Métis jurisdictions must also be considered and included
in preparedness planning and broad consultation with Indigenous peoples should be part of such
planning.

Any and all future preparedness planning should explicitly include First Nations, Inuit, and Métis
jurisdictions as well as broad consultation with all Indigenous peoples. Indigenous peoples are often the
most knowledgeable groups regarding local and regional conditions which would affect preparedness
planning. Indigenous peoples are also often the closest people to spills in remote locations; therefore, they
can be well-placed to be among the first responders (if capacity exists in Indigenous communities).

Recommendation: Any future relative risk assessments should include weather as a factor within
the risk scenarios.

It was logical that local or regional-scale weather data were not included in the TCRA high-level
assessment. However, consideration of the relative frequency of extreme weather events among the zones
used in the assessment would have added value since the weather can be an important determinant of
consequences (e.g., due to delays in response, weather effects on efficacy of response measures and
weather effects on dispersion of the oil).

Recommendation: The TCRA should be revised to include risks of spills from U.S.-bound tanker
traffic in Canadian waters.

Risk in the Pacific sector was related largely to tankers going to refineries in Washington State, and the
authors stated that, therefore, these spills would not be subject to Canada’s spill response regime, despite
close proximity to southern British Columbia’s coastal areas. There is also concern that the risks of spills
from U.S.-bound tanker traffic within Canadian waters was not considered.

Recommendation: The Panel has made several recommendations with respect to research on the
fate and effects of diluted bitumens. Results of this research should be included in future risk
assessments, particularly for scenarios where research indicates that the fate or toxicity of diluted
bitumens would differ significantly from conventional crude oil.

The modeling of the behaviour of spills did not distinguish diluted bitumens from conventional crude, and
it was not clear whether critical differences between marine and freshwater systems with respect to the
processes were included in the model (drifting, spreading, evaporation, photooxidation, natural
dispersion-dissolution of oil in water, water-in-oil emulsification, biodegradation and sedimentation;
Chapter 2).

Recommendation: The current lack of inland shoreline sensitivity information in the open
literature is a serious gap affecting the ability to apply risk-based preparedness and response
planning. Prioritization of areas requiring mapping could be based upon evaluation of the intensity
of human use, the current knowledge of the relative sensitivity of ecosystems, and the availability of
information from various sources. For example, high-priority areas for mapping might include
areas upstream and downstream of hydroelectric dams.

Shoreline mapping is still incomplete, particularly inland, including along the Great Lakes. The Panel is
aware that industry is doing some mapping and that Environment Canada continues to work on mapping

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 337
marine coastal areas. The Panel also understands that Environment Canada has been working with
Alberta, British Columbia, and CN Railway on mapping of rail corridors and that similar work was done
in Quebec (S. LeBlanc, pers. comm). Environment Canada has been approached by Canadian Pacific
regarding conducting similar work along their railway corridors (S. LeBlanc, pers. comm.).

Recommendation: National-scale relative risk assessments used for preparedness and response
planning must be based upon reliable and consistent data and analysis. Therefore, standardized
methodology for environmental sensitivity mapping is an urgent requirement.

Apparently, even when environmental sensitivity data are available, they are obtained using different
methods. For example, according to the TCRA, survey methods used by each regional department of the
DFO may differ, and bird data are obtained using different types of datasets, resulting in regional
disparities.

Recommendation: Future refinements to relative risk assessment must include explicit modeling of
freshwater systems, taking into account the unique features of freshwater environments that affect
exposure and effects.

The TCRA assumption that effects in fresh water will resemble marine spills is problematic, as illustrated
by the case studies reviewed earlier in this chapter and the review in Chapter 4.

Recommendation: The Panel recommends that the relative risk assessment method could be
applied at a smaller scale for the purposes of identifying the critical areas for protection and
preparedness, and could be used to identify critical data gaps.

As the authors state, rare events are difficult to predict—especially the really large spills. Therefore,
preparedness planning should focus on events where consequences would be highest.

Recommendation: The Panel strongly recommends the development of a national, consensus-based


set of indicator species for each of Canada’s high-risk major marine offshore and inshore zones.
Once these indicator species have been selected, coordinated research programs among academia,
industry and government should be developed that: a) compile existing information; b) identify and
prioritize critical data gaps; and c) conduct research to fill the priority data gaps. (See
recommendations in the Section 8.15 for additional details.)

Recommendation: The Panel recommends that future national or regional-scale relative risk
assessments should include indices of current status of receiving environments, with emphasis on
existing levels of anthropogenic stressors. Future risk assessment could also combine risks of tanker
spills with risks of pipeline, rail and road spills for a cumulative oil spill risk in the most sensitive
receiving environments.

The overall TCRA environmental risk index appeared not to address cumulative risks or the existing state
of the environment (which may already be degraded and at increased risk; reviewed in Chapter 4).

8.13.4 Major Conclusions and Recommendations from the Transport Canada Risk Assessment
Report

Some of the key conclusions and recommendations presented in the TCRA report were:

 Spills in the 10 to 1000 m3 range generate higher overall risks both respect to frequency and
consequence;
 Eight of 10 highest risk areas are near-shore and two of the 10 are intermediate—this conclusion
has implications for response capabilities and preparedness;

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 338
 The Marine Pollution Incident Reporting System database on pollution incidents maintained by
the Canadian Coast Guard should be examined for improvements regarding comprehensiveness
and quality of data;
 In future risk assessments, consideration could be given to effects of spill response according to
oil type and spill size, as well as remoteness, weather and other factors;
 All regions should attain the same level of detail regarding environmental data, and these data
should be in GIS (geographic information system) format;
 Regarding physical sensitivity, more information on evolution of coastal areas facing climate
change impacts, in particular for sensitive components such as coastal wetlands, is urgently
required. Integrating data on littoral geomorphology to capture erosion impacts would be a
valuable refinement– these data are often available at the provincial level;
 Regarding biological resources, data should be incorporated for all protected marine areas under
various jurisdictions and Ecologically and Biologically Significant Area (EBSA) or equivalent
layers, and risk analysis for fresh water should be considered separately from marine analyses due
to differences in the type of data available; and
 Regarding human resource use, information should be incorporated for provincial and municipal
conservation/protected areas, for freshwater intake and utilization (drinking water, agricultural
and industrial) and for archeological and cultural heritage sites. The current risk estimates should
be overlaid with information on Aboriginal communities.

The Panel supports the recommendations of the TCRA report.

8.14 Risk Assessment for Marine Spills in Canadian Waters North of 60th Parallel

A relative risk assessment for oil spills in Arctic waters was conducted, but only refined products were
considered because crude oil is not currently transported in the Arctic (WSP and SL Ross 2014b).
Therefore, the Panel reviewed this Arctic Marine Risk Assessment (AMRA) report as a general indicator
of risk of oil spills in the Arctic, with emphasis on drivers of risk and uncertainties within the risk
calculations.

The Panel’s review focuses on the environmental risk index component of the AMRA since calculation of
spill frequencies appeared to be the best possible given the available data. The authors discussed the
overall poor coverage of Arctic navigational charts (except for the Beaufort Sea and Foxe Basin) but
decided not to apply increased weight to spill frequency due to poor navigational information in specific
subsectors. The reason is that there had not been a significant number of reported spill incidents related to
navigational issues and, for the most part, marine traffic in the Arctic is performed by long-time operators
who use up-to-date charts and are familiar with their routes.

The logic behind the use of the same basic methods for calculating risk for North of 60 as for South of 60
is debatable. The PSI represents the degree of difficulty involved in coastal cleanup due to factors such as
remoteness and/or distance from first-response centres. Therefore, the weighting attached to the PSI for
Arctic conditions arguably should have been higher than for the south. Furthermore, the HRI might need
upward adjustment due to use by Indigenous peoples.

Recommendation: Shoreline mapping should be conducted in areas where oil and gas exploration
and development is already occurring or is expected, e.g., Beaufort Sea and the Mackenzie Delta.
The mapping should also consider other human activities that would affect relative risk, such as
mining, where large port facilities have been proposed involving not only transport of the minerals
out but fuel in, and increased commercial shipping traffic through the Northwest Passage, which
brings greater risk of fuel oil spills. The current rapid rates of coastal erosion should also be noted
and mapped for the purposes of continued monitoring.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 339
The AMRA authors assumed that ice-free shorelines are the most sensitive because the presence of ice
prevents the oil from reaching the shoreline. This was a simplistic assumption as there could be times
when the shoreline is partially ice-covered and the situation will be very dynamic. The highest risk would
probably be during the seasonal changes when multiple processes might control fate and effects (e.g.
erosion, scouring, slumping that would take oil and mix it in with shoreline materials or near-shore
sediment materials).

The AMRA authors noted the lack of information regarding Arctic shoreline types.

Recommendation: The Panel acknowledges that shipping traffic volume is low in the Arctic.
However, shipments of fuel oil to Arctic coastal communities do occur, and the risks of spills of fuel
oil should be assessed. Furthermore, in preparation for increases in Arctic traffic the AMRA
methodology needs to be revised and tailored for the Arctic.

Risks of spills during shipment of fuel oil by ship and barge as well as transfer of fuel from barges to
storage facilities require assessment in order to inform preparedness planning. The methods used for the
AMRA biological resources index were basically the same as for the South of 60° TCRA study. The
Panel questions the use of the same methods for the Arctic as for more southerly zones, particularly
sensitivity weights. Sensitivity in the Arctic should consider more than marine species and habitats, as
river deltas and the rivers themselves are important for anadromous fish 7 and other sensitive uses
(including human uses).

The calculation of the HRI involved the following:

 Coastal population index (CPI);


 Tourism index (TI);
 International freight tonnage index (IFTI);
 National freight tonnage index (NFTI); with
 HRI = 0.7(CPI) +0.2(TI)+0.05(IFTI)+ 0.05(NFTI)

The AMRA authors state that, “since the majority of the population in the area of study is Inuit, the
coastal population indicator captures explicitly the human use for these communities.”

This assumption that the coastal population captures human use of the coastline is an over-simplification
because many coastline areas with zero population can have high traditional use value. The Panel
questions whether the CPI would be found to be sufficient by Inuit people.

Recommendation: Additional research should be conducted in the near future to capture both
increases in marine traffic and the activities of commercial fisheries in the regional risk assessment.

The AMRA report acknowledged the value of royalties paid to Nunavut by offshore turbot and shrimp
fishing companies and also acknowledged the value of Arctic char and sealing to communities and state
that these activities are vulnerable to marine spills. It was felt that the large scale of the assessment
limited the incorporation of commercial fisheries as their activities may not be associated with coastal
environments.

Recommendation: Current environmental risk index (ERI) values are low or very low because of
projected low spill frequency. Therefore, the Panel is uncertain whether ERIs, in their present
form, are useful for planning. The Panel suggests using the ESI component of the ERI equation as a
means of identifying priorities. However, the way that the ESIs are calculated would have to be

7
Anadromous fish hatch in fresh water, spend most of their life in the sea, and return to fresh water to spawn; e.g., salmon.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 340
revisited using input from communities and scientists because the Panel is not convinced that the
current ESI methods are relevant and representative of Arctic conditions.

The authors state that, from an emergency planning point of view, the ERI can be interpreted as a relative
measure of the importance of risks associated with oil spills. A large ERI value implies a relatively higher
risk of economic and environmental damage. Therefore, the index could be applied to damage reduction
in Arctic waters if combined with emergency planning by authorities. However, the results of the ARMA
did not provide a particularly useful level of resolution, given the scale of the Arctic and the location of
population centres that may serve as the location for Arctic preparedness.

The AMRA report gave little consideration to risks in the Mackenzie River and Great Slave Lake (and
other freshwater areas), except with respect to the Mackenzie River delta where it enters the Beaufort Sea.
Traffic is low in these areas; however, the same can be said for the Beaufort Sea and the southern part of
Hudson Bay, including James Bay and the Foxe Basin, where ERIs were high in some cases. Given the
description of the ecology of the Mackenzie River and Great Slave Lake in Section 4 of the AMRA
report, and the presence of EBSAs, it is puzzling that the calculated risks for these two areas did not
exceed the “relatively low” category. There were no comments about the future risks for the River or the
Lake in Appendix 2 of the AMRA report.

8.14.1 Recommendations from the Arctic Marine Risk Assessment Report

The Panel concurs with the following recommendations made in the AMRA report:

 Matching vessel codes from the Automatic Identification System database and Transport Canada
commodity database would allow consideration of the exact trajectory of specific cargo—this
should be considered for the Arctic as well as other regions;
 The current analysis is based on yearly estimates of spill frequencies. The estimate could be
refined by taking into account the length of the shipping season within each subsector and thereby
considering seasonal variations that influence risk estimates;
 Evaluation of coastal areas facing climate change impacts is required;
 A separate analysis for fresh water (Mackenzie River and Great Slave Lake) is necessary to
account for differences in critical processes affecting fate and transport of oil as well as effects;
and
 Commercial fisheries information should be incorporated into the HRI.

Recommendation: The Panel also recommends that future relative risk assessments for the
Canadian Arctic include spills from exploration and production in adjacent U.S. waters of the
Beaufort Sea as well as future drilling in Canadian waters.

8.15 Assessment of Marine Oil Spill Risk and Environmental Vulnerability for the State of
Alaska

The Panel reviewed an assessment of oil spill risk for Alaska (Reich et al. 2014) to compare and contrast
the methods used with those used for the two Canadian risk assessments reviewed above. The results of
the Alaska risk assessment are briefly described in Chapter 7.

The Alaska assessment was at a smaller spatial scale than the Canadian AMRA assessment, but still
covered a large area. This scale required the use of similar generalizations regarding spill behaviour and
effects as those used in the AMRA.

Recommendations based on the differences between the Alaska and Canadian (AMRA) assessments that
should be considered for future risk assessments in Canada are presented below.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 341
Recommendation: Future relative risk assessments conducted for the Canadian Arctic include
season-specific influences on sensitivity.

The Alaska assessment included season-specific vulnerability, which is very important in Arctic regions.

Recommendation: Future risk assessments for the Canadian Arctic should include consideration of
crude oils, including exploration and oil production facilities.

The Alaska assessment considered crude oils, heavy oils, light oils and distillates, whereas the Canadian
assessment focused on refined products only.

Recommendation: The Panel considers the inclusion of dispersion modeling to be an essential


component of risk assessments of oil spills, particularly at the regional scale. However, the Panel
notes that the dispersion modeling done in support of the Canadian assessment was basic and
would require refinement for application to finer scales (with the accompanying requirement for
additional data including oil interactions with ice, and biodegradation potential as bioremediation
may be the most feasible cleanup strategy in remote Arctic areas).

Weathering processes and the amount of oil reaching shore were not modeled in the Alaska assessment.
However, the report recommended that for regions with high relative risk, trajectory and fate modeling
should be conducted.

Recommendation: The Panel notes that the vulnerability method used in the Alaska assessment is
applied at a finer-scale and does not rely on broad sensitivity categories applied to entire groups;
therefore, this approach may be more appropriate if finer-scaled relative risk assessments are
conducted in Canada.

Vulnerability was assessed for Alaska receiving environments using habitat (shoreline, bottom marine
and sea ice) and species sensitivity and recovery information compared with “sensitivity” in the Canadian
assessments, which combined physical, biological and human-use indices.

Recommendation: The Panel suggests that an assessment of future risks in the Canadian Arctic
include a critical review of the preparedness and response/mitigation options applicable to Arctic
conditions and the likelihood of these measures being effective in reducing both the probability and
consequences of oil spills, with special consideration of spill site remoteness and distance from first-
responder centres.

The Alaska assessment assumed that risk mitigation would reduce the probability of an incident becoming
a spill event and also assumed increased effectiveness of spill prevention and risk mitigation measures to
reduce spillage.

Recommendation: The Panel suggests that a Spill Risk Calculator tool for Canada would be
beneficial to preparedness planners and responders.

The authors of the Alaska study provided a Spill Risk Calculator Tool and noted that the transparency of
the method allows for quick updates of results as more data become available.

The Alaska assessment method was very data-intensive. A similar quantity of relevant data would not be
available for the Canadian Arctic except for a few relatively small areas.

Recommendation: The Panel strongly recommends a national database of information required for
relative risk assessment in the Arctic. This database would include government, academic and
industry sources, as well as traditional knowledge sources.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 342
Recommendation: The Panel suggests that a more detailed assessment of the Beaufort Sea be a
priority for the next round of relative risk assessments. The assessment could be used to identify
critical data gaps for parameters that drive risk in the Beaufort Sea.

Both the Alaska and Canadian assessments identified the Beaufort Sea as a high relative risk area for
crude oil spills in the future.

Recommendation: Although it would require some educated guesses (preferably in collaboration


with industry), incorporation of cleanup costs would add value to the results of relative risk
assessments for the Arctic.

Neither the Alaska nor the Canadian assessments included cost of cleanup because information on such

Ecological risk assessment (Figure 8.13) is conducted using a standard framework that starts with
Problem Formulation, proceeds through exposure and effects analysis and concludes with
characterization of the risk. Problem formulation includes: development of risk management goals and
objectives; identification of assessment and measurement endpoints; determination of the spatial and
temporal context; screening for stressors of concern; and selection of valued ecosystem components
(VECs). Engagement of interested parties in problem formulation can reduce the level of criticism and
debate because the range of values can be represented via selection of goals, endpoints, receptors, etc.
Exposure assessment is conducted via measurement and/or modeling. Effects assessment is usually
conducted by comparing predicted or measured exposures with derived effects thresholds based on
toxicity test results. Risk characterization estimates the level of effect, most commonly through the
calculation of so-called hazard quotients (the ratio of the predicted exposure and the effects threshold).

costs was lacking (due to the fortunate lack of major spills).

In summary, the Alaska analysis provided a conservative relative environmental risk map based on
maximum most probable discharge or worst-case discharge. The assessment did not account for
weathering and did not model amounts reaching shorelines, but it did include seasonal differences, a
fairly large number of discrete habitat types, and a cross-section of species or species groups that are
assessed for vulnerability using several parameters. The Alaska analysis did not provide socioeconomic
analysis, which can be regarded as a significant weakness.

8.16 Assessment of Risks of Hypothetical Spills from the Proposed Northern Gateway Pipeline

Detailed ecological risk assessments were conducted to evaluate risks to freshwater and estuarine
resources from hypothetical spills along the proposed Northern Gateway pipeline (NGP) route (Green et
al. 2015). The risk assessments conducted for the NGP were quantitative and detailed, in contrast to the
relative risk assessments reviewed above. The purpose of the NGP assessment was to evaluate the risks
associated with a full bore pipeline rupture at discrete locations.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 343
Figure 8.13 Framework for Ecological Risk Assessment. Image from EPA (1998).

These detailed assessments included modeling of dispersion and resulting concentrations of whole oil and
oil components on the water surface, shorelines, in the water column and in sediments of four different
receiving streams that were selected on the basis of environmental, socioeconomic, cultural and
traditional values (Green et al. 2015). The receiving streams also reflected a range of watercourse types
and had limited access, thus increasing response times. The predicted concentrations in and on water, in
sediments and along river and estuary shorelines were then assessed against thresholds for acute and
chronic effects for a range of species.

The four locations selected for the NGP assessment were:

 Chickadee Creek – low gradient interior river tributary discharging to the Athabasca system
upstream of a populated centre in Alberta (Whitecourt);
 Crooked River – low gradient interior river with wetlands entering a lake system in the Interior
Plateau, BC;
 Morice River – high gradient river along the western boundary of the Interior Plateau with
sensitive fisheries resources and downstream human population; and
 Kitimat River near Hunter Creek – high gradient coastal tributary discharging to a large
watercourse with sensitive fisheries resources, downstream human population and discharging to
Kitimat River estuary (Green et al. 2015).

The Panel’s review of the NGP assessment focuses on key findings of the dispersion modeling and
toxicity assessment and compares these findings with the spill case studies presented earlier in this
chapter. However, before presenting our comments on the detailed findings, we have some comments
regarding important aspects of the spill scenario.

The spill scenario was a so-called ‘full bore rupture’ releasing a volume of between 974 to 3,321 m3,
which would place the releases into the “moderate” or “large” spill categories used in the relative risk
assessments conducted for tanker spills by WSP and SL Ross (2014a,b). The hypothetical volumes
released depended on the volume of oil that would be contained within the pipeline from the full-bore
rupture point to the nearest shutoff valves in the pipeline on either side of the rupture. It was assumed that

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 344
13 minutes would elapse before shutoff valves were activated. “While spill durations would likely be
longer, as oil would drain from the pipeline, all oil was released in 13 minutes to represent conservatively
high in-water concentrations” (Horn and French-McCay 2015).

The NGP risk assessment did not assume any efforts to prevent oil from entering the river or migrating
downstream, nor did it assume any recovery of the oil. This ‘worse-case scenario’ for the conditions
outlined maximized impacts for those conditions; however, the overall degree of conservatism in the
assessment is subject to discussion and debate.

8.16.1 Review and Recommendations Based on the Key Findings of the Northern Gateway Pipeline
Ecological Risk Assessments

Recommendation: Risk assessment scenarios that assume a more likely combination of factors (less
than full-bore rupture with a longer elapsed time to response) would be more relevant and
comparable to actual spill incidents.

In the Panel’s view, the assumed time period to shutoff is not conservative, given recent experience; e.g.
the 17-hour delay for stopping the Kalamazoo spill (Section 8.10) and an unknown period of time (but
possibly several days) for the Nexen emulsion spill (Section 8.13). The assumed 13-minute elapsed time
before shutoff of oil flow significantly constrains the spill volume, thus fundamentally affecting all
subsequent outputs from the risk assessment.

Recommendation: Detailed risk assessments conducted in support of applications for new projects,
such as pipelines and production facilities, should include winter conditions, including full or
partial ice cover and spring flood conditions with accompanying ice jams.

The NGP modeling did not consider the presence of ice on or in the receiving streams. It is likely that the
presence of ice would be a ‘risk driver’. Oil might flow on top of the ice to impact river shorelines above
the normal flow height. Heated dilbit might locally melt ice and thereafter travel below the ice where it
would be difficult to detect and intercept. The presence of snow and ice could decrease evaporation—
alleviating explosion hazards but increasing toxicity risk at spring melt. Any assessment of ecological
risks in Canada should consider seasonal differences in parameters contributing to fate and behaviour as
well as to toxicity.

Recommendation: Future detailed risk assessments should consider the influence of animal
behaviour on exposure to oil spills.

Animal behaviour was not considered in NGP exposure modeling. As illustrated by some of the case
studies reviewed above, fidelity to particular habitats can be an important contributor to animal exposure,
especially to oil deposited to sediments or stranded on vegetation or organic debris. It would have been
interesting to compare risks to receptors such as salmonids with and without assumptions of fidelity to
particular locations in the receiving streams (e.g., spawning sites). British Columbia coastal rivers attract
grizzly bears during salmon spawning migrations, providing another example of a behaviour-related risk
scenario.

Recommendation: Differences and similarities between modelled and observed fate and transport
of dilbit, synthetic oil and condensate should be used to develop future research projects as well as
guidance for standardized monitoring of oil spills.

Several results of modeling of the fate and transport of three products—dilbit, synthetic oil and
condensate—were consistent with observations during actual spill events (Horn and French-McCay
2015). The dilbit selected was the dense and highly viscous MacKay Heavy Bitumen diluted with
Synthetic Light Oil. The synthetic oil was an intermediate Syncrude Synthetic Light Oil. The condensate

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 345
was a highly aromatic and low viscosity CRW Condensate Blend (Table 2.2). The following list presents
the NGP predictions and the corresponding observations from case studies (in italics):

 Extensive shoreline oiling within confined reaches of small creeks and parts of the larger
rivers was predicted, as was observed in the Pine River (light crude) and Kalamazoo River
(dilbit) spills. However, there was little consideration of the interaction between river stage,
exposed shoreline length and amount of shoreline oiled on falling or rising flow stages;
 The magnitude of river current during the spill event was of primary importance—higher
flows resulted in more extensive oiling of shoreline and transport further downstream
transport, as was observed in the Pine River and Kalamazoo River spills;
 Higher wind cases for a hypothetical dilbit spill resulted in oil deposition to sediments as a
result of wind preventing entrained oil from resurfacing and allowing more time for
interaction with suspended sediments and eventual sinking. The Wabamun Lake spill case,
although it was a lake rather than a river, also illustrated the importance of wind in
dispersing and mixing heavy crude in the water column as well as stranding heavy crude
along shorelines;
 Hypothetical dilbit spill cases typically resulted in the most extensive oiling of shorelines
compared to the other oils, as was observed in the Kalamazoo spill; however, the Pine River
spill of light crude also resulted in extensive oiling of backwaters and side channels; and
 Condensate was predicted to stay buoyant with little or no deposition –slicks were maintained
longer under low flow and low wind; and extensive evaporation was predicted. These
predictions mirror observations of the diluent in the Kalamazoo spill.

Other predictions did not have analogous spill cases, either because of the type of oil or because of the
type of receiving environment. These predictions included:

 Vertical turbulence would be induced by bottom roughness of the stream bed and,
secondarily, winds would entrain oil into the water, modified by the oil viscosity. This was
not observed in the river case studies, although there were no explicit attempts to examine
entrainment in the Pine River case;
 Floating dilbit that was not entrained or stranded ashore was predicted to emulsify quickly to
form highly viscous mousse, which prevented additional entrainment and sinking as it was
carried farther downstream. Dilbit mousse formation was not reported for the Kalamazoo
spill, nor has it been observed in open-tank experimental spills (Chapter 2). Dilbit is not as
amenable to emulsification or mousse formation as are conventional crude oils;
 Dissolution of soluble and sparingly soluble aromatics into the water was predicted to be
enhanced by entrainment as small droplets. This particular mechanism was not studied in the
Pine River case and there are no published data regarding this mechanism for the
Kalamazoo spills. However, it was suggested that wind-induced formation of Langmuir
troughs in Wabamun Lake might have contributed to vertical mixing of the spilled heavy fuel
oil;
 Under high flow and high wind conditions, bottom roughness and wind-induced mixing was
predicted to keep large amounts of condensate entrained in the water column. The
predominant fate of diluent (Kalamazoo River) and lighter fractions of light crude (Pine
River) was rapid surface dispersion and evaporation; however, temperature and current
regimes in the NGP assessment were very different;
 Above a certain site-specific threshold, dependent on streambed roughness and total
suspended solids, higher winds were predicted to maintain the entrained oil in the water
column allowing large amounts of deposition to the sediment. The mechanisms producing
deposition to the sediment were different in Pine River (where the primary drivers were slow

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 346
current and fine-grained sediments) and the Kalamazoo River (where the primary drivers
were factors leading to the production of OPAs); and
 The predicted behaviour of synthetic oil was distinctive and contributed to higher risks being
associated with this type of oil. The predicted behaviour of synthetic oil included:
 A relatively large percentage that became entrained and potentially settled into sediments.
Synthetic oil under both high flow and low flow conditions resulted in the largest
amounts of oil deposited to sediments;
 The entrained oil and high concentrations of dissolved aromatics were in many cases
transported down the full length of the modeled river; and
 As a result aquatic biota were exposed to acutely toxic concentrations along the modeled
reaches.

Recommendation: The Panel recommends that post-spill monitoring guidance include protocols for
examining hyporheic flow.

The Panel notes that the NGP assessment did not include a detailed consideration of entrainment of
dissolved and particulate oil into coarse sediments by downward hyporheic flows. These flows
characterize salmonid rivers and may explain the fate of much of the unaccounted for oil in spill case
studies. The NGP risk assessment mentions hyporheic flows only in passing as an inflow of groundwater
that would act as a diluent of any dissolved oil in gravel beds. To date, the Panel is not aware of any
reports of post-spill systematic investigations of oil distribution in stream gravel, particularly as it relates
to hyporheic flow down-welling zones. The same principles apply at the river’s edge where lateral
hyporheic flows generated by river bends transfer river water (and potentially oil) into ground water of
riparian zones.

Recommendation: Differences and similarities between modeled and observed acute and chronic
toxicity of dilbit, synthetic oil and condensate should be used to develop future research projects as
well as guidance for standardized monitoring of oil spills.

Predicted acute toxicity to water column organisms was greater for synthetic oil and condensate than for
dilbit because of the high viscosity of dilbit (which impeded entrainment), the lower concentrations of
low molecular weight compounds and PAHs in dilbit, and because of the smaller water areas and river
lengths affected by dilbit. However, acute injury to wildlife and emergent vegetation was usually highest
in modeled dilbit cases due to higher viscosity that kept it floating on the surface or adhering to the
shorelines.

These predictions are consistent with the spill case studies reviewed by the Panel. Acute toxicity to
aquatic life appeared to be minimal during the Kalamazoo spill, with no reported fish deaths. In contrast,
thousands of fish were killed due to the spill of light crude into the Pine River. Acute injury to wildlife was
low during the Kalamazoo River spill. Death of emergent vegetation was observed during the DWH spill,
but not during any of the freshwater spill cases reviewed.

Assessment of the potential for chronic toxicity extended for up to one or two years post-spill. The
chronic NGP assessment addressed: weathering on shoreline soils and in stream sediments; partitioning of
hydrocarbons from sediment back into pore water in streambed spawning gravels; potential effects to fish
eggs and embryos present in spawning habitat; and chronic risk to receptors that are higher level
consumers (Stephenson et al. 2015). Two models were used to evaluate chronic fate and bioavailability:
these models predicted the fate as a mixture of 61 hydrocarbons, including the Canada-wide Standard
aliphatic and aromatic TPH fractions, BTEX and other monoaromatic hydrocarbons, PAHs and other
miscellaneous hydrocarbon compounds. Long-term exposure to a full suite of alkylated and non-alkylated
PAHs that may be dissolved in the interstitial water of stream bed gravels was considered for potential
effects on developing fish eggs and larvae. Chronic exposure point concentrations were calculated for

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 347
water, sediment, oil, vegetation and fish tissue at time points of four weeks and one to two years
(Stephenson et al. 2015).

Chronic effects benchmarks were developed from the Target Lipid Model (TLM) of Di Toro et al. (2000)
and Di Toro and McGrath (2000). Canada-wide standards for PAHs in soils were used for the assessment
of risks to soil invertebrates and plants and CCME guidelines were used to assess toxicity of fresh crude
to livestock (Stephenson et al. 2015).

As discussed in Chapter 4, graphs of predicted toxicity using the TLM versus observed chronic toxicity
indicate very large variances (10-fold or greater). The versions of the TLM used by Stephenson et al.
(2015) have been superseded by newer publications; however the newer TLM predictions are still
imprecise and are based on many untested assumptions.

Predicted total PAH concentrations in sediment pore water were less than 0.2 µg/L after four weeks and
less than 0.05 µg/L after 1-2 years. Predicted river water TPH concentrations were mostly less than 0.01
mg/L. No significant risks were predicted for fish or benthic invertebrates, assuming additive toxicity of
the various hydrocarbon compounds (Stephenson et al. 2015). The exception was in the Chickadee Creek
scenario where fine-grained sediments were predicted to create conditions leading to higher
concentrations in sediment pore water. The authors stated that in gravels most likely to be used by
salmonid fish as spawning habitats, the expected average total PAH concentrations were below
benchmarks established to identify potential for blue sac disease or mortality of developing eggs (Chapter
4).

Although the statement regarding spawning gravels is logical if oil droplets are not entrained in
sediments by hyporheic flows, it does not acknowledge that some fish species (including salmonids such
as mountain whitefish) may use finer-grained sediments for spawning. Observed toxicity to lake whitefish
and northern pike embryos from exposure to PAHs in near-shore sediments of Wabamun Lake four to
seven months after the spill of heavy fuel oil illustrates this point.

Stephenson et al. (2015) discussed the possibility that crude oil could become entrained into and trapped
in river gravels as a result of hyporheic flows (Figure 2.3). They speculated that although this fate is
possible it would depend upon:

 Exchange between river water and subsurface pore water. This can occur within river reaches at
scales of hundreds of metres to kilometres, or locally at the scale of a single salmon redd.
Depending upon the season and local topography, rivers may alternate locally and seasonally
between ‘gaining’ and ’losing’ water from or to groundwater of hyporheic system;
 Whether upwelling dominates - for rivers like the Kitimat it is likely that hyporheic flows will be
dominated by inflowing groundwater and that such groundwater will be upwelling through the
gravels, thus limiting the potential for ingress and retention of crude oil; and
 Whether crude oil becomes dispersed by turbulent river flow as fine droplets that would not
rapidly resurface and would be available to become entrained into hyporheic flow should
conditions allow—such dispersion could occur for condensate although it has low density and is
likely to rapidly resurface and could occur for synthetic oil, but is less likely for dilbit, which is
more viscous. Stephenson et al. (2015) noted that the magnitude of predicted oil spill effects at
the population level would depend upon the timing and location of the spill with respect to fish
and their key habitat, as well as the availability of unaffected habitats either upstream or in
tributaries.

The Panel notes that fish demonstrate habitat fidelity so even if there were unaffected habitats, the fish
might not use those habitats but rather remain in the contaminated areas. This was the concern in the

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 348
Pine River spill (Birtwell 2003); however, follow-up monitoring of population-level indicators in the Pine
River was insufficient to either confirm or reject this hypothesis.

The Panel understands the emphasis on sensitive salmonid species when conducting risk assessments
since the standard practice is to select the lowest effects thresholds for the contaminants of concern (and
these are usually for salmonids). However, other species, such as white sturgeon or eulachon, would be
more vulnerable if the last few remaining spawning habitats were oiled or the food supply was reduced in
critical rearing habitats. Furthermore, species at risk would be less resilient to the loss of a year class (or
even a portion of a year class).

8.17 Assessment of Risks of Hypothetical Spills from the Proposed Trans Mountain Pipeline
Expansion Project

8.17.1 Semi-quantitative assessment of spills at selected locations

Kinder Morgan (2013) conducted a semi-quantitative risk assessment of the Trans Mountain Pipeline
Expansion Project (TMPEP). Quantitative estimates of failure frequency and qualitative estimates of
consequences were combined and applied to hypothetical spills at selected locations along the pipeline
route. Risk results were used to identify the need for additional pipeline design measures to reduce the
probability and/or the consequences of failure (e.g. relocation of a valve site to reduce potential outflow in
a high consequence area). A reliability approach was used to estimate failure rather than incident
statistics. Areas with higher potential consequences were identified based on land use or location with
respect to water bodies.

A worst case full-bore rupture was used for the TMPEP assessment; however, as for the NGP assessment,
the elapsed time to achieve valve shutdown was short—in this case only 10 minutes. The TMPEP authors
stated that 10 minutes for a readily identifiable catastrophic rupture is conservative because the Control
Centre Operation would recognize the event immediately.

The Panel has the same comment as for the NGP assessment—the assumption of 10 minutes to valve
shutdown significantly (and possibly unrealistically) constrains the spill volume, thus fundamentally
affecting the results of the risk assessment.

Recommendation: The Panel suggests that risk assessment scenarios that assume a more likely
combination of factors (less than full-bore rupture with a longer elapsed time to response) would be
more relevant and comparable to actual spill cases. These scenarios would also be viewed as more
credible by stakeholders sensitized by recent spills that experienced long delays before shutdown.

Eight case studies were assessed in the TMPEP report to provide relevant information for use in
developing assumptions and inputs to the qualitative assessment of consequences. The cases included the
Kalamazoo River, Wabamun Lake and Pine River spills. All of the case studies used are presented in
Table 8.2.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 349
Table 8.2 Case Studies Considered for the Trans Mountain Pipeline Expansion Project (TMPEP) Assessment of
Environmental Effects of Oil Spills
Volume
Oil Spill Location Year Spill Source Oil Type
(m3)
Pipeline Full- bore Diluted
Kalamazoo River Michigan, US 2010 3,200
Rupture Bitumen
Wabamun Lake Alberta, Canada 2005 Rail Accident Bunker “C” 712

East Walker River California/Nevada, US 2000 Truck Accident Bunker “C” 14

Pipeline Full- bore


Pine River BC, Canada 2000 Light Crude 985
Rupture
Pipeline Full- bore
Yellowstone River Montana, US 2011 Light Crude 240
Rupture
Pipeline Full- bore
OSSA II Bolivia, South America 2000 Mixed Crude 4,611
Rupture
DM932 Louisiana, US 2008 Barge Accident Bunker “C” 1,070
Heavy
Pipeline Third Party
Westridge Burnaby, BC 2007 Synthetic 224
Damage
Crude

The spill scenarios involved releases into the Upper Athabasca, North Thompson and Lower Fraser
Rivers. “Credible worst case spill volumes” ranged from 1,250 to 2,700 m3. The crude spilled was Cold
Lake Winter Blend dilbit (see Table 2.2).

Recommendation: Differences and similarities between the assessed consequences of a dilbit spill
into various locations along the Athabasca, North Thompson and Lower Fraser Rivers and case
study results should be used to develop future research projects as well as guidance for
standardized monitoring of oil spills. Some suggested topics for further investigation are indicated
in italics.

The qualitative assessment of consequences varied with season and spill location. For locations on the
Athabasca River and Thompson River, consequences in winter were assessed as low (except for localized
effects on aquatic invertebrates) because many of the receptor organisms would be dormant and overland
flow would be slowed with some oil absorbed into the snowpack. Furthermore, it was assumed that
environmental effects may be minimized because most of the oil is recoverable. The consequences in
summer in the Athabasca and Thompson Rivers were predicted to be much greater because of rapid
dispersion due to high flows, entrainment due to turbulence, and deposition along shorelines. In addition,
the TMPEP assessment assumed that as the oil is transported downstream and weathers, it becomes more
viscous and dense and interacts with shoreline sediments, e.g., forming OPAs and resulting in oil that
submerges in low energy areas such as eddies and backwaters. The TMPEP authors stated that although
water in the Athabasca and Thompson Rivers is somewhat turbid, the suspended sediment load is not
particularly high, little oil would be entrained in the water column, and the water has no appreciable
salinity. Thus, OPA formation would not be a dominant factor in the fate of the spilled oil. The extent of
dispersion was assumed to be greatest in the summer, extending as far as 100 km downstream. The
magnitude of effects on aquatic biota ranged from low to high depending upon the group and the
proximity to the spill.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 350
Consequences in spring or fall were considered to be intermediate in the Athabasca and Thompson Rivers
due to lower and less turbulent flows than during the summer. Some submerged oil was assumed to be
present in eddies and backwaters, as was predicted for summer conditions. The extent of downstream
dispersion was assumed to be about 25 km.

Actual seasonal differences in the consequences of spills may differ from those predicted by the TMPEP
assessment. For example, winter time spills involve much slower loss of volatiles and a longer time for
dispersion to occur, and recoverability of a winter spill will depend upon the logistic challenges
associated with the spill location.

The conclusion of the TMPEP authors that OPA formation would not be a dominant factor in the fate of
the spilled dilbit is subject to debate. If a spill occurs during high flow when suspended sediment is
elevated and/or if the flow-path involves interaction with riparian and floodplain materials (as was
observed in the Kalamazoo spill), OPA formation could be much more significant.

The Fraser River has ice-free conditions during the winter; thus, predicted winter consequences in the
Fraser River differed from those predicted for the Athabasca and Thompson Rivers. While some receptors
would be dormant, birds such as bald eagles and some waterfowl could be present and potentially
experience high magnitude consequences. High to medium magnitude consequences were predicted for
aquatic receptors in side channels.

Summer consequences in the Fraser River location also differed because a side channel at this location
would be heavily affected during high flows. Much of the oil was predicted to become stranded along
shorelines and in riparian areas. The potential for OPA formation and submerged oil in the side channel
may have been underestimated given the experience with the Pine River, Kalamazoo and Wabamun
spills.

Spring and fall consequences were also predicted by the TMPEP assessment to be greater in the Fraser
River location because flow conditions would disperse the oil as much as 60 km downstream during a
time when many migratory birds would be present and semi-aquatic mammals would also be present and
active. Effects were assumed to be insufficient to affect regional bird populations. However, this
conclusion may not apply to species-at-risk, to situations where a population has insufficient resiliency to
withstand the loss of a substantial portion of breeding pairs, or if sublethal effects of residual oil on food
supplies essential to migrating birds results in reduced fitness.

In summary, the qualitative assessment of consequences of modeled dilbit spills in the TMPEP report was
reasonably consistent with case study data, with appropriate conservatism added to account for
uncertainty. The potential for OPA formation may have been underestimated given the experience with
the dilbit spill into the Kalamazoo River, as well as the behaviour of heavy fuel oil in the Wabamun Lake
spill. The recovery period was assumed to be from 12 months to five years post-spill for almost all of the
spill scenarios. No evidence was presented in support of this assumption; however, the limited
information on recovery from case studies suggests that this range is reasonable for most receptors; e.g.
fish in the Pine River (Section 8.8).

8.17.1.1 Stochastic modeling of spills at locations near and at the mouth of the Fraser River

Recommendation: The results of the stochastic oil spill fate and transport modeling for a spill at the
Fraser River and Delta location should be used as input for spill preparedness planning, as well as
for identification of priority requirements for verification of the model.

The TMPEP assessment included stochastic oil spill fate and transport modeling in support of the risk
assessment for a spill at the Fraser River and Delta location near the Port Mann Bridge. Stochastic
modeling was performed due to the complexity of the situation at that particular hypothetical spill

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 351
location. The modeling results by season provide valuable insights regarding the most probable
consequences and implications for cleanup. The results are summarized below, accompanied by
comments from the Panel, which are in italics.

Winter: >80% of oil is stranded within three days of spill; 11% evaporates; < 5% remains on the surface;
< 1% is submerged, biodegrades or dissolves; < 0.1% forms OPAs.

 The implications are that the spill response will be focused on stranded oil, with potential for
substantial disturbance of habitat during the cleanup. Also, if most of the oil cannot be removed
before higher spring flows, re-mobilization might occur.
 The assessment assumed that many of the receptors are absent or dormant in winter—this would
be true for plants (dormant) and some (but not all) invertebrates, and perhaps for amphibians,
but many birds overwinter at this location.

Summer: < 60% of oil is stranded, about 10% evaporates; about 30% remains on the water surface; < 1%
is submerged, biodegraded or dissolves; < 0.1% forms OPAs.

 There is still a substantial amount of stranded oil, resulting in medium to high effects on riparian
vegetation and in marsh areas; trade-offs between the effects of the stranded oil and the effects of
the cleanup will require very careful evaluation.
 The oil remaining on the water surface is predicted to result in medium to high effects on ducks
and geese and high effects on semi-aquatic mammals; which illustrates the importance of rapid
and effective removal of floating oil.
 OPA formation may have been underestimated, given the sediment load of the Fraser River.

Spring and Fall: 70% of oil is stranded; 10% evaporates; 20% remains on the surface; a small amount of
oil (< 1%) is predicted to become submerged, undergo biodegradation or dissolve; OPA formation is not
predicted (< 0.1% of oil volume).

 There is an implied assumption that spring and fall represent one condition, which does not
account for the highly stochastic weather, discharge, ice state and biological events that typify
these seasons. One condition does not encompass the range of conditions, nor the change in
conditions during or immediately after a spill. Thus, a source of significant uncertainty does not
seem to be recognized. The same argument applies to winter and summer, although the size and
frequency of changes are not as large.
 The assessment once again focused on the effects of recovery of stranded oil. This common theme
again highlights the importance of having a sufficient understanding of the receiving environment
to enable confident evaluation of the trade-offs between the effects of residual oil and the effects
of further cleanup.
 The TMPEP assessment assumed that riparian habitat along the mainstem of the Fraser River
would be remediated with less intrusive methods and a greater emphasis would be placed on
natural attenuation of low levels of residual oil. This is a debatable assumption given the
potential effects of the modeled spill on socioeconomic and cultural values. The Panel suggests
that the net environmental benefit calculations would be challenging and subject to considerable
uncertainty.
 High effects on semi-aquatic mammals (due to oiling) were predicted and medium to high effects
on ducks and geese were predicted (including reproductive effects caused by transfer of oil to
eggs or effects on habitat quality caused by disturbance arising from oil spill response efforts).
 Particular attention was paid to the western sandpiper. Hundreds of thousands of these birds may
congregate in the assessment area, feeding on biofilm and invertebrates—the fate modeling

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 352
showed that the probability of oiling areas used by the birds is very low—therefore the
sandpipers are unlikely to be significantly affected and any effects are likely reversible.
 The Panel notes that the reliability of the TMPEP fate model is paramount regarding risk to
western sandpiper.

Recommendation: The Panel recommends the use of stochastic fate modeling for complex, sensitive
receiving environments because of the additional important information provided to decision-
makers regarding the probability of the various levels of effect on specific VECs. This information
can assist decision-makers in focusing preparedness and response plans. Stochastic modeling also
provides a basis for prioritization of environmental data collection.

8.17.1.2 Comparative assessments of Unmitigated and Mitigated Spills at the Westridge Marine
Terminal

Assessment of an unmitigated and a mitigated spill of heavy synthetic crude oil blend at the Westridge
Marine Terminal produced a mass balance comparison, presented in Table 8.3 (Kinder Morgan 2013).
Table 8.3 Mass Balance Comparison for a 160 m3 Spill at the Westridge Marine Terminal
Amount (m³) Unmitigated Case Mitigated Case
On shore after 1 day 16.6% 10.7%
Left on water after 1 day 0.4% 0.1%
Evaporated after 1 day 1.2% 0.9%
Dissolved after 1 day 1.8% 1.2%
Biodegraded after 1 day < 0.1% < 0.1%
Inside the containment area but not yet
80% 0%
recovered
Recovered inside the containment boom n/a 80%
Recovered at sea n/a 7.1%

After one day, no oil was predicted to remain on the water with mitigation, compared to 80% if
unmitigated (although oil inside the containment area would be recovered over subsequent days).
Mitigation reduced shoreline oiling but did not eliminate it (10.7% on the shoreline with mitigation versus
16.6% on the shoreline without mitigation).

Recommendation: The Panel urges incorporation of lessons learned from incidents as well as
assessments conducted for proposed project into preparedness and response planning by
responsible authorities and industry. The Panel notes that common themes among incidents and
assessments include communication, coordination, adequate resourcing and training.

The authors noted that the evaluation of responses to a spill at the Westridge Marine Terminal contributed
to the following ’lessons learned’ (these lessons learned are in addition to those presented in Section
8.11).

 Proximity of ready response capability to a spill site together with site-specific response plans
(which responders have exercised) help to greatly increase the effectiveness of response. In the
case of the simulation study, the modeling runs helped WCMRC gain a good understanding of
key requirements to effectively improve response. This is similar to the iterative learning
achieved through oil spill exercises.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 353
 There is no substitute for establishing an early line of defense by rapidly booming a release or
damaged vessel, when possible. This knowledge is tempered by the reality that health and safety
conditions, suitable nearby anchoring sites and operational constraints may not always make this
outcome possible.
 Recovery assets should be located in relatively close proximity to the spill, as would be the case
for Westridge Marine Terminal.
 Use of model-derived trajectory and slick thickness information to direct skimmers can help
identify optimum recovery locations. While remote sensing offers considerable opportunities for
spill detection, it also has limitations. The combination of numerical modeling and remote-
sensing data provides the most powerful approach to both current and future predictions of slick
positions.

The WCMRC has proposed a harbour base that would be continuously staffed if the TMPEP is approved
(Kinder Morgan 2014).

The Panel notes that some of the above lessons learned would also apply to the Marathassa spill which
occurred in English Bay in 2015 (Section 8.12).

8.18 Assessment of Accidents and Malfunctions for the Energy East Pipeline Project

The Panel reviewed the assessment of accidents and malfunctions for the proposed Energy East Pipeline
project (EEPP), which included consideration of the fate, transport and effects of three representative
crude oils: Bakken crude oil, Husky Synthetic and Western Canadian Select diluted bitumen (Table 2.2).
Four categories of rivers and streams were assessed based on magnitude of flow and width.

EEPP assumptions included:

 The release of the entire volume of the spill directly into the water body;
 Complete, instantaneous mixing; and
 The entire volume of Constituents of Interest dissolved into the water column (Energy East
2014a).

These assumptions were referred to as ’conservative’. However, the spill case studies illustrated that
release to a terrestrial flow path leading to a watercourse can have significant consequences in terms of
fate and behaviour of spilled oil as well as consequences to receptors along the flow path (see also the
TMPEP assessment). Complete, instantaneous mixing and complete solubilisation are not conservative
assumptions because there would be no floating oil to cause acute mortality of birds and semi-aquatic
mammals. These assumptions do not consider acute effects of floating oils or other important fate
mechanisms that have been shown to significantly affect consequences (e.g., the formation of sinking oil).

The EEPP assessment was based on simplistic estimates of hydrocarbon concentrations, reliance on low
likelihood of a spill occurring at particular locations, and blanket assurances regarding mitigation, which
is illustrated by the following quote:

“The environmental effect of a crude oil spill would vary both temporally and spatially depending
on the volume, timing and location of the spill. Localized effects could occur from virtually any
size of crude oil spill. In general, the chance of a spill occurring at any discrete location along the
pipeline or marine transportation of oil is low and if a spill were to occur, it likely would be
relatively small (4 bbl or less). In the event of a spill, Energy East would respond according to
applicable regulations and their Environmental Response Plan which will contain provisions for
protecting and mitigating potential effects to environmental receptors. In addition, in the event of

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 354
a spill, Energy East will consult with regulatory agencies to determine the appropriate and
preferred approach to cleanup and monitoring” (Energy East 2014b).

The Panel was initially encouraged to see that the EEPP assessment considered locations along the
pipeline route that are rarely evaluated in the context of oil spills; e.g. the South Saskatchewan River, the
Red River, a lake in Ontario (Trout Lake), the Rideau River, and the Iroquois River, NB (Energy East
2014b). These locations represent a wide variety of receiving environments with important differences
that can have substantial effects on oil fate, behaviour and toxicity. Unfortunately, the simplistic nature of
the assessment did not address these characteristics in any meaningful way (e.g., sediment loads in the
South Saskatchewan River, morphometry of Trout Lake). Instead, the same statement that is quoted
above was repeated for each location, with slight (if any) variations.

Statements regarding the low probability of a spill were also very similar for all assessed locations. The
rationale for low probability included such points as:

 The Project’s pipeline materials and design, as well as the strategic placement of valves, are
expected to minimize the likelihood of a spill and reduce the volume of oil released in the event
of a spill;
 Rivers will be crossed by horizontal directional drilling, which will reduce the probability of a
rupture by reducing threats to the pipeline given the depth of the pipe below the river. The depth
and overlying materials also would help confine a release and reduce the possibility of the crude
oil reaching the river;
 Valves are strategically located along the EEPP route to reduce the amount of crude oil that
potentially could be spilled. The location of valves, spill containment measures and emergency
response procedures would mitigate adverse effects to both surface water and groundwater; and
 The assessment used conservative assumptions to overestimate risk. Many assumptions, such as
all benzene from the oil would be instantly dissolved into the water, are unrealistic but help to
screen for the potential for effects.

The Panel found the EEPP assessment to be overly simplistic, with blanket assurances regarding
mitigation, generalized assumptions that mitigation would always be effective, and a reliance on low
probability to reduce risk. The claim that the assessment used conservative assumptions to overestimate
risk is incorrect since assuming instantaneous mixing of a spill would be conservative with respect to fish
but liberal with respect to birds and semi-aquatic mammals. The EEPP assessment did not add any value
with respect to risk-based prioritization of mitigation measures along the pipeline route.

Recommendation: The Panel suggests that the sites of interest used in the EEPP assessment merit
being re-analysed using analytical tools similar to those used for Trans Mountain or Northern
Gateway pipelines—particularly with respect to fate and transport and the level of sophistication
with respect to acute and chronic toxicity, shoreline effects and overland flow effects with season.

8.19 Summary Comments and Recommendations Regarding Risk Assessments of Oil Spills
within Environmental Impact Assessments

The NGP and TMPEP risk assessments are impressive documents in their scope and detail and are among
the best of their kind. The fate modeling in support of the NGP assessment and the use of stochastic fate
modeling for particularly sensitive and complex receiving environments done for the TMPEP assessment
provided valuable details with respect to the relative proportion and location of floating, dissolved,
entrained, stranded and submerged oil. The use of case study information to develop a qualitative effects
assessment for the TMPEP provided a useful alternative to the calculation of quantitative hazard indices
using single point-estimates of effects thresholds.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 355
Although the NGP and TMPEP risk assessments had several strengths, they were subject to the same
uncertainties and challenges as any large-scale assessment. The results of the assessments can be used as
valuable input in identifying priorities for preparedness and responses as well as for focusing on critical
uncertainties associated with factors which drive the risk. However, the results of the assessments cannot
be assumed to be reliable predictors of what will actually happen during a spill. The state-of-the-science
of risk assessment has not yet achieved the required level of verification via comparison of predicted
versus observed effects. Furthermore, uncertainty in risk assessment will always be substantial, and in
some cases irreducible. “Personal judgment and expert opinion will necessarily be a part of such risk
assessments because they handle rare events in complex systems” (Hague et al. 2014).

The development of the scenarios to be used is one of the most difficult challenges in risk assessment, and
the NGP and TMPEP assessments illustrate this challenge. In both cases, a full bore pipeline rupture was
assumed and described as “worst case”. However, the legitimacy of the description of the risk scenarios
as “worst case” is subject to question because of the assumed short time period between detection of a
pipeline breach and shutting off of valves to stop the flow of oil. Other sources of uncertainty regarding
the “worst case” nature of the risk scenarios include:

 Lack of assessment of winter spills for the NGP;


 Choices of spill sites which, in some cases, reduced the amount of oil reaching the most
sensitive areas (e.g. reducing oil reaching the estuary at Kitimat) (Hodson and Martin 2012);
 Inclusion of receptor species that may be sensitive because of their habitat, conservation
status or cultural value (e.g. eulachon) (Hodson and Martin 2012);
 The use of average physical and chemical characteristics of river receiving environments
(Hodson and Martin 2012); and
 Missing or inadequate representation of potentially important processes, such as hyporheic
flows.

The development of risk scenarios is a fundamental issue because they represent high stakes, uncertain
facts and conflicting values (Hauge et al. 2014). As Hauge et al. point out, uncertainty makes different
interpretations possible, and values may be embedded in knowledge production and interpretation (as
discussed in Section 8.6 with respect to the EVOS). The ‘appropriate’ size, location, flow-path, time-
scale, receiving environment and receptor species for each oil spill assessment is subject to debate, not
only regarding which particular combination might best represent a worst case, but also whether a worst
case is what is needed to support decisions.

In some contexts, worst case scenarios are so highly unlikely as to be of little value for decision-making.
It may be more useful to choose more likely scenarios which occur more frequently.

Reduction of some of the uncertainties associated with risk assessment is possible. For example,
including migration patterns and other behavioural characteristics (e.g. fidelity to feeding grounds and
spawning sites) and combining these patterns with stochastic dispersion modeling would offer new
insights into the potential exposure of aquatic receptors. Unfortunately, data required for more
sophisticated spatially and temporally-explicit exposure modeling is often lacking. The use of chronic
toxicity data from the most relevant test species, combined with field information from spill case studies
would produce less uncertainty than using acute:chronic ratios. Careful attention to all exposure routes for
each receptor helps ensure that exposure is not underestimated.

Several sources of uncertainty are very difficult to reduce. Major oil spills are rare and each of these rare
events is unique - making generalizations from one case to another very difficult. Stochastic processes
such as weather influence the fate and effects of oil spills in unpredictable ways.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 356
The framework and tools for ERAs originally were developed for relatively small sites; their applicability
to 1,000’s of kilometres of pipeline crossing highly variable habitats with major physical and chemical
gradients is questionable. Of necessity, risk assessors must select representative sites along these routes
and the assessed risk is ultimately a comparison of modeled concentrations of hydrocarbons in water,
sediment (or sometimes tissue) to a single benchmark of toxicity. Rarely, modeled concentrations are
compared to a dose-response curve. Effects of exposure to several chemicals are usually assumed to be
additive, although the validity of this assumption is highly questionable, particularly when chemicals with
very different modes of action are combined. In particular, the interactive effects of oil and municipal and
industrial effluents have not been considered for risk assessments in freshwater and coastal ecosystems.
The existing condition of the receptor is almost never explicitly considered; e.g., if a fish population is
already exhibiting poor condition due to the existence of other stressors in the receiving environment.

Given the limitations of the toxicological approach to risk assessment, there have been many calls for the
incorporation of ecology into risk assessment. The use of spatially-explicit exposure modeling that
incorporates seasonal habitat use by each receptor and superimposes contaminant gradients or patterns
has been a significant advance; however, such modeling rarely appears within impact assessments. More
commonly, a weight-of-evidence approach is used, which combines laboratory, field and modeling lines
of evidence into an overall risk estimate. The evaluation of effects of exposure to multiple stressors is the
subject of active research; however, the results of this research seldom are reflected in risk assessments.

In summary, although many improvements are needed in both exposure and effects evaluations within
risk assessment, the most fundamental challenges centre on the framing of the risk assessment. Risk
framing can be from the perspective of informing management, or it can be from the perspective of
deciding whether the risk is acceptable to society (Hague et al. 2014). Public concerns about impacts may
be very different from a petroleum company’s concerns. Often, risk assessments are part of processes
where questions can be asked of science and yet cannot be answered by science (Weinberg 1972 cited in
Hague et al. 2014). Risk assessments cannot provide the level of confidence often demanded by the public
and political decision-makers—to achieve the required confidence would take decades of research, and
even then, we may not be able to adequately represent the complexity of ecosystems.

8.20 Assessment of Risks Associated with Transport of Crude by Rail

A major increase in crude-by-rail transport started around 2009. Since then, crude-by-rail transport in the
US has increased by 40-fold or more (Etkin 2015). The bulk of the oil is crude oil from the North Dakota
Bakken fields and the Alberta oil sands. A 42% increase has occurred in average annual spillage of oil by
rail in the US but at a reduced rate per volume transported (Etkin 2015) (Table 1.4). However, accidents
involving crude-by-rail are of particular concern because of the potential for fires and explosions and
impacts on urban areas traversed by rail lines.

Etkin (2015) cited analyses showing a steep decline in the frequency of derailments per train mile, with
about one mainline derailment for every 3.85 billion tons of freight (including but not limited to crude-by-
rail). It has been estimated that about six crude-by-rail related derailments would occur annually.
Consequences were determined by oil type, volume, location and timing.

For Bakken crude, the property of greatest concern is volatility leading to explosion and fire, e.g. the Lac
Mégantic, QC, disaster. For diluted bitumen the greatest public concern is that oil will become submerged
when it reaches a body of water. Bitumen blends vary considerably depending on the source, blending
procedures and diluent used (Table 2.3); therefore, the potential for sinking will vary. However, the
Kalamazoo spill and the results of modeling in support of risk assessments illustrate that sinking is a
definite possibility when dilbit is spilled, depending upon weathering-related density changes, turbulence,
salinity and the amount of sediment in the water.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 357
Estimated spill volumes for crude-by-rail Unit Trains are presented below (from Etkin 2015). The ranges
are based on a less conservative 9.2% release rate to a more conservative 16.7% release rate:

 25th percentile – 21 to 38 m3 involving 2 cars


 50th percentile – 52 to 95 m3 involving 5 cars
 75th percentile – 104 to 190 m3 involving 10 cars
 90th percentile – 178 to 322 m3 involving 17 cars
 95th percentile – 240 to 436 m3 involving 23 cars
 Worst case – 1,274 to 2,313 m3 involving 122 cars

In actual crude-by-rail spill cases in Canada and the US in 2013 and 2014, the lowest volume spilled was
17 m3 and highest was 2,909 m3, which was greater than the estimated worst case (Etkin 2015).

The higher percentile spill volumes (and the highest actual spill volume) are substantial spills, comparable
to tanker spills and the larger pipeline spills. Therefore, preparedness and response plans should not
discount the potential for higher-volume spills from crude-by-rail accidents. Challenges related to
response to spills of Bakken crude transported within Canada include:

 Fires in populated areas or in locations where wildfires may be triggered; and


 The adequacy of first responder systems in locations that have seen a sudden increase in crude-
by-rail (e.g., from the Estevan area of south-eastern Saskatchewan).

The Canadian Association of Fire Chiefs is reviewing the preparedness for crude-by-rail spills (Etkin
2015).

Responses to Bakken spills have included the use of sorbent and containment booms along with water
spraying to corral oil for skimming and vacuum pumping. Sorbent pompoms and pads have been used in
some areas (Etkin 2015).

Response to spills of dilbit would include the requirements for general preparedness for inland locations,
consideration of the flammability of diluent, and development of capabilities for submerged oils.

Etkin (2015) noted that risk mitigation through prevention is key. Prevention measures would include
training of railway personnel and first responders, positive train control, improvements in braking
systems, and track inspections and maintenance. Newer tank car designs would reduce the potential for
leakage or spillage. Reducing volatility of the oil (e.g., via conditioning of Bakken crude) would be
another preventive measure. Canada and the US recently jointly announced forthcoming wide-ranging
changes to rail safety regulations including braking systems, speed limits, improved tank car
specifications, and rules regarding securing of unattended trains8.

8.21 Overall Conclusions Regarding Assessment of Risks of Oil Spills

Assessments of the risks of hypothetical spills conducted in support of the environmental impact
assessments of proposed pipeline projects in Canada have used unrealistic assumptions regarding
spill response time, based upon experience to date. The more detailed and comprehensive risk
assessments, such as those conducted for the proposed Northern Gateway Pipeline, were sufficient for
risk-based prioritization of preparedness and response planning as well as mitigation via pipeline design.
However, the consequences were likely underestimated given the assumption that it would take only a
few minutes to recognize a pipeline breach and deploy shutoff valves.

8
e.g., http://www.cbc.ca/news/canada/montreal/lisa-raitt-u-s-counterpart-announce-broad-new-standards-for-rail-safety-
1.3057150 (May 1, 2015).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 358
The review of risk assessments also revealed a wide “range of practice” applied to assessments by
different practitioners. The Panel was concerned about the breadth of this range, particularly since the
range was demonstrated in submissions to the same regulator (the NEB). Therefore, in addition to
research and monitoring to provide information needed to support risk assessments, there appears to be a
need for an explicit ‘standard of practice’ regarding the assessment of risk of oil spills. This is particularly
important given the high level of concern associated with the production and transport of oil.

There is an urgent need for investigations in Arctic Canada regarding the fate and behaviour of oil
under Arctic conditions, the sensitivity (and resilience) of Arctic environments and the effectiveness
of available spill response measures. Arctic marine systems include unique habitats, including the
epontic habitat. Epontic communities are currently the subject of research; however, the potential effects
of oil spills on these important communities should be investigated. The results of Arctic marine research
should be consolidated periodically to serve as a baseline for monitoring activities (SL Ross 2014).

Investigations of oil spill effects must include collaboration with Indigenous peoples to ensure that
traditional knowledge is incorporated into our overall understanding of the risks of oil spills.

There is a critical need for a coordinated and integrated database of information relevant to the
assessment of risk of oil spills in Canada. The data needed for input to risk assessments in Canada are
often either absent or widely scattered among government agency, industry, and academic sources. The
highest priority should be accorded to assembling, and if necessary, sampling to build an understanding of
pre-spill conditions at locations that are under the highest threat of contamination. The Panel suggests that
efforts initially focus on geographical areas at risk that have received less attention with respect to oil
spills (e.g. inland rivers, portions of the Great Lakes with the highest oil transport-related activities, and
the Gulf of St. Lawrence).

There is a compelling need for coordinated environmental sensitivity assessment in Canada using
standard tools and analytical approaches. This assessment must include integration of traditional
knowledge and information about traditional uses of resources. The gap in sensitivity mapping is larger in
inland Canada. Some industries are conducting their own sensitivity mapping; however, this information
is generally proprietary. Sensitivity mapping efforts should be prioritized using screening tools that can
identify the most likely sensitive watercourses subject to oil spills because of their location in proximity
to production or transportation routes, characteristics that would affect fate and behaviour in such a way
that risks are increased, and the presence of sensitive environmental components. Partnerships among
industry, government and academia will be necessary if adequate coverage of receiving environments in
Canada is to be achieved.

Regarding the potential for spills in the Canadian Arctic, SL Ross (2014) suggested:

 Consolidate the rapidly expanding knowledge about the aspects of the ecology of Arctic species
relevant to monitoring (spatial distribution, populations, habitat use, seasonality);
 Prepare a review of research into effects of oil spills on Arctic marine and anadromous fish
species (e.g., Arctic and polar cod, Arctic herring, Arctic char, Arctic cisco). The Panel adds the
need for consultation with Aboriginal peoples regarding their knowledge of fish species and the
areas most sensitive to oil spills, e.g., char spawning areas;
 Develop a decision tree to help decide whether a field monitoring program is required during a
spill incident. The Panel suggests that field monitoring would be required for all but small spills;
therefore, the decision tree could be designed to assist in the determination of the scope of the
monitoring program; and
 Prepare a readily accessible up-to-date directory for all regional environmental emergency
coordination centres in Canada, including the names and contact information for the senior
personnel. The Panel strongly endorses this recommendation.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 359
In addition to the above more general recommendations, SL Ross (2014) suggested that it would be useful
to develop an understanding of the sensitivity of Canadian species to PAH-induced immunosuppression
and the potential long-term consequences of it in Canadian fish populations during spills.

Focusing research and management on ecological resilience would be useful both in principle and
for specific spill scenarios. As the NRC (2013) pointed out, resilience provides a valuable conceptual
framework for managing complex systems because it focuses attention on how systems are affected by
short-term disturbances and long-term stresses. Resilience can be either increased or decreased through
management decisions. Analysis of the trade-offs inherent in management decisions would benefit from
the use of a resilience framework.

Adopting resilience as a fundamental framework will require much more attention to the acquisition of
baseline data, both for the natural environment and for the socioeconomic environment.

The Panel suggests that our focus should be on prevention of large spills and rapid and effective
response to smaller ones because we will never be able to eliminate spills. Therefore, the Panel
recommends:

 Identifying where most of the spills are occurring and why (e.g., pipeline spills into wetlands are
more common than into rivers; truck spills are more likely to enter storm sewers and then rivers
rather directly impact rivers; rail lines often run parallel to rivers and derailments may occur
more frequently with certain load and geographical configurations, etc.);
 Examining past spill response records, plus the current risk management processes and
regulations and their effectiveness and then honing in on the weak spots;
 Filling in critical data gaps regarding environmental sensitivity;
 Conducting relative risk assessments of the common spill location types overlain on the most
sensitive receiving environments; and then
 Identifying the combination of spill source, receiving environment and level of preparedness that
most urgently requires attention.

8.22 Research Needs and Recommendations

The following list presents knowledge gaps and corresponding recommendations, in order of priority.

8.22.1 High priority research needs (short-, medium-, and long-term)

8.22.1.1 Gap: Baseline information

Recommendation: Collect and Evaluate Baseline Information from High-Risk Areas (Short-term
and Extending to Long-Term)

Research is needed regarding the current status of sensitive species and vulnerable habitats for specific,
pre-defined locations in Canada representing a range of human disturbance, from relatively undisturbed to
highly disturbed. This information can then be used as the base case for assessment of any future spill.
The priority locations for collection of baseline information are those identified as high-risk in recent
assessments conducted for Transport Canada (WSB and SL Ross 2014a,b), combined with information on
areas with proposed new human activities that may increase the likelihood of spills and/or contribute to
cumulative effects of spills (e.g., LNG terminals along the British Columbia coast). Input from
Indigenous peoples and from key stakeholders should be part of the selection process for baseline
collection areas.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 360
8.22.1.2 Gap: National, accessible databases for use in assessment of oil spills

Recommendation: Establish a National Database of Information Relevant to Risk Assessment of


Oil Spills, Preparedness and Response (Short-Term)

Baseline Data: There is a critical need for a coordinated and integrated database of information relevant to
the assessment of risk of oil spills in Canada. (e.g., data collected for EIAs, EEM programs, compliance
monitoring (e.g., for municipal or industrial discharges), research programs, etc.). The Panel is aware that
some databases are already established (e.g., for EEM data); however, there is no readily accessible,
national portal to metadata on freshwater systems. Since the assembly of such a database would be a huge
undertaking, the Panel suggests that efforts initially focus on geographical areas at risk that have received
less attention with respect to oil spills (e.g., inland rivers, the Great Lakes).

The Panel strongly recommends that the national database include data for the Arctic.

Experimental and Monitoring Data: To support predictive numerical models and operational guidelines
for spill response, there is a need for a rigorous database on the fate, behaviour and effects of various
types of oil spilled and the efficacy of current and emerging oil spill countermeasures over a range of
environmental conditions.

8.22.1.3 Gap: Shoreline sensitivity in the Arctic

Recommendation: Extend Shoreline Sensitivity Mapping to Selected Arctic Locations (Short-term


to Medium-Term)

The Panel strongly supports the continued mapping of shoreline sensitivity and calls for this mapping to
be extended into Arctic areas.

Shoreline mapping should be conducted in areas where oil and gas exploration and development is
already occurring or is expected, e.g., Beaufort Sea and the Mackenzie Delta. The mapping should also
consider other human activities that would affect relative risk, such as mining, where large port facilities
have been proposed involving not only transport of the minerals out but fuel in, and increased commercial
shipping traffic through the Northwest Passage, which brings greater risk of fuel oil spills.

8.22.1.4 Gap: Shoreline sensitivity in inland Canada

Recommendation: Extend Sensitivity Mapping to Inland Freshwater Habitats (Short-Term to


Medium-Term)

The current lack of inland shoreline sensitivity information in the open literature is a serious gap affecting
the ability to apply risk-based preparedness and response planning. Prioritization of areas requiring
mapping could be based upon evaluation of the intensity of human use, the current knowledge of the
relative sensitivity of ecosystems and the availability of information from various sources. For example,
high-priority areas for mapping might include areas upstream and downstream of hydroelectric dams.

Sensitivity mapping should be coordinated and should use standard tools and analytical approaches. This
assessment must include integration of traditional knowledge and information about traditional uses of
resources.

8.22.1.5 Gap: Efficacy of oil spill responses for various types of oil

Recommendation: Conduct Experiments on Efficacy of Current and Emerging Oil Spill


Countermeasures (Short-Term to Medium-Term)

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 361
More data are required regarding the efficacy of current and emerging oil spill countermeasures for
various types of oil spilled over a range of environmental conditions. This will require support for the
conduct of field trials that incorporate controlled releases of oil. The consensus of scientists in oil spill
countermeasure research is that major advances in the development and validation of spill response
technologies are being hampered by our inability to conduct controlled field experiments with oil. Such
field experiments involving intentional releases of oil must use scientifically valid and statistically sound
experimental designs that use multiple replicate control and treated plots to enable calculation of
experimental error.

8.22.1.6 Gap: Sufficient information to conduct NEBA for Arctic spills

Recommendation: Compare the Risks of Various Methods and Intensities of Cleanup under Arctic
Conditions (Short-term to Medium-term)

Research comparing sites that have been subjected to various methods and intensities of cleanup to sites
with minimal or no cleanup should be conducted under conditions relevant to the Canadian Arctic,
building upon the experience in Prince William Sound. Experimental designs should include the range of
crude oils that may be produced and transported in the Arctic, as well as the specific fuel oils already
transported to Canadian Arctic communities. Well-controlled field experiments would be of great benefit;
therefore, the current prohibition of experimental oil spills in the field should be re-examined.

8.22.1.7 Gap: Finer-scale relative risk analysis that allows focused preparedness and response
planning

Recommendation: Conduct a Series of Focused Risk Assessments as a Follow-up to Recent


Assessments (Short-Term)

The Panel recommends that a set of focused relative risk assessments be conducted building upon the
Transport Canada marine spill results by focussing on high-sensitivity areas within each of the assessment
zones in order that preparedness and response plans can include explicit plans for the areas with the
highest potential consequences.

Follow-up risk assessments should be focused on specific areas at a finer scale than recent relative risk
assessment. In addition, areas can be selected based on environments with multiple anthropogenic and
natural stressors and measureable levels of impact and/or potential for future human disturbance. For
example, these areas might include portions of the Bay of Fundy, the Gulf of St. Lawrence and the Strait
of Georgia. Areas with high natural stressors, or which are sensitive to global factors that, at present, do
not have high levels of anthropogenic stressors might include the Labrador Sea, the MacKenzie River
delta and Hudson Strait. Once the priority areas for follow-up risk assessments have been identified,
methodology refinements should be applied. Recommended refinements are presented in bold in Sections
8.11, 8.12 and 8.13.

8.22.1.8 Gap: Limited data from past spills in Canada

Recommendation: Take Advantage of ‘Spills of Opportunity’ (Short-term and Extending to Long-


term)

If spills do occur, maximum advantage should be derived from the opportunity to study the fate,
behaviour, and effects of the spill in the short, medium, and long-term. Studies of the relative
effectiveness of response measures should also be part of a suite of investigations associated with all
significant oil spills in Canada. Pre-approved funding should be put aside for spills of opportunity that
would incorporate a combination of research and monitoring. Responsibility for each component of spill
studies must be clear and research and monitoring efforts must be well coordinated to eliminate

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 362
redundancy and to ensure that all required study components are implemented. It may be appropriate to
select areas that would not receive treatment; these ‘set aside’ areas would then serve as a medium and
long-term source of data for comparison of treated versus untreated sites.

8.22.2 Medium-priority research needs (medium-term and long-term)

8.22.2.1 Gap: Risks of residual oil in Arctic ecosystems

Recommendation: Evaluate the Consequences of Long-Term Residual Oil in High-Risk Marine


Habitats (Medium-term)

The consequences of long-term residual oil in high-risk marine shoreline, estuary, marsh and lagoon areas
both north and south of the 60th parallel require further research. Chedabucto Bay is a good candidate for
part of this evaluation. Other study areas should be selected on the basis of where consequences might be
highest. The results of the research would provide much-needed input to NEBA regarding the most
appropriate level of cleanup.

8.22.2.2 Gap: Risks of deposition to deep sea sediments

Recommendation: Evaluate the Consequences of Deposition of Oil to Deep Sea Sediments


(Medium-term)

Research into the consequences of deposition of oil to deep sea sediments in the Arctic, as well as deep
sea sediments south of the 60th parallel in Canada, would provide part of the required knowledge base for
support of decisions regarding offshore oil and gas exploration and oil transport. The current state of
knowledge is totally inadequate with respect to supporting confident assessment of risk to deep sea
systems.

8.22.2.3 Gap: The potential contribution of hyporheic flow to oil spill risk in rivers

Recommendation: Investigate the Role of Hyporheic Flows in Contributing to Exposure to Oil


(Medium-term)

Research is needed on the relative role of hyporheic flows in contributing to exposure of fish to
hydrocarbons after a spill. A range of experimental conditions using relevant concentrations and both
weathered and unweathered oil should be tested on salmonid species under conditions that represent
actual field conditions in terms of variables such as flow, temperature, substrate and dissolved oxygen.
Experimental results can be used to evaluate whether monitoring of hyporheic flows after spill events
would be feasible and justifiable.

8.22.2.4 Gap: Data supporting assessment of effects on ecosystem services

Recommendation: Investigate the Socioeconomic Impacts of Oil Spills in Support of Assessment of


Effects on Ecosystem Services (Medium-term to Long-term)

Measurement of socioeconomic impacts of oil spills as a first step in implementing an ecosystem services
approach to oil spill impact assessments. A standard set of indicators of socioeconomic effects should be
developed for marine and freshwater systems in Canada, in consultation with First Nations, industry,
academics and government agencies. These indicators should have explicit links to ecosystem services. If
possible, measurement of effects on culturally important ecological services should be included.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 363
8.22.3 Long-term research needs

8.22.3.1 Gap: Effects of oil spills on trophic dynamics

Recommendation: Conduct Research on the Effects of Oil Spills on Trophic Dynamics of Aquatic
Ecosystems

In the Canadian context, the current level of understanding of trophic dynamics in identified high-risk
offshore and inshore marine habitats requires evaluation in order that critical gaps in understanding are
identified. For example, there may be specific critical features of Bay of Fundy trophic dynamics at
specific locations (e.g., estuaries) that, if affected by oil spills, would have much greater consequences
than spills in other areas of the Bay. Alternatively, the pelagic trophic dynamics of the Bay of Fundy are
fundamental to maintenance of many highly-valued ecosystem services. Currently, it would be extremely
difficult to differentiate risks among inshore and offshore areas (and thus to guide cleanup and follow-up
monitoring programs).

8.22.3.2 Gap: Potential for population-level effects

Recommendation: Conduct Long-term Research on Effects of Different Oils on Populations of


Aquatic Biota

The Panel recommends that there be support for long-term research into effects of different oil types on
populations of aquatic biota, especially fish, marine mammals, and waterfowl. Such long-term research
requires stable funding and effective collaboration among disciplines such as environmental chemistry,
toxicology and ecology. Spills of Opportunity could provide the impetus for the initiation of long-term
research.

8.22.4 Operational preparedness needs

Standardized Sensitivity Mapping

National-scale relative risk assessments used for preparedness and response planning must be based upon
reliable and consistent data and analysis; therefore, standardized methodology for environmental
sensitivity mapping is an urgent requirement.

Post-Spill Monitoring Program Development and National Guidance

The Panel strongly recommends that national guidance regarding monitoring after oil spill events be
developed, involving consultations among industry, government, Indigenous communities and
community stakeholders. The guidance would be designed to produce information that is reliable,
adequate, credible and consistent. The guidance should include provisions for adjustment in response to
specific characteristics of the receiving environment. The guidance should also include requirements for
standard baseline datasets (tailored to specific receiving environments). The guidance could be divided
into two parts: 1) which information to collect without exception; and 2) which information gathering can
be deferred until the full scope of the spill and its potential effects are better understood. The guidance
should include information on data quality requirements, including determination of minimum sample
sizes, standard sampling protocols and laboratory quality assurance/quality control.

The Panel recommends the development of a national, consensus-based set of indicator species for each
of Canada’s major marine offshore and inshore zones (perhaps using the zones developed by WSP and SL
Ross (2014). Once these indicator species have been selected, coordinated research programs among
academia, industry and government should be developed which: a) compile existing information; b)
identify and prioritize critical data gaps; and c) conduct research to fill the priority data gaps.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 364
Standard guidance for monitoring the fate of spilled oil in all major environmental compartments of lotic
systems is required. This guidance would include measurement of oil in sediments, as noted above, but
would extend to measurement in water, pore water and biota.

Standard guidance is required for measurement of oil contamination of sediments in flowing waters in
Canada. This guidance should consider the range of lotic receiving environments and logistic constraints
common to remote locations. The guidance should include the de minimis level of investigation required
for decision-making regarding cleanup requirements and techniques, as well as for determination of
acceptable residual levels of oil.

The Panel supports the recommendations of SL Ross (2014) regarding monitoring programs in Canadian
aquatic systems. These recommendations are listed in Section 8.14.

Update and Refine Risk Assessment Methods for Use in Support of Project Applications or
Preparedness Planning

The Panel’s review of risk assessments performed as part of pipeline applications revealed opportunities
for improvement.

 Risk assessment scenarios should assume a more likely combination of factors (less than full-bore
rupture with a longer elapsed time to response) that are more relevant and comparable to actual spill
incidents. These scenarios would also be viewed as more credible by stakeholders sensitized by recent
spills that experienced long delays before shutdown.
 Detailed risk assessments conducted in support of applications for new projects such as pipelines and
production facilities should include winter conditions.
 Stochastic fate modeling for complex, sensitive receiving environments is recommended.
 The sites of interest used in the EEPP assessment merit being re-analysed using more sophisticated
analytical tools and approaches—particularly with respect to fate and transport and the level of
sophistication with respect to acute and chronic toxicity, shoreline effects and overland flow effects
with season.

Apply Research Results, Engagement with Indigenous peoples and Stakeholders and Economic
Analyses to Address the Longstanding Remediation Question “How Clean is Clean Enough?”

Effects on ecosystem services can be evaluated using data obtained from laboratory studies, large-scale
open-air experiments, long-term field research and baseline data surveys. These data should be compiled
and interpreted with the intention of determining acceptable, science-based chemical and biological
endpoints of spill remediation. Long-term monitoring of follow-on effects after experimental spills and
spills of opportunity would inform regulatory guidelines for cleanup endpoints.

The benchmarks for an acceptable level of residual oil should always include social and economic factors.
Social factors must include input from Indigenous peoples. Social and economic factors could include
residual effects on ecosystem services such as recreational or commercial fishing, drinking water sources
and parks or other protected areas.

Inclusive Preparedness Planning

Preparedness planning should consider information from Indigenous, provincial, territorial and municipal
sources in order that smaller-scale, higher-risk locations can be identified. Many of these areas have
already been mapped at a finer scale than presented in the Transport Canada relative risk assessment.
Provincial laws and regulations must be considered in preparedness to ensure harmonization and
avoidance of redundancy.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 365
Develop a Spill Risk Calculator Tool

The Panel suggests that a Spill Risk Calculator tool for Canada would be beneficial to preparedness
planners and responders. This tool can be based upon methods used for the assessment of oil spill risks in
Alaska (Reich et al. 2014).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 366
CHAPTER 9: SUMMARY CONCLUSIONS AND RECOMMENDATIONS

The Royal Society of Canada (RSC) Expert Panel on The Behaviour and Environmental Impacts of Crude
Oil Released into Aqueous Environments was asked to address the following questions:

1. How do the various types of crude oils, including diluted bitumens, compare in the way they
behave when mixed with fresh, brackish or salt waters under a range of environmental
conditions?
2. How do the various crude oils compare in their chemical composition and toxicity to organisms in
aquatic ecosystems?
3. How do microbial processes affect crude oils in aquatic ecosystems, thereby modifying their
physical and chemical properties, persistence and toxicity?
4. Are the research and oil spill response communities able to relate, with reliable predictions, the
chemical, physical and biological properties of crudes to their behaviour, persistence, toxicity and
ability to be remediated in water and sediments?
5. How should these scientific insights be used to inform optimal strategies for spill preparedness,
spill response and environmental remediation?
6. Given the current state of the science, what are the priorities for research investments?

During its deliberations the Panel considered accidental releases of conventional (light, medium and
heavy) crude oils and unconventional crude oils (including diluted bitumen), at exploration and
production sources or as they are transported between ports and storage sites, to refineries or to users.

Following a review of available information, the Panel concluded that large spills of crude oil into marine
or freshwater systems in Canada from oil production facilities, tankers, pipelines, rail and truck transport
that cause significant ecological damage are infrequent, and the probability of spills decreases with
increasing spill size (Chapter 1). The Panel noted that comprehensive statistical records for spills into the
marine environment of Canada were much more readily available than for inland spills into fresh water.
Thus, in the case of freshwater spills, the Panel had to search through individual investigative reports
from the Transportation Safety Board (TSB), Alberta Energy Regulator (AER) and the media.
Furthermore, as illustrated by the AER (2013) report on pipeline performance that provided data on spill
incidence and volume, spills of crude oil were not always distinguished from refined product spills.
Instead, statistics were presented as ‘hydrocarbon liquids’. This disparity in spill record keeping for the
freshwater environment of Canada should be resolved.

Despite the relative infrequency of crude oil spills into aqueous environments in Canada, the
consequences of spills into sensitive waterbodies can be substantial, not only economically but also
impacting human health, safety and the environment. This concern is commensurate with the key
questions posed by the sponsors and stakeholders.

With the anticipated growth of Canada’s offshore oil and gas industry and the production and transport of
unconventional oils, such as diluted bitumen, accidental releases of oil from offshore platforms, pipelines
and other sources were considered to be within the scope of this report. The Panel focused on the
Canadian environment. However, it reviewed and considered the applicability of case studies from the
United States and other countries to ensure full coverage of pertinent information. For example, current
knowledge on the effects of and response to subsurface blowouts was provided by the Gulf of Mexico
Deepwater Horizon (DWH) oil spill, and the consequences of and response to a release of dilbit spilled
into a freshwater environment was represented by the Kalamazoo River, MI, oil spill. The Panel has made
an effort to acknowledge the large number of research papers and environmental impact statements
recently published on oil spills, as well as major ongoing research initiatives. The conclusions and
recommendations made by the RSC Expert Panel in this report, under its remit to address the questions of

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 367
its Terms of Reference, are subject to revision as new findings from environmental impact assessments
and oil spill response technologies emerge.

Question 1: How do the various types of crude oils, including diluted bitumens, compare in the way
they behave when mixed with fresh, brackish, or salt waters under a range of environmental
conditions?

The Panel provided an overview on the chemical composition of petroleum hydrocarbons, the types of
conventional and unconventional oil (including dilbit) transported in Canada, their bulk properties, and
information on the processes that influence their weathering in the environment (Chapter 2). The Panel
concluded that the behaviour of these oils when mixed with fresh, brackish or salt waters is dependent on
their inherent physical and chemical properties, which are influenced by the environmental conditions at
the specific areas of concern. Our understanding and ability to accurately predict the behaviour of some
oils in the environment following accidental spills are still limited. This is due to analytical challenges,
incomplete experimentation and complex interactions among oil weathering processes, remediation
strategies and environmental conditions.

Each oil spill will be a unique combination of oil and receiving environment characteristics. However, the
review of oil spill cases revealed that some generalizations can be made (some having more uncertainty
associated with them than others).

Season and weather are dominant factors determining both the initial and ultimate fate and effect of all oil
types. The lighter the oil, the more it is affected by spreading and evaporation and the easier it is to treat
effectively, regardless of the environment in which it is spilled. The higher the ambient temperature, the
faster is the evaporation and dissolution and the more extensive is the spreading. Medium oils (e.g., most
conventional crude oils) are generally less volatile than lighter oils, which increases the potential for
impacts on waterfowl and wildlife. However, medium and light oils are amenable to relatively rapid
biodegradation. Heavy oils and diluted bitumens (e.g., Bunker C, dilbit, synbit and railbit), which have
fewer components that dissolve in water or evaporate to the atmosphere are more resistant to evaporation
and biodegradation and thus their potential long-term damage to the environment, waterfowl and fur-
bearing animals is greater. Cleanup of heavy oils and bitumens is extremely difficult for both marine and
inland spills because of their specific gravity, viscosity, flash point properties and high asphaltene
content.

In addition, the direct effect of weather conditions on turbulence influences the physical dispersion of oil
and the likelihood of successful dispersant use, especially for light and medium oils where some degree of
mixing-energy is required for breakup of oil into small droplets. However, turbulence can also increase
emulsification of oils, especially heavier oils, causing moussification or the production of water-in-oil
emulsions, which resist evaporation, spreading and biodegradation. Dilbit is an exception because it
resists the formation of water-in-oil emulsions. Turbulence is also an important factor in the formation of
oil-particle aggregates (OPAs; significant with respect to some diluted bitumens and their potential for
sinking).

High winds with associated wave action and currents can deposit oil onto windward shorelines and into
more protected areas such as lagoons. High waves, tides and seiches can result in oil burial along sandy
shorelines followed by re-exposure. Wind-driven surges can drive oil into the high intertidal zone or, in
freshwater systems, deep into marginal wetlands or riparian habitats. Spills occurring immediately after
high rainfall events can transport and subsequently deposit oil over wide areas of the floodplain of rivers,
resulting in significant stranding when flows subside. These effects are applicable to all oil types,
regardless of their properties.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 368
The persistence of residual oil along shorelines correlates with shoreline energy level (e.g., waves, tides
and ice), with oil persisting longer in low-energy and low-temperature environments. Cobble/boulder
beaches with fine-grained interstitial materials, lagoons with fine sediments, backwater channels and
shoreline marshes have been linked to the sequestration of conventional crude oil in a relatively un-
weathered state for long periods of time due to protection from physical weathering and/or biodegradation
in hypoxic conditions. Heavier oils are less amenable to this sequestration action as they tend to remain
on the surface and resist seeping into the interstitial area of the shoreline.

The fate of oil spilled into rivers is affected by river flow, river channel characteristics and sediment
characteristics and processes. High-flow conditions at the time of the spill will increase turbulence and the
presence of suspended particulate matter that, in turn, contribute to the formation of OPAs that will sink
and accumulate in deposition zones where they may be re-suspended by subsequent high-flow events.
Biological processes within sediments can also elicit the re-suspension of oil and associated sheen
production on the water surface. River channels that have been subject to bank hardening (e.g., by
placement of riprap or rubble) have higher velocity flow, leading to the potential for transport of spilled
oil farther downstream and/or stranding of oil among the bank materials. There is also a potential for
entrainment of oil into porous bed sediments of shallow, high gradient rivers by surface-ground water
interactions (hyporheic flows), although this possibility has not yet been evaluated in detail.

The fate of oil spilled into lakes is affected by lake morphology and mixing characteristics. Conventional
crude oils spilled into shallow, wind-swept, well-mixed lakes will be subject to rapid dispersion if the
spill occurs during open-water seasons. Oil, especially the heavier types, spilled into deeper lakes with
thermal stratification may, in part, be sequestered in cold, low-oxygen, deepwater sediments with
subsequent low biodegradation rates if OPAs are formed and subsequently sink.

Brackish and saline waters have greater density than fresh water, which affects the sinking of oil. In
theory, as a dilbit loses its light components to evaporation, it may reach sufficient density to sink even in
seawater, especially if it forms OPAs. The longer-term fate and effects of oil deposits within marine
sediments is the subject of current research.

Question 2: How do the various crude oils compare in their chemical composition and toxicity to
organisms in aquatic ecosystems?

The Panel provided an overview of what is known about oil toxicity and ecological impacts of oil spills to
aquatic environments based on previous review articles and current literature (Chapter 4). Numerous case
studies (e.g., Exxon Valdez) have documented various levels of habitat destruction that have resulted from
oil spills. Although an overview is provided on the toxic effects of oil spills on various biota (i.e., plants,
invertebrates, fish, reptiles/amphibians, birds and mammals) within the Panel’s report, a major focus was
given to fish, which were the subject of the greatest number of research papers. This is not surprising as
the concern over oil spills is largely focused on impacts to fisheries due to their socioeconomic
importance. Furthermore, many of the interactions between oil chemistry and its toxicity to fish may also
apply to other species.

Studies with fish have shown that acute lethality is associated primarily with low molecular weight
(LMW) components of oil, such as monoaromatics (e.g., BTEX), diaromatics (naphthalene and alkyl
naphthalenes) and short-chain alkanes (< C12). LMW petroleum hydrocarbons are sufficiently
hydrophobic to partition rapidly into lipid membranes and to cause mortality by narcosis. Typically, as oil
weathers following a spill, most narcotics hydrocarbons are lost within two days by volatilization,
biodegradation and dilution. As a consequence, lethal concentrations in water are not maintained for
prolonged periods unless the discharge of oil is continuous. Thus, observed fish kills are typically brief
and localized because of the rapid loss of the acutely lethal LMW oil components through dilution and
weathering. However, extensive fish mortalities have been observed in rivers where a point source of oil

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 369
was rapidly transported downstream before significant weathering occurred (e.g., pipeline break, Pine
River, BC).

The acute toxicity of hydrocarbons to aquatic biota can be estimated from octanol-water partition
coefficients (Kow), the ratio at equilibrium of the concentrations of a compound in n-octanol (a water-
insoluble phase that mimics tissue lipids and floats on water) to the concentrations in the underlying
water. Mortality typically occurs when tissue concentrations reach a ‘critical body residue’ or median
lethal dose (CBR or LD50) of approximately 5 mmoles/L of tissue lipid. These predictable relationships
have been used as the basis for mathematical models for assessing the risk of acute lethality following
potential oil spills from proposed developments, such as pipelines, and for assessing the potential lethality
of hydrocarbon mixtures measured during actual oil spills.

Chronic and sublethal effects of oil on fish are associated with either chronic exposures (e.g., oil stranded
in sediments) or with delayed or lingering effects of acute exposures. Over the lifetime of a species,
exposures to oil may occur via respiratory uptake of hydrocarbons from water (most common), direct
contact with oil droplets (an emerging area of research) or the diet and maternal transfer to eggs (not well
studied). Effects range from initial cellular genetic and molecular responses to long-term organismal
impacts on rates of reproduction, growth, disease, mutations/cancer and survival.

In contrast to acute toxicity, 3-5 ring medium and high molecular weight (MMW and HMW,
respectively) polycyclic aromatic hydrocarbons (PAHs), including aromatic heterocycles, such as alkyl
dibenzothiophenes, are thought to be primary causes of chronic toxicity to fish. Embryos appear to be the
fish life stage most sensitive to oil exposure, presumably due to interference with the large number of
genes that must be up-regulated or activated at precise stages during embryonic development for normal
development to occur. Most studies report the effects of waterborne oil, usually by reference to the
concentrations of total PAHs (TPAHs) and total petroleum hydrocarbons (TPHs) in test solutions that
affect 50% of test organisms. The EC50 and LC50 values for chronic exposures have been reported to
range from 0.3 - 60 µg/L TPAH and 0.03 - 11 mg/L TPH. Because of their lower volatility, alkyl PAHs
persist in water longer than monoaromatics continuing to partition from oil into water over weeks to
months if residual oil persists.

It might be expected that the relative risks of acute and chronic toxicity of different oils could be
predicted by simply ranking them in order of the concentrations of their different constituents.
Unfortunately, the expected correlations were not observed in studies with marine fish and invertebrates.
Toxicity may be confounded by the relative proportions of different alkyl PAHs, the presence of
unidentified compounds (including additives) or by generation of polar hydrocarbon derivatives during
weathering of the oil.

Differences in the relative proportions of different PAHs in oil may influence the toxicity of diluted
bitumens, such as the product dilbit. Dilbits are prepared from bitumen, a highly-weathered oil depleted in
alkylnaphthalenes, which are the alkyl PAH of lowest chronic toxicity. Thus, dilbit contains a higher
proportion of the more toxic 3- to -5-ring alkyl PAH mixture than do conventional crude oils, so that
chronic toxicity at a given concentration of TPAH may be greater than expected. Only one report has
been published on the chronic embryo toxicity of dilbit. In this study with Japanese medaka, chronic
toxicity across an array of crude and refined oils suggested that dilbit may be slightly more embryo-toxic
than average. Until equivalent data for other species have been reported, any conclusions about the
perceived toxicity of dilbit may be confounded by species differences in sensitivity to oil. There is some
evidence dilbit elicits acute toxicity in benthic invertebrates and that its toxicity varies with sediment
characteristics.

It is often difficult to express quantitative relationships between PAH concentrations in oil and toxicity
because the physical characteristics of oil modify toxicity by controlling droplet formation and the rate of

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 370
partitioning of hydrocarbons from oil to water. The proportion of waxes, resins and asphaltenes will
control the physical characteristics of oil, which in turn affect the dispersability and biodegradability of
oil in water. Thus, relative to light oils, heavy oils including dilbit are more difficult to disperse
mechanically in water, are less likely to form fine droplets, biodegrade more slowly, and may appear less
toxic than light oils because their toxic constituents are less bioavailable.

The Panel’s report includes tables and references that contain compositional and physical data for crude
oils that represent petroleum types transported in Canada and globally. Based on current analytical
capabilities, the petroleum hydrocarbons produced and/or transported within Canada appear to fall along a
chemical continuum spanning condensates to bitumen. Condensates are considered acutely toxic due in
part to elevated BTEX concentrations. Conventional oils with lower proportions of BTEX can exhibit
chronic toxicity associated with alkyl PAHs and resins.

The results of studies to date suggest that the direct toxic effects of dilbit in both freshwater and marine
ecosystems do not seem to differ dramatically from other oils. However, the binary nature of dilbit
currently confounds extrapolation of toxicity from its individual components: diluent and bitumen. The
acute toxicity associated with the diluent may be dispelled by evaporation, and the heaviest residues (the
asphaltenes) in the bitumen are likely biologically inert. However, the likely extent of diluent evaporation
(conversely, the biological availability of diluent in weathered dilbit) is currently unknown. Furthermore,
the toxicity of resins and/or products of photooxidation is poorly understood.

The impacts of spilled oil on aquatic species depend as much on the extent of their exposure to oil as on
the inherent toxicity of oil. Exposure to oil will vary with the interactions among the fate and behaviour of
spilled oil and the habitat characteristics, behaviour and physiology of different species. Research is
needed to identify those species that will be most exposed to oil with extreme properties, including very
light (e.g., Bakken crude) and very heavy oils (e.g., HFO), and the ecological factors that will enhance
exposure (e.g., turbulent mixing of rivers or weathering). Research is also needed to better understand the
differences between dilbit and Bakken light crude versus conventional crude oils in terms of their acute
and chronic toxicity. The hydrocarbon composition of various oils can play a major role in explaining
these differences. Although acute lethality can be attributed to LMW components of oil and chronic
toxicity to the alkyl-substituted 3- to 5-ring PAHs, research is needed to identify the most toxic
constituents of oil, particularly polar hydrocarbons generated by weathering, biodegradation or
photooxidation that are not often analyzed in oil or water.

It is important to note that most vertebrate species (e.g., fish, birds, mammals) can metabolize and excrete
hydrocarbons. In contrast, invertebrate filter feeders do not metabolize or excrete petroleum hydrocarbons
quickly; thus, contaminated zooplankton can accumulate significant quantities of droplet oil. Although
contaminated invertebrates can contribute to the dietary exposure of predators, it appears that petroleum
hydrocarbons do not typically biomagnify to high concentrations in food webs.

Question 3: How do microbial processes affect crude oils in aquatic ecosystems, thereby modifying
their physical and chemical properties, persistence and toxicity?

The Report outlines in Chapters 2, 3 and 6 how microbial communities in aquatic environments interact
with spilled oil to change its properties, behaviour and effects on other biota. Interactions that affect
physical properties include stabilizing emulsions via cell attachment to oil:water interfaces, enhancing oil
dispersion in water through secretion of biosurfactants and biodegradation.

The chemical composition of oil dictates its susceptibility to biodegradation (i.e., the proportions of
biodegradable alkanes and aromatics versus recalcitrant steroids, resins and asphaltenes). In competition
with abiotic evaporation and dissolution processes, biodegradation selectively converts certain lighter
alkanes and aromatics to gases (CO2 and/or methane), water and biomass while leaving the heavier,

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 371
complex oil components as residues. This altered chemical composition changes the physical properties
of the residual oil, making it more dense and viscous than the original. In turn, this physical change
progressively reduces further biodegradation of the residual oil since biodegradation is also influenced by
the physical state of the oil (e.g., viscosity that affects the oil:water interfacial surface area available for
microbial attachment and attack). Biodegradation, being an enzymatic process, is directly affected by
temperature, and because microbial cell growth is the ultimate outcome of biodegradation, its rate is
affected by availability of growth-promoting nutrients such as soluble nitrogen and phosphorus sources.
Even under ideal environmental conditions, biodegradation of the multitude of components within a crude
oil is a relatively slow process that may take weeks, months or even years to remove a significant mass of
spilled oil, and oil that is buried in anaerobic sediments may persist for decades or longer. A residue of
original petroleum components, including asphaltenes and resins, will always remain, in addition to
partially degraded hydrocarbons and metabolic end products (e.g., CO2, CH4, H2S), the mass of which
will depend on the interplay between environmental conditions, petroleum chemistry and microbial
community composition.

The predominance of aerobic versus anaerobic conditions dictates which enzymatic pathways are
effective, the rate of microbial attack and the end products. Oil biodegradation is faster and more efficient
under aerobic conditions, such as in a well-aerated water column or in the uppermost layer of sediment,
but it is slower in environments dominated by anaerobic conditions, such as buried sediments. The range
of susceptible hydrocarbons under anaerobic conditions appears to be more limited than with aerobic
degradation, and the end products and residual oil composition therefore differ from aerobic processes.
For this reason, wetlands, salt marshes and mud flats that have significant anaerobic strata are among the
most sensitive and difficult ecosystems to clean after an oil spill. Bioremediation may be the best
approach for wetland cleanup because it is among the least intrusive strategies for these sensitive
environments. However, even bioremediation requires some human intervention that may cause
additional harm to the contaminated site by trampling of oil into the anaerobic sediment, which may
diminish bioremediation of the site and promote oil persistence in situ. Natural attenuation may be the
cleanup method of choice if trampling cannot be avoided. The beneficial interaction of hydrocarbon-
degrading microbes and plant roots in soils and wetlands (phytoremediation) may enhance biodegradation
and decrease persistence of buried oil.

The types of microbes present at a spill site dictate the lag time before biodegradation begins, the
subsequent rate and extent of biodegradation and hence the behaviour and persistence of the residual oil.
Hydrocarbon-degrading bacteria are ubiquitous, but they tend to be present in higher proportions in
chronically oil-impacted sites because they have an advantage over microbes that cannot utilize
hydrocarbons. For the same reason hydrocarbon-degraders may flourish after an oil spill even in a pristine
site, given favorable conditions. Thus, the history of a spill site may indirectly influence progression of
bioremediation or, conversely, oil persistence. Extensive and rapid rate of biodegradation requires the
metabolism of a consortium of hydrocarbon degraders working in conjunction with non-hydrocarbon
degrading microbes.

Given these general factors, it is evident that light oils are inherently more biodegradable than heavier oils
and leave lighter residues. In contrast, only small proportions of heavy crudes are readily biodegradable,
and the residues will persist even with active intervention. In the case of dilbit, it is not yet clear whether
all of the diluent is actually available for biodegradation, is protected by virtue of molecular association
with the bitumen or is simply unavailable to microbes because of the limited surface area of viscous
weathered dilbit. Considerable research is required to address this knowledge gap under different
environmental conditions.

In the process of selectively biodegrading petroleum components, microbes can reduce or increase the
toxicity of an oil spill. For example, very light oils (e.g., jet fuels and gasolines) contain high
concentrations of acutely toxic compounds that are susceptible to biodegradation. Although such light

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 372
compounds may be quickly lost by abiotic processes (particularly evaporation and dissolution),
biodegradation may reduce acute toxicity if they are retained in an environment (e.g., in deep water or
buried in sediment or trapped under ice). In some conventional oil spill scenarios natural attenuation can
effectively lower the risks to aquatic life and human consumers of fish and shellfish, despite persistent
residues, because: a) biodegradation may produce primarily non-toxic residues; b) the residual oil is not
being released back into the water in amounts sufficient to significantly affect aquatic life; or c) residual
oil compounds (e.g. asphaltenes) are not bioavailable. This was well illustrated by the Arrow tanker oil
spill case study. Microbial degradation by indigenous bacteria (i.e., natural attenuation) contributed to the
reduction of toxic fractions within the heavy Bunker C fuel oil stranded within the beach sediments at
some sites. At present, information is insufficient regarding the effectiveness of monitored natural
attenuation as an oil spill countermeasure for a dilbit spill in the aquatic environment.

Conversely, in unfavorable conditions microbes may be unable to completely convert hydrocarbons to


gases, water and biomass. Instead, partially oxidized metabolites that are more polar (more water-soluble)
and more mobile in an aqueous system may accumulate. Because of the diverse hydrocarbon isomers
present in petroleum, an enormous number of possible metabolites could be produced. The toxicity of
most is untested either as pure compounds or, more realistically, as mixtures of polar metabolites. In some
cases the metabolites generated by microbial attack are known to have carcinogenic potential. By analogy
to the resins fraction, many are likely to have chronic toxic effects. Additional research is needed to
address this question of chronic toxicity due to incomplete microbial metabolism of crude oil
components.

Another factor affecting chronic toxicity of the oil itself (versus its metabolic products) is that as lighter
components of the oil are biodegraded, the proportion of alkyl PAHs and resins in the residual oil
increases, thus increasing the chronic toxicity associated with these components. In addition to these
chemical effects, the increase in oil density and viscosity due to biodegradation of light fractions can
increase the likelihood of detrimental physical impacts on higher organisms including plants.

The mechanism by which microbial processes affect dilbit toxicity has yet to be fully resolved because
the binary nature of dilbit confounds simple extrapolation of toxicity from the individual components.

Question 4: Is the research and oil spill response community able to relate, with reliable
predictions, the chemical, physical and biological properties of crudes to their behaviour,
persistence, toxicity and ability to be remediated in water and sediments?

The environmental behaviour of unconventional oils and blended oils currently cannot be predicted with
confidence, which affects spill response planning and decisions. Each oil spill is unique, due to the
specific chemical nature of the oil and the complex interaction with different environmental and climate
factors. Thus, to account for the environmental variables that might be experienced in a dilbit spill, a
multitude of combinatorial trials would be required before empirical rules of its behaviour can be
developed.

The toxic effects of oil are well described for a limited array of aquatic organisms, most notably acute
toxicity in fish embryos due to their high sensitivity, inability to avoid exposure and the obvious concerns
over the impacts of oil spills on the fisheries. Much less is known about the effects of oil on other
responses of fish (e.g., sexual maturation, reproduction, behaviour) or on other species, in part because
research has not been focused and systematic (e.g., see review of toxicity to algae, Chapter 4) or it is
simply lacking. In addition, the interactions of oil exposure and toxicity with environmental stressors,
including temperature, salinity, oxygen concentrations, pH, spawning migrations and other anthropogenic
contaminants, would be important variables in this research. It is critically important to determine if acute
sublethal exposure of aquatic organisms (e.g., fish embryos) to oil causes delayed or lingering toxicity
evidenced as a reduced abundance in the years following an oil spill.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 373
A major impediment to quantitative predictions of oil toxicity is the quality of toxicity data for risk and
impact assessments. Problems include a lack of methods to distinguish dissolved from droplet oil, the
high cost of analytical methods to adequately characterize exposures, and time-varying concentrations in
toxicity tests. The application of toxicity data in ecological risk assessments is also limited by the
availability of models capable of linking acute and chronic toxicity to the chemical characteristics of
hydrocarbons. The current models that predict acute lethality reasonably well do not apply to situations
where concentrations of oil in water vary over time, as would be the case during an oil spill. The routine
applications of these models also do not generate estimates of the error of calculated toxicities, despite
considerable variations between observed and predicted toxicities for individual compounds. Models to
predict chronic toxicity have not been developed from data on chronic toxicity, but rely on the incorrect
assumption of a constant ratio between acute and chronic toxicity.

Ecological impact assessments of spilled oil, especially dilbit, are also limited by a lack of ecological
research in oil-contaminated ecosystems. Long-term time-series research ranging from days to years on
‘spills of opportunity’ or experimental oil spills is needed to establish mechanistic and statistical links
among hydrocarbon exposure of aquatic species, biomarker responses, the prevalence of signs of toxicity
and pathology, the distribution, abundance and production of individual species, and ecosystem structure
and function. In particular, research is needed on the fate, behaviour and effects of oil in freshwater
ecosystems using designs and methods that reflect differences in spatial scales, processes, water
movements and biological communities relative to marine ecosystems.

A knowledge gap remains in understanding the potential impacts of oil, or chemically-dispersed oil, in
deepwater and Arctic ecosystems where dispersant use is contemplated. The complex interactions among
extreme low temperatures, the unique physiologies of coldwater species, the solubility kinetics of
hydrocarbons in water, exposure times, pharmacokinetics of hydrocarbon uptake and toxicokinetics for
different mechanisms of acute and chronic toxicity of oil must be investigated. With respect to freshwater
and terrestrial ecosystems, our understanding of the impacts of spilled oil on permafrost ecosystems and
northern wetlands, which are characterized by complex water flows and plant and animal species highly
adapted to harsh conditions, is severely lacking.

Chronic impacts may also result from a cascade of effects originating with acute exposure to the spill
and/or to lingering effects of residual oil. For example, mortality induced damage to the structure of a
population or community (e.g., a killer whale pod) will alter food web dynamics (e.g., elimination of a
predator leading to increased abundance of the prey species and increased feeding pressure on primary
producers). Toxicity can also cause indirect effects, such as energetic costs associated with the
metabolism of PAHs and other hydrocarbon compounds, that result in lower body weight and increased
mortality or higher susceptibility to predation. Such effects and their ecological significance are not fully
understood and hence not properly considered in risk assessment models.

In terms of modeling, the oil spills of interest for the community are those that cover large areas because
the smaller ones can be more or less contained and treated. Therefore, the thrust of modeling oil transport
from a response point of view has been on parameterizing small-scale processes while attempting to
capture the bulk of the large-scale transport. In other words, the small-scale information has been
sacrificed in favour of the large-scale. Unfortunately, this means that most oil spill models designed for
response, such as NOAA’s GNOME, are not designed to provide concentrations that could be readily
used for predicting toxicity effects. For example, a model that conserves mass and uses a computational
block of 1 km x 1 km would not distinguish whether the mass of oil is spread within that block or whether
it is localized within a 10 m x 10 m sector of that block. While the mass is the same in both cases, the
latter provides a concentration that is 100 times greater. Also, oil spill models developed for response
cannot be used to design the optimal parameters for applying dispersants. In contrast, models designed for
research are sufficiently sophisticated to allow their use in toxicity studies and for evaluating remediation
scenarios. Nevertheless, as modeling is becoming more acceptable as a predictive tool, sophisticated

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 374
models are becoming more common. The goal is to achieve effective communication between the
response community and the academic community that can lead to collaboration, such as occurred with
the droplet model from jets VDROP-J developed in academia and currently being adopted by NOAA in
their model ADIOS2.

To support the application and accuracy of predictive models following spill events, research and
implementation of autonomous, remotely operated detection devices (e.g., AUVs and ROVs) and other
remote sensing technologies should be encouraged, as well as the development of new or refined
biomarkers and fingerprinting methods for conventional and unconventional oils (including dilbit) and oil
spill treatment agents such as dispersants, in water, sediment and biological materials recovered from
freshwater and maritime systems. Uncertainty, in the form of incomplete information, errors in sampling,
subjective judgement, random variations of and dynamic interactions among operating factors,
approximations and assumptions in measurement, and changes of environmental conditions, will always
be a challenge in oil spill response operations. Reduction in the level of uncertainty is needed in the
measurement of environmental factors, spill modeling and response decision-making and implementation.

The Panel recommends the conduct of biodegradability studies on conventional and unconventional oils
commonly transported in Canada, especially dilbit, to delineate the maximum biodegradation rates
expected under actual environmental conditions. Comparison to multiple control conditions will be
essential for rigorous interpretation of the results. A controversy still remains over the use of oil
dispersants for spill response. Consideration should be given to the use of mathematical decision-making
tools to optimize the dose, logistical and operational processes for their application, as well as the
development of more effective and eco-friendly dispersant formulations.

The existing models for simulating response processes (e.g., booming, in situ burning, skimming,
dispersion and bioremediation) and predicting their effectiveness (individually and collectively) are
limited in terms of accuracy and adaptability to different environmental conditions. This is especially true
for cold waters and Arctic environments. Furthermore, once the response has ended, it becomes a matter
of evaluating the effectiveness of natural attenuation (i.e., a reduction in contaminant levels and habitat
recovery) and risk from residual oil. The development of standardized methodologies and targeted post-
spill monitoring programs are needed to accurately determine the levels of response efficacy and the risks
of residual oil.

The challenge exists on how to determine what level of preparedness is appropriate for each area of
response and of concerns across Canada. Despite the importance of oil type, the Panel concluded that the
overall impact of an oil spill, including the effectiveness of an oil spill response, depends mainly on the
environment and conditions (weather, waves, etc.) where the spill takes place and the time lost before
remedial operations.

In terms of prevention, there is a need to assess the current level of preparedness in light of relative risks
in order to identify major geographic areas of concern and those with deficiencies in response capability.
The risk assessment commissioned by Transport Canada should be expanded from marine spills to
freshwater environments. In addition to cumulative impacts of response actions, socioeconomic factors
should be more effectively reflected in the net environmental benefit analysis (NEBA) process to
minimize impacts on a region’s environment and economy and the wellbeing of its people. In terms of
risk assessments, it is important to stress that our ability to distinguish the effects of oil spills (even on the
scale of the DWH) depends upon the availability of pre-spill baseline information, an understanding of
natural variability and an understanding of the effects of multiple anthropogenic stressors. The risks of oil
spills must also be evaluated in the context of large-scale oceanic and climatic phenomena and changes
that are occurring at unprecedented rates.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 375
In the marine environment, preliminary data suggest that the early post-spill behaviour of dilbit would not
differ substantially from the behaviour of conventional crude oil. However, dilbit fate and behaviour
depends on various and sometimes unpredictable environmental factors, such as the level of mixing
energy and level of contact with suspended particles. Thus, questions remain about our level of
confidence in selecting and developing suitable response strategies and cleanup equipment for both
contingency planning and response management. To reduce the level of uncertainty, rapid and effective
response operations are paramount. Regardless of oil type, rapid response will limit the spatial extent of
floating oil, prevent extensive contact with soils, sediments and organic materials along the spill flow
path, and minimize shoreline stranding. The nature of the crude oil is less important than the rapidity and
efficacy of the response. For example, rapid on-water response by an oil recovery unit to the spill of
diluted bitumen from the punctured Kinder Morgan pipeline in Burnaby, BC, minimized subsequent
environmental damage. However, if such a spill had occurred in a more remote area, the consequences
might have been more substantial.

Cold and harsh environments in northern Canada and the Arctic are usually characterized by a wide range
of wind speed and direction, limited visibility, low temperature, rough water surface, ice coverage, etc.,
posing substantial challenges for oil spill prevention, preparedness and response. To protect the habitat
and its resources within the Arctic and sub-Arctic regions, vulnerability and risk analyses are needed to
support contingency planning and response operations. Currently, there is a drive within Canada and
internationally (e.g., Northern Region Persistent Organic Pollution Control Laboratory – Memorial
University; Churchill Marine Observatory – University of Manitoba; Arctic Joint Industry Partnership –
International Association of Petroleum Producers) to increase understanding of the fate and behaviour of
oil in water and ice conditions (solid, slush and frazil ice) and during active periods of formation and
breakup of annual and multi-year ice, as well as the impacts on ecosystems. This experimental
information on the effect and efficiency of different response methods under Arctic conditions is urgently
needed to improve the ability to forecast and manage oil spill response operations in a more timely, eco-
friendly and cost-effective manner.

Most of the knowledge of oil spill response technologies has been developed through results from studies
in the laboratory, mesocosm test systems and actual case studies (with limited controls and treatment
replication). It is the consensus of the Panel and general scientific community that controlled field trials
are needed to advance the development, validation and acceptance of operational spill response strategies,
especially for subsurface blowouts, Arctic oil spills and freshwater shorelines.

Question 5. How should these scientific insights be used to inform optimal strategies for spill
preparedness, spill response and environmental remediation?

Baseline data are needed to distinguish oil impacts from natural environmental variability and for the
selection of spill response options to be taken (e.g., dispersant use). The identification of operational
endpoints for oil spill response operations raises the difficulty of addressing the question “how clean is
clean?” Usually this is a site-specific question because habitat recovery depends largely on a variety of
species at risk identified through baseline surveys.

Anticipating the challenges of oil spill response in remote, northern areas and cold waters and planning
for each contingency would help decrease the likelihood of undue delays and inappropriate response
techniques. Recovery of oil spilled in the Arctic can be expected to be relatively ineffective; therefore,
prevention must remain a major focus.

Delays in response typify oil spill case studies. Common causes of delays include lack of preparedness
and response capabilities, deficiencies in regulatory oversight and lack of communication and
coordination among government agencies, Indigenous communities and industry. Involvement of

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 376
Indigenous communities in decision-making regarding cleanup and monitoring is occurring and will
continue to contribute to the baseline knowledge required to support focused preparedness planning.

Human intervention often can cause more harm than the spill itself. Aggressive shoreline cleanup
methods, such as high-pressure washing and physical removal of oiled vegetation and debris along the
shore of lakes and rivers, can impart substantial negative effects on biological communities, including
indirect and cascade effects that can delay recovery. Thus, responders should be sure to consider the
impacts of their response actions will have on the ecosystem even from the contingency planning stage. In
some cases, monitored natural attenuation (natural recovery) may be the best response option for stranded
or submerged oil since the negative effects of cleanup may be significant. Consultation not only with
Aboriginal communities but also with ecologists knowledgeable in the contaminated environment is
essential for protecting these important habitats. Furthermore, NEBA should also consider sociological
factors because of public concerns over matters such as the return of a high amenity sites (e.g. highly-
valued beaches and watercourses) to their pre-spill state.

Research is needed to further understanding of the effects of spilled oil in Arctic and permafrost
ecosystems and how best to develop appropriate response strategies that mitigate the damage without
further harm. Little has been done to advance knowledge of how to deal with a subsurface blowout in ice-
covered marine environments. While the National Energy Board has identified the need for additional
precautionary measures (e.g., same season relief well capacity), further research is needed to aid in
development of oil detection and response decision support. With the advent of Arctic drilling for oil and
hydrocarbons, improved methods are needed to detect and monitor the behaviour and transport of spills
on, in and under ice and within the water column in the event of a subsurface deep sea blowout. In
addition, oil interactions with permafrost and spring melt are associated with a number of challenges and
unknowns that include site access for spill response, rapid spreading of oil during freshet, slow rates of
weathering and how hydrocarbons interact with ice and suspended particulate matter.

Current technologies, equipment and response personnel are needed to be fully capable of effectively
responding to major inland or offshore spills in northern Canada and Arctic regions. To advance
operational spill response guidelines in these extreme environments, improvements in the understanding
of the ecological effects of various oil spill cleanup measures need to be developed.

More research is needed to rapidly determine the highest-risk combination of environmental


characteristics when dealing with dilbit. For example, further studies are needed to understand under what
circumstances a dilbit product is most likely to sink and its weathering would be slowest. In addition,
more research is needed to understand the behaviour of dilbit and other heavy fuel oils in a range of
environments, including near-shore and offshore marine, estuarine and freshwater lakes, rivers and
wetlands, under various combinations of climate, water chemistry and biological community conditions.
This new information will help to optimize the selection of current spill response strategies and to support
the development and validation of emerging technologies, including biodegradation of mousse (water-in-
oil emulsions). Details on the factors that influence mousses and tar balls are needed to better understand
their formation, properties and treatment.

Based on the Panel’s review of the physical and chemical properties of petroleum hydrocarbons, the
overall behaviour of both conventional and unconventional oils currently transported within Canada fall
along a continuum for which existing oil spill technologies have been used. Unfortunately, a full
understanding of the fate and transport of conventional and unconventional oil, such as dilbit, and
associated impacts under different environmental conditions is insufficient. This compromises our ability
to select and develop suitable response strategies and cleanup equipment.

In terms of priorities, just as conventional oils have been studied for years, greater focus should be given
to dilbit, heavy fuel oils, emulsified oils and possibly tar balls in a range of environments, with special

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 377
emphasis on identified high-risk coastal areas and cold northern parts of Canada. The results of these
studies will help to identify and predict the level of impact following a spill, aid in the selection of the
optimal spill response strategies, support the development and validation of emerging technologies and
provide needed assistance in the assessment and understanding of the risks following a high impact spill.

Most literature on oil persistence is from marine environments. More information on oil persistence in
riverine, lacustrine and mudflat environments would be beneficial for risk assessments and response
operations. Also, more information is needed on high gradient, high energy rivers and hyporheic zones
connected to groundwater resources where there is a potential for mixing of oil into water by physical
dispersion and widespread distribution.

Emphasis has recently been given to in situ ‘green’ technologies for the remediation of oil-contaminated
sites that have higher public acceptance. Although much progress has been made over the last decade,
green spill response technologies, such as phytoremediation (a technique based on enhanced oil
biodegradation rates by plants and their associated rhizosphere microorganisms) and biodispersion (by
potentially eco-friendly dispersants based on biosurfactants which are surface-active chemicals produced
by microorganisms), should be tested and validated as cost-effective cleanup options. Although anaerobic
biodegradation of hydrocarbons is still poorly understood, methods to accelerate the process are being
considered for the treatment of oil sequestered within riverine, lacustrine, estuarine or marine sediments
where it could persist for decades. Innovative methods (such as introduction of bioventing, air sparging or
dissolved air flotation as response options) to stimulate anaerobic processes in the environment are
needed to accelerate the rate of oil-impacted sediment remediation and habitat recovery.

Responses to oil spills usually consist of a series of dynamic, time-sensitive and complex processes
subject to time and resource limitations, as well as environmental and technical constraints. The
effectiveness of the response relies on the efficient use of available information (oil, weather, wave, etc.)
and resources (devices, manpower, money, etc.) in response decision-making and implementation.
Inadequate response decision support is one of the major challenges that limit the efficiency of current
response practices. Due to limited attention and investment, existing response decision support systems
are rare and lack dynamic and interactive support from other assessment tools (e.g. risk analysis, impact
assessment, spill modeling, process simulation, etc.) and field validation. These are especially true for the
Arctic regions and waters where harsh environmental conditions limit the time period available for
response.

Question 6. Given the current state of the science, what are the priorities for research
investments?

The behaviour, fate and effects of spilled oil are highly site-specific and decisions on oil spill cleanup and
on assessing environmental and socioeconomic impacts are highly dependent on science. As illustrated in
this RSC Panel report, the science is often lacking or highly uncertain due to the large number of
variables that affect oil behaviour, fate and effects. There is an urgent need in Canada to develop science-
based guidance and protocols for oil spill impact, risk assessments and cleanup. Guidance documents
should provide scientific information and technical data on the various types of hydrocarbons transported,
their potential effects and baseline information on potential sites exposed to high-risk -- material essential
for the application of science-based decisions in oil spill response operations. This initiative requires a
national level of coordinated scientific research in both the laboratory and the field (including
experimental field trials with controlled releases of oil) that involves both support and funding from
Indigenous communities, federal and provincial government agencies, private sector industries, academia
and the public. The core needs of this proposed program are listed within seven high-priority research
themes identified by the RSC Panel to improve spill prevention and the capability and effectiveness of oil
spill response operations in Canada:

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 378
1. Research to better understand the environmental impact of spilled crude oil in high-risk and poorly
understood areas, such as Arctic waters, the deep ocean and shores or inland rivers and wetlands.

2. Research to increase the understanding of effects of oil spills on aquatic organisms at the population,
community and ecosystem levels.

3. A national, priority-directed program of baseline research and monitoring to develop an


understanding of the environmental and ecological characteristics of areas that may be affected by
oil spills in the future and to identify any unique sensitivity to oil effects.

4. A program of controlled field research to better understand spill behaviour and effects across a
spectrum of crude oil types in different ecosystems and conditions.

5. Research to investigate the efficacy of spill responses and to take full advantage of ‘spills of
opportunity’.

6. Research to improve spill prevention and develop/apply response decision support systems to ensure
sound response decisions and effectiveness.

7. Research to update and refine risk assessment protocols for oil spills in Canada.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 379
APPENDIX A. ROUTINE LABORATORY ANALYTICAL METHODS FOR CHEMICAL CHARACTERIZATION
OF CRUDE OIL CONTAMINATION IN THE ENVIRONMENT

Crude oils are complex mixtures of thousands of chemicals; various combinations and proportions of
these chemicals impart different properties on the oils (Chapter 2). Analysis of such mixtures, particularly
complex high molecular weight (HMW) components, including resins and asphaltenes, remains
technically challenging. No single method is sufficient to reveal and identify all components
simultaneously and for some purposes it is unnecessary. Therefore, methods including those described
below can be used singly or in combination to provide a range of information from bulk composition to
detailed chemical ‘fingerprints’ of a crude oil spilled in the environment. A summary of analytical
methods used in spill response (a subset of all available methods) is presented below and in Table A1. A
recent review of additional methods used for analysis of crude oils (‘petroleomics’) is available (Rogers
and McKenna 2011).

The concept of Total Petroleum Hydrocarbons (TPH) is important in oil analysis. TPH is a single value
for each distinct crude oil, and represents the total mass of all hydrocarbons in that oil, including the
volatile and extractable (non-volatile) hydrocarbons that are recovered and detected using a specific
method. When used in the context of an environmental sample, TPH often refers to the total hydrocarbons
that can be extracted from the sample (EPH or extractable petroleum hydrocarbons). Many methods are
available, and each will yield a different TPH value for a sample depending on the properties that the
method analyzes, as described below. Therefore, the TPH value must be accompanied by a definition of
the method used to obtain the value, and only recognized, standard methods, such as those vetted and
archived by ASTM International (www.astm.org; formerly the American Society for Testing and
Materials), the US-EPA, and ISO standards, should be used.

A potential complication relevant to spilled oils recovered from the environment for analysis is that non-
petroleum hydrocarbons, such as plant oils and waxes, may be co-extracted from soils rich in vegetation
or plant debris, and subsequently be detected by methods outlined below, thus contributing to TPH values
without accurately reflecting the oil itself.

Infrared (IR) spectroscopy is a simple, non-destructive, rapid and inexpensive analytical technique
developed for analysis of oils and greases that provides bulk composition data and allows estimation of
physical properties of oils (Khanmohammadi et al. 2012). Typically, petroleum is extracted from an
environmental sample (usually water or soil) using a solvent (historically Freon 113, now replaced with a
non-fluorocarbon solvent like cyclohexane; ASTM [2011]), and a beam of infrared light is passed through
the extract. Absorption of particular IR wavelengths yields a spectrum based on the type and abundance
of chemical bonds in the extract and provides a measure of TPH. In theory, most hydrocarbons will be
detected quantitatively using this analysis, except for the very light and very heavy compounds because
the former may evaporate and the latter may not be soluble in cyclohexane. A drawback is that the
standard method as applied to crude oils does not identify specific chemicals in a mixture and the
presence of water (e.g., in oil samples recovered from the environment) interferes with accurate
measurements.

Fluorescence spectroscopy (FS) is simple, non-destructive and non-specific, like IR, but employs
ultraviolet (UV) wavelengths (Steffens et al. 2011). When some aromatic hydrocarbons are irradiated
with UV they re-emit that light at a different wavelength. This light can be captured and analyzed as a
proxy for TPH concentration. The simplicity of the measurement lends itself to remote sensing of crude
oil in aquatic systems. Several instruments in different configurations have been developed to detect and
estimate petroleum concentrations in real time. For heavy oils the fluorescence signal is broad and short-
lived, although it is more robust for lighter oils (Steffens et al. 2011). FS only gives an estimate of the
aromatics present in a sample and thus is inaccurate for estimating total hydrocarbons.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 380
Gravimetric analysis is a rapid, inexpensive laboratory technique for determining TPH, typically after
solvent extraction of the oil from an environmental sample. The mass of material that dissolves in the
solvent and then remains after the solvent is evaporated is weighed. Unless combined with another
analytical method, gravimetry cannot determine which specific chemicals within a class are present (or
have been lost to the environment), but it can determine bulk changes in the oil and therefore indicate
changes in gross properties. This method is unsuitable for light oils or dilbits that suffer large evaporative
losses, but is more appropriate for heavier crudes and may be particularly useful for analyzing oil samples
that present difficulties for other techniques, such as heavy oils and bitumens that have high resin and
asphaltene contents and resist detailed chemical characterization. However, the measurement will include
any non-petroleum materials in the sample that are extracted by the solvent used.

Liquid-Solid Column Chromatography and Fractionation. Oil that is recovered from the environment
using extraction solvents subsequently may be subjected to chromatographic fractionation. Typically, the
light fractions in the oil are depleted by artificial weathering in a laboratory fume hood (‘topping’ the oil)
to make subsequent steps easier and because these light molecules would be lost during the fractionation
procedure anyway. After determining the mass of oil that has been lost during topping (by mass
difference), the asphaltene fraction is removed by precipitation using either n-pentane or n-hexanes, dried
and weighed. The de-asphalted oil (also called the ‘maltenes’) is then passed through a glass column
containing adsorbent(s) like silica. Discrete fractions (e.g., SARA; Chapter 2) comprising chemicals in
different solubility classes are then eluted using solvents of different polarity (Figure A1). The masses of
the fractions are determined after evaporating the solvents so that gravimetric changes, such as those
caused by weathering (Chapter 2), can be discerned. Additionally, the saturate and aromatic fractions can
be subjected to detailed chemical analysis using, for example, GC-FID or GC-MS, described below, to
yield additional insight into the oil’s composition. Thus, chromatography and fractionation can provide
quantitative information, as well as composition, within technical limitations of detection.

Crude oil

Deplete light components


(VOLATILES) in a fume hood

“Topped” oil
Precipitate
ASPHALTENES
in n-pentane or
n-heptane
Maltenes
(de-asphalted oil)
Adsorption
chromatography

SATURATES AROMATICS RESINS

Figure A1 Liquid-solid chromatography and fractionation yielding SARA (saturates, aromatics, resins and
asphaltenes) fractions of crude oil (Chapter 2), plus volatile hydrocarbons determined from mass of initial oil
minus mass of topped oil.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 381
Gas Chromatography (GC) uses principles similar to those of the adsorption chromatography
fractionation method described above, but uses elevated temperature and gas flow across the adsorbent
phase rather than liquid solvents at ambient temperature to separate individual chemicals in the oil. Long,
small-bore capillary columns (e.g., fused silica columns of 30 m length, 0.25 mm diameter with an ultra-

are available to separate specific chemical classes, providing a characteristic ‘fingerprint’ (chromatogram)
of the oil where, ideally, each peak represents a single chemical (Figure A2). In practice, a finite range of
molecular sizes and composition having boiling points below 524 °C can be detected using GC, routinely
from C6 to C30. Only materials having no or low polarity are resolved, i.e., many saturates and aromatics,
but few resins (asphaltenes are removed before GC analysis to prevent fouling of the column and
instrument). Therefore, TPH values obtained by this method are limited to only GC-detectable and -
resolvable compounds. This is important when analyzing weathering losses from heavy oils in which the
large majority of material is not GC-detectable, because any changes observed by GC represent only
minor mass changes in the total oil composition.

GC methods are in common use because of the broad range of hydrocarbons that are detected selectively
and sensitively, and because identification of individual compounds can be achieved by coupling various
detectors to GC analysis, including flame ionization detectors (GC-FID) and mass spectrometers (GC-
MS), among others.

GC-FID uses a non-selective hydrogen flame to ionize and detect a broad range of compounds that are
separated by and elute from the GC column. It is a sensitive and reproducible method commonly used for
detecting ~C6 to C30 hydrocarbons and some resins, but is not suitable for large, polar molecules such as
asphaltenes. It is commonly applied to analysis of the most readily biodegradable fractions, the saturates
and aromatics. However, even within these fractions, the presence of myriad isomers can render
resolution of individual compounds difficult; in that case, the chromatogram is characterized by a ‘hump’
comprising unresolved peaks called the Unresolved Complex Mixture (UCM). The UCM is the dominant
feature in chromatograms of heavy oils. Examples of GC-FID chromatograms are shown in Figure A2.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 382
WEATHERED

PARTIALLY BIODEGRADED

Figure A2. Examples of GC-FID chromatograms showing the GC-detectable saturate fraction of an Alaska
North Slope crude oil. Top: weathered oil, showing evaporative loss of ‘light ends’ below C12; Bottom: weathered
and partially biodegraded oil, showing selective removal of n-alkanes below C24. Numbered peaks: n-C16 and n-
C24= n-hexadecane and n-tetracosane, respectively; Pr and Ph= pristane and phytane, respectively (biomarkers
that are slow to degrade); S= squalane, a reference hydrocarbon added to the oil as an internal standard. The
remaining taller peaks represent the n-alkane series, differing by a single –CH2 group, and the shorter peaks
represent multiple different isomers of iso- and cyclo-alkanes. The ‘hump’ toward the right-hand side represents the
unresolved complex mixture (UCM), comprising isomers that are not resolved by the GC conditions used.

GC-MS: The mass-selective detector yields highly detailed information about individual chemicals as
they elute from the GC column, including inferred compound identity. However, although it is very
sensitive, it may not be able to discern differences between multiple isomers that are common in crude
oils. GC-MS is not used to measure bulk properties, such as TPH, but is more of a ‘fine tool’ to discern
specific GC-detectable changes in oil composition.

A major disadvantage of all GC-based methods is that only certain classes of chemicals are resolved on a
given adsorbent column and the detectors are only suitable for certain classes of petroleum components.
For example, whereas saturated and aromatic hydrocarbons within a molecular weight range (e.g., ~C 6–
C30) are routinely detected using GC-FID and GC-MS, larger saturated and aromatic compounds and the
resins and asphaltene fractions are particularly difficult to resolve and/or are undetectable by either
method. GC-MS is particularly useful for routine analysis of changes in the specific chemical
composition of an oil as it weathers or biodegrades, but usually does not provide comprehensive
information, as it does not detect or identify large compounds.

An important consideration of GC analysis is that ‘disappearance’ of a peak from the oil chromatogram
(‘oil fingerprint’) does not necessarily mean that the corresponding chemical has been completely
removed from the environment. It is possible that the compound has been only slightly altered, e.g., by
photooxidation (see Chapter 2) or by partial biodegradation by microbes, with the majority of the
hydrocarbon skeleton remaining intact but being too polar to be detected or resolved by GC. Thus, GC

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 383
chromatograms must be analyzed carefully so as not to over-interpret the fingerprint and thereby
overestimate degradation.

High Temperature Simulated Distillation by Gas Chromatography (GC-HT SimDis or HTSD)


separates the components of petroleum in the order of their boiling points (Roussis and Fitzgerald 2000),
generating bulk compositional information about the oil. It is particularly useful for analysis of oils with
high contents of waxes, resins and asphaltenes that cannot be measured using conventional GC methods.
The curve generated by detecting boiling ranges as the analytical temperature increases to >700 °C can be
used to compare different oils or to refer to calibration curves to estimate the oil’s composition. The shape
of the boiling point distribution curve (i.e., cumulative % mass loss with increasing temperature) reflects
the proportions of low- and high-boiling point components in the oil.

An example is shown in Figure A3, comparing HTSD for a conventional oil (ASMB), an intermediate
fuel oil (IFO-180) and two dilbit samples (Cold Lake and AWB; Table 2.2). A simplistic interpretation of
the distillation curves is that a greater proportion of ASMB components have lower boiling points (are
lower molecular weight) than those of IFO-180; Cold Lake and AWB have similar compositions and
include light ends present in much lower proportions than the bulk of the oil and are similar in content to
those in ASMB.

Figure A3. High Temperature Simulated Distillation (HTSD) plot showing the simulated boiling point
distributions for Cold Lake (CLK) and Access Western Blend (AWB) dilbits, Alberta Sweet Mixed Blend (ASMB)
conventional crude and Intermediate Fuel Oil 180; see Table 2.2 for oil descriptions. Figure courtesy of Dr. H.
Dettman, National Resources Canada – CanmetENERGY, Devon, Alberta.

Non-routine analytical techniques

Two other techniques used for petroleum analysis are ElectroSpray Ionization Fourier transform ion
cyclotron resonance mass spectrometry (ESI FT-ICR-MS) and Nuclear Magnetic Resonance (NMR)
(Rogers and McKenna 2011). Both are highly sophisticated and require extremely expensive instruments,

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 384
as well as skilled operators, but they provide fundamental insights into petroleum composition that is not
afforded by more routine analytical techniques.

Other methods are available (e.g., high performance liquid chromatography [HPLC]; Table A2) but they
are typically used for more specialized analysis.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 385
Table A1. Summary of chemical analysis methods commonly used in oil spills

Name of method Description of method Advantage of the method Disadvantage of the method Reference
Infrared  Provides miscellaneous information about  Can analyze the hydrogen  Gives more qualitative than Brown et al. 1975;
spectroscopy (IR) complex mixtures (e.g., hydroxyl and carbonyl bonding in the crude oil quantitative information Butt et al. 1986;
groups) mixture  Cannot detect some compounds Fernández-Varela
 Measurement of a great  In some cases should be et al. 2006; Gautam
number of structural combined with other et al. 1998; Lynch
parameters spectroscopic techniques, such and Brown 1973
 Helps determine oil age as NMR spectroscopy
 Includes standards only for ‘oils
and greases’, not specifically for
petroleum
Gas  Separates and analyzes complex oil components  Quick analysis; small sampling  Limited use in heavy crude oil Butt et al. 1986; de
Chromatography that can be vaporized without decomposition amount analysis Andrade et al.
(GC)  Sample is carried through a column by the gas  Well analyze light oil  FID: ideal for limited 2010; Gaines et al.
phase; chemical and physical properties of  FID: very sensitive and compounds; compounds are 1999; Kenkel
components differentiate their rate passing selective destroyed 2002; Kim et al.
through the column, resulting in different  TCD: universal detector; low  TCD: not very sensitive; 2013; Payne et al.
retention times selectivity, low sensitivity compounds are destroyed 2008;
 Separated substances leaving the column are  FPD: identify heavily Simanzhenkov and
quantitatively detected by an electronic signal, biodegraded oils, sulphur Idem 2003; Yim et
which can be later detected by using detectors heterocycles and others al. 2011
with different sensitivity and selectivity, such as
flame photometric detector (FPD), flame
ionization detector (FID), and the thermal
conductivity detector (TCD).
Mass  Gives information on compound structures, such  Identify compound structure  Instrument can be expensive Eide and Zahlsen
spectrometry as molecular weight and formula  Results may require technically 2005; Li et al.
(MS)  Used for compound identification and new sophisticated interpretation 1997; Rostad 2006;
compound detection Sun et al. 2009;
 MS techniques include electron impact (EI), Tzing et al. 2003;
chemical ionization (CI), field ionization (FI), Wang et al. 2002
and fast atom bombardment (FAB)
 Combine with chromatographic methods: GC-
MS, HPLC-MS
High performance  Use high pressure to force chemical compounds  Can analyze compounds with  Limited use in crude oil analysis Fish et al. 1984;
liquid to pass through a metal column low volatility that are due to the limited compound Kenkel 2002;
chromatography  Detector used: UV absorption (UV), refractive inappropriate for GC analysis species determined Pasadakis et al.
(HPLC) index (RI) or fluorescence (F)  Can analyze some compounds  Unable to detect some 2001; Speight and
that GC is unable to analyze compounds, depending on the Speight 2002
(e.g., heavy crude oil) with detector used
precision in a short time  UV and F only detect some
species; F lacks sensitivity and
is influenced by temperature

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 386
APPENDIX B

Table B1 Major properties of oil, analytical methods and relevance to oil behaviour
Relevance with application and
Parameter Description of parameter Analysis method References
cleanup measurements
Viscosity  Measure of flow resistance  Viscometer: determines the flow  Viscous oil spreads on water surface Aske et al. 2002;
 Liquids with lower viscosity flow faster rate of the oil under standard slowly and penetrates Beal 1946; Beggs
 Dependent on oil composition; hence, the conditions; higher flow rate soils/sediments slowly and Robinson
viscosities of crude oils vary. indicates lower viscosity  Oils with high viscosity affect the 1975; Lambert
 Oils with greater saturates and aromatics  Rheometer: measures liquid flow in efficacy of pumps and skimmers 2003; Morrison
fractions and lower asphaltenes and resins response to applied force and Murphy 2006;
content have lower viscosity  ASTM method D445 ASTM 2015
 Weathering of oil increases its viscosity
 Generally exponentially dependent on
temperature
Density  Mass of a unit volume of oil, expressed as g/mL  Density meter (densitometer)  Light grade and heavy oils Aske et al. 2002;
or kg/m3  ASTM method D 5002  Submerged oil detection (e.g., ASTM 2005;
 Density of oil typically 0.7-0.99 g/mL, (versus acoustic remote sensing); surface Fingas et al. 2006;
seawater density; 1.03 g/mL) oil detection (e.g., optical remote
 Most oils float on seawater, except for a few sensing)
bitumens  Heavier oils are more persistent and
 Heavily weathered oils tend to have greater may pose greater challenge for
density remediation
 Oil density is more temperature-dependent than
is water
Specific  Ratio of oil density to receiving water density  ASTM method D287 -12b: Standard  Gives approximate hydrocarbon ASTM 2012a;
gravity  Varies with temperature test method for API gravity of crude composition and heat of combustion Fingas 2010
(relative  Quality indicator of oil by American Petroleum petroleum and petroleum products  Affects oil storage, handing and
density) Institute (API), or API gravity (hydrometer method) combustion
Surface  Net stresses at the boundaries between different  Tensiometer  Affects oil spreading Aske et al. 2002;
tension substances  du Noüy ring method  Relates to the final size of the oil ASTM 2012b;
 Expressed as the increased energy per unit area,  ASTM method D971 slick du Noüy 1925;
or as force per unit length (mN/m)  Lower surface tension leads to
thinner and larger extent of slick
Flash point  The temperature at which the vapor over the  Flash point analyzer  Safety indicator for spill cleanup ASTM 2007;
liquid can be ignited  ASTM method D1310 (for low process (light fuels such as gasoline ASTM 2013
 Flammable liquid viscosity oil), and ASTM method can ignite under most ambient
D93 (for heavier oil) conditions)
Pour point  The temperature at which oil ceases to flow in a  ASTM method D97  Limited use as an indicator of oil ASTM 2012b
five second-period status
 Ranges from -60 °C to 30 °C for crude oils  Important for oil recovery and
 Lighter, less viscous oils have lower pour transport
points

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 387
APPENDIX C
Table C1: Additional details of composition and properties of relevant crude oils and refined products. Data
excerpted from sources listed in references.
Oil type API specific % % % % % adhesion %
(source province gravity gravity sulphur saturates resins asphalt- waxes (g/m2)* dispersion
or country) (°) at 15°C enes (agent)
Condensates and refined products used as diluents
Sable Island 40 0.82 0.03 88 0 1
condensate (NS)
Fort 68 0.67 0.14 73
Saskatchewan
condensate blend
(AB)
Cold Lake Diluent 69 0.70 0.25
(AB)
Naphtha (AB) 57 0.75 0.07 85
Southern Lights 80 0.68 0.03
Condensate
Diluent (US)
Suncor Synthetic 33 0.86 0.2 82 1 0 40
Crude Crude Oil/ (Corexit
SCO (AB) 9500)
Light crude oils
Alberta Sweet 36 0.84 0.6 65 5 33 4 13-61 40
Mixed Blend (Corexit
(ASMB) 9500)
Reference oil #4
(AB)
Macondo 35 0.84 0.3 0
(MC252) (USA)
Norman Wells 38 0.83 0.4 86 2 1 2 35
(NWT) (Corexit
9500)
Transmountain 34 0.86 0.8 81 2 4 7
Blend (BC)
Statfjord 38 0.84 0.3 68 6 2 8 14-62 40
(Norway) (Corexit
9500)
Arabian Light 32 0.86 1.9 76 6 4 3 17-35 19
(Saudi Arabia) (Corexit
9500)
West Texas 36-41 0.85 0..9 79 6 1 3 12-32 28
Intermediate (Corexit
(USA) 9500)

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 388
Medium crude oils
Alaska North 30 0.89 1.1 52 9 5 28-33 46%
Slope (ANS) (Corexit
middle pipeline 9500)
(USA)
Atkinson Point 24 0.91 0.9 48 14 3 1
(NWT)
Prudhoe Bay-A 29 0.88 1 53 10 4 4 24-29 10
(USA) (Corexit
9500)
Gullfaks, Troll 29 0.88 0.3 60 5 1 4 23-35 25
(Norway) (Corexit
9500)
Shale oils
Bakken (USA) 42 variable;
usually
<0.3
Waxy crude oils
Hibernia (NL) 35 0.85 73 7 2 5 12-160 11-21
(Corexit
9500)
Terra Nova (NL) 36 0.85 0.4 62 6 2 9 10-80 14
Sour crude oils
Midale (SK) 31 1.5 51 9 5 5 25
(Corexit
9500)
BC Light (BC) 41 0.82 0.6
Western Canadian 22 0.92 3
Blend (WCB)
(AB)
West Texas Sour 30 0.78 1.5 51 9 5 5 21-25 25
(USA) (Corexit
9500)
Heavy oils and diluted bitumen
Bunker C fuel oil 11 0.99 0.5 25 17 11 2 85-421 6-14
(Fuel Oil No. 6) (Corexit
(international) 9500)
Cold Lake blend 23 0.92 4.7 13-977
(dilbit) (AB)
Albian Heavy 20 0.94 2.3 42 26 6 15
Synthetic (AB) (Corexit
9527)
Access Western 23 0.92 4
Blend (AWB)
(AB)

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 389
Wabasca Heavy 16 0.96 5 41 28 9 <10
(AB) (Corexit
9527)
Western Canadian 19-22 0.93 3
Select (WCS)
(AB)
Synbit Blend (AB) 21 0.93 2.9
Athabasca 8 1.01 5 5
bitumen
(undiluted) (AB)
Cold Lake 9.8 1.0002 6.9 13 2 0
bitumen (Corexit
(undiluted) (AB) 9527)
Orimulsion-100 8 1.01 2.3 17 16 20
ultraheavy oil and
water (Venezuela)
* range of adhesion values represents different weathered states

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 390
APPENDIX D
Table D.1. Summary of effects of petroleum products (crude oil, refined/residual oil, synthetic oil, oil sands bitumen) on fish.
Oil type Weathered? Species Life stage Response Type Time Method Fresh/ Lab/ Nominal Measured Notes Reference
Seawater Field
Crude oil Unknown Nile tilapia (Oreo- Adult Blood (RBCC Acute 72 h WAF Fresh Lab Hemoglobin Nuclear swelling of Al-Ayed
(Source chromis niloticus) and hematocrit dropped at 2000 erythrocytes did not 2001
unknown) value) ppm after 48 h recover after remov-
al of crude oil expo-
sure, unlike hemo-
globin, hematocrit

Hungary No Common carp Adult Cytochrome Acute 3d and IP injection Fresh Lab 2mL/kg for 3d: Deer et al.
crude (from (Cyprinus carpio), P-450 enzyme and 8d of 2 mL/kg EROD elevated 6- 2010
Szeged- Silver carp (Hy- (EROD, chronic fold over control in
Algyo, phothalmichtys ECOD) carp and 4-fold over
Hungary) molitrix), European induction; control in eel; LP
eel (Anquilla an- antioxidant reduced in carp.
quilla) parameters
(only lipid 2 mL/kg for 8d
peroxidation (Carp): Drop in
(LP) affected) EROD, but still
higher than controls.

4mL/kg for 3d or
8d: Same increase in
EROD as 2mL/kg
ANS crude Yes; 70 °C Pacific herring Adult TPAH in Chronic 16-18 d WAF from Seawater Lab Significant correla- Carls et al.
until 20% loss (Clupea pallasi) tissue; preva- oiled gravel tions between PAH 1998
lence of viral columns exposure and VHS
haemorhagic prevalence; cumula-
septicemia tive mortality
(VHS) and
mortality
ANS crude Yes; 70 °C Pacific herring Adult BSD; mor- Chronic 16 d WAF from Seawater Lab No effects on ova 82-94% of PAH Carls et al.
until 20% loss (Clupea pallasi) tality oiled gravel from 16 d expo- accumulated to ova 2000
columns sure of adult fe- in exposed adults
males to 58 ug were naphthalenes,
PAH/L indicating that
HMW and alkyl
PAH were metabo-
lised and/or exposure
occurred before cell
differentiation; indi-
cates that embryos
are more sensitive
than gametes

ANS crude Yes; heated to Pacific herring Adult Cytochrome Chronic 16 d WAF from Seawater Lab Pre-spawn fish more Thomas et al.
70 °C until (Clupea pallasi) P450 (EROD) oiled gravel vulnerable (less 1997
20% loss induction columns EROD activity)

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 391
Oil type Weathered? Species Life Response Type Time Method Fresh/ Lab/ Nominal Measured Notes Reference
stage Seawater Field
Exxon Yes, naturally Dolly varden Adult Reproductive Chronic 1- 3 y Stranded Seawater Field Elevated biliary Sol et al.
Valdez oil in field (Salvelinus malma parameters oil in field FACs in 1989, de- 2000
malma), (plasma estra- creased by 1991;
yellowfin sole diol, GSI, reproductive parame-
(Limanda aspera), GTH-1), ters were consistently
pollock (Pol- biliary fluo- low at high FAC
lachius sp.) rescent aro-
matic com-
pounds
(FACs)
Exxon Yes, naturally Pacific herring Adult Histopatholog y Chronic ?? Unknown Seawater Field Hepatic necrosis Hepatic necrosis in Marty et al.
Valdez oil in field (Clupea pallasi) (hepatic necro- in oiled sites in herring from oiled 1999
sis, spleen, 1989 only sites in 1989 may
kidney), have been a result of
tissue con- viral hemorrhagic
centration s of septicemia (VHS);
PAH see Carls et al. 1998

Exxon Yes (in field) Salmonids Adult Population Chronic Unknown Unknown Seawater Field Mean fish densities Barber et al.
Valdez oil density were low in 1990; 1995
same as control (un-
oiled sites) by 1991,
suggests recovery of
PWS
United Yes (in field) Rockfish (Sebas- Adult Hepatic Chronic 5d Field expo- Seawater Field Muscle alkyl Biliary PAH metabo- Jung et al.
Arab tes schlegeli), CYP1A induc- sure to oil PAH = 284 ng/g lite concentrations 2011
Emirates marbled flounder tion; biliary from Hebei dw (5 d post- decreased to back-
Upper (Pseudopleuronec PAH metabo- Spirit oil spill) ground by 11 m post-
Zakum, tes yokohamae) lite concentra- spill (2007) spill.
Kuwait tions
Export
Crude, and
Iranian
Heavy
Crude
Bass Strait No Crimson-spotted Adult Reproductive Acute 3 d oil; WAF; Fresh Lab BSD EC50 = Pollino and
crude rainbowfish performance 14 d CEWAF 0.43 mg HC/L; Holdway
(Melanotaenia (egg production) clean genotoxicity 2002
fluviatillis) water LOEC = 0.24
mg/L
Bunker C No Japanese flounder Adult Immune system Acute 3d 3.8 g/L oil Seawater Lab Down-regulation of Nakayama et
(Paralichthys gene expression in water immune response al. 2008
olivaceus) genes, may be used
as biomarkers?

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 392
Oil type Weathered? Species Life Response Type Time Method Fresh/ Lab/ Nominal Measured Notes Reference
stage Seawater Field
Campos No Tetra (Astyanax Adult Histopathology Acute 12, 24, WAF Fresh Lab Liver: necrotic Also measured ace- Akaishi et al.
Bay crude, sp.) (gill, liver) 96 h areas and loss of tylcholinesterase 2004
Brazil hepatocyte cell activity (neurotoxici-
limits at 15% ty): decreased activi-
WAF; ty at 15% WAF for
96 h and 33% for 24
Gill: aneurysms h
at 33% WAF
ANS crude Yes; heated Zebrafish (Danio Adult Heart shape; Chronic 96 h Static WAF Fresh Lab Subtle changes See Incardona (2005) Hicken et al.
to 70 °C rerio) reduced and WAF in heart shape for embryo expo- 2011
until 20% swimming from strand- and reduced sures
loss performance ed oil col- swimming abil-
(reduced cardi- umns ity, indicative of
ac output) reduced cardiac
output

Prudhoe Yes (in field) Pacific herring Adult Reproductive Chronic ?? Field collect- Seawater Field Unknown Artificial fertiliza- Kocan et al.
Bay crude (Clupea pallasi) success and ed adults; tion of embryos from 1996b
histopathology artificial "exposed" adults in
fertilization PWS; oiled - re-
and rearing duced hatching
of embryos success, increased
deformities

Prestigelike No Rainbow trout gill Cells Cytotoxicity; Acute 24 h 5 - 45 g oil Fresh Lab LOEC = >70 µg No cytotoxicity in Navas et al.
heavy fuel cell line EROD per 100 ml TPAH/L RTG-2 cells, but 2006
oil (IFO growth me- dose-dependent
380) dium (pro- increases in EROD
duced were observed
WAF)
Alberta oil No Fathead minnow Embryo Mortality; Acute 96 h Direct expo- Fresh Lab Fathead minnow: Fathead min- EC50 (CYP1A in Colavecchia
sands (Pimephales cytochrome sure to con- LC50 = 0.06 now: LC50 = 1 kidney) was more et al. 2007
bitumen promelas), P450 (CYP1A) taminated (WWP) - 0.91 (S- - 29 ng sensitive than eye
and tailings White sucker induction; eye sediments Nat) g sediments/ TPAH/L pathology
pond sedi- (Catastomus com- pathology (renewed L (calculated)
ments mersoni) daily)
White sucker: White sucker:
LC50 = 0.38 (E- LC50 = 14 - 36
Nat) - 1.15 (S- ng TPAH/L
Nat) g/L (calculated)

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 393
Oil type Weathered? Species Life stage Response Type Time Method Fresh/ Lab/ Nominal Measured Notes Reference
Seawater Field
Arabian Yes; sparged Sheepshead min- Embryo Mortality Acute 96 h WAF; Seawater Lab Sheepshead min- Fuller et al.
medium with air at now (Cyprino- CEWAF now: Continuous 2004
crude room T, until don variegatus), from stirred NOEC (mortality)
25% loss Inland silverside aspirator = 10 mg oil/L;
(Menidia berylli- flasks - both declining NOEC =
na) 36 mg/L

Silverside min-
now: continuous
NOEC (mortali-
ty) = <1.2 mg/L;
declining NOEC =
26 mg/L
Bass Strait No Australian bass Embryo Mortality Acute 96 h WAF; Seawater Lab CEWAF: Gulec and
crude (Macquaria no- CEWAF LOEC (mortali- Holdway
vemaculeata) (24 h stir, 1 ty) = 3.13 mg/L 2000
h settle - no
renewal)
MESA Yes; sparged Mummichog Embryo Survival; Acute 96 h WAF; Seawater Lab (Estimated) LC50: (Estimated) Couillard et al.
with air for (Fundulus hetero- body length; CEWAF WAF >1 g oil/L; EC50 (body 2005
24-48 h at clitus) cytochrome CEWAF = 0.35 length and
room T, until P450 enzyme g/L EROD) = 150
20% loss activity ng PAH/mL
(EROD) Body length
EC50: WAF = 0.1
(Estimated)
g/L; CEWAF =
LC50 = 200-
0.025 g/L
400 ng
PAH/mL
EROD EC50:
WAF = 0.1 g/L;
CEWAF = 0.05
g/L
No 2 fuel oil No Sheepshead min- Embryo Mortality Acute 96 h WAF from Seawater Lab LC50 (WAF) = 80 Adams et al.
+ Omni- now (Cyprino- 20-30 s (20 ppt) mg/L; (CEWAF) 1999
Clean dis- don variegatus) electric = 80 mg/L (esti-
persant stirring mated)
South Lou- 17.5 h mixing Japanese medaka Embryo Mortality Acute 4d Microcosms Fresh Lab Diesel was more Bhattacharyy
isiana with de- (Orzyias latipes) (water + toxic than South a et al. 2003
crude, die- ionized soil); 1, 7, Louisiana crude; no
sel fuel water, 10 or 31, 186 d toxicity data provid-
60 min after prepa- ed aside from corre-
settling, ration lations among varia-
weighed to bles
determine %
loss

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 394
Oil type Weathered? Species Life Response Type Time Method Fresh/ Lab/ Nominal Measured Notes Reference
stage Seawater Field
Alberta oil No Fathead minnow Embryo Mortality; Chronic 12 d (5 d Indirect Fresh Lab LOEC: 0.05 to 0.8 65 (WWP) - Sediments were Colavecchia
sands (Pimephales pro- BSD; growth to exposure to g sediment/L 176 (A-Nat) µg dominated by HMW et al. 2004
bitumen melas) hatch, 7 contaminated TPAH/L (Cal- 3- to 6-ringed alkyl
and tailings post- sediments culated from PAH; very little
pond sedi- hatch) (renewed TPAH of stock alkyl naph - most
ments daily) sediments) sensitive response
was mortality
Alberta oil No White sucker Embryo Mortality; Chronic 35 d Indirect Fresh Lab LOEC: 0.1 36 (S-Nat) - 100 Most sensitive Colavecchia
sands (Catastomus BSD; growth (14 d to exposure to (weight) to 0.4 (E-Nat) µg response was growth et al. 2006
bitumen commersoni) hatch, contaminated (edema, hemor- TPAH/L (Cal- and BSD and EROD
and tailings 21 post- sediments rhage) g sedi- culated from activity; premature
pond sedi- hatch) (renewed ment/L stock sediments hatch occurred at all
ments daily) - LOEC across exposures; pericar-
LOEC (EROD all endpoints dial edema respon-
activity) = 0.002 sible for observed
g/L (WWP) mortality
ANS crude "Yes"; biode- Inland silversides Embryo Teratogenicity Chronic 7-10 d WAF Seawater Lab LOEC (develop- Middaugh et
graded 14d (Menidia berylli- recovered mental) = 0.75 al. 2002
using Phe#6, na) after degra- mg/L WAF
Hexaco#2, dation
EI2V

ANS crude Yes; and Inland silversides Embryo Teratogenicity Chronic 7-10 d WAF Seawater Lab LOEC (develop- Middaugh et
biodegraded (Menidia berylli- recovered mental) = 1% al. 1996
na) after degra- WAF
dation
ANS crude No Pacific herring Embryo Hatching Chronic 4d WAF; Seawater Lab 96-h NOEC Oil + dispersant Barron et al.
(Clupea pallasi) success and CEWAF (yolk sac edema (high energy WAF) 2003
timing; vitali- embryos) = 17 exposures were
ty; abnormali- µg TPAH/L followed by expo-
ties 96-h NOEC (oil sure to UVA. Expo-
and oil+disp) = sure to UVA caused
9.2 µg TPAH/L 50-fold increase in
96-h NOEC toxicity
(oil+disp+UV)
= 0.2 µg
TPAH/L
ANS crude Yes; 70 °C Pacific herring Embryo LOEC: Chronic 16 d WAF from Seawater Lab LWO: LOEC = Based on initial or Carls et al.
until 20% loss (Clupea pallasi) mortality; oiled gravel 9.1 µg peak concentrations, 1999
BSD columns TPAH/L; which declined with
EC50 = 18.4 µg time. MWO (more
TPAH/L weathered oil) -
(swimming same oiled gravel as
ability) originally used in
LWO (less weath-
MWO: LOEC = ered oil)
0.41 µg
TPAH/L; EC50
= 0.33 (cranio-
facial)

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 395
Oil type Weathered? Species Life stage Response Type Time Method Fresh/ Lab/ Nominal Measured Notes Reference
Seawater Field
ANS crude No Pink salmon Embryo Growth Chronic 10 d WAF (glass Seawater Lab 0.18-0.35 mg Birtwell et al.
(Oncorhynchus extraction HC/L 1999
gorbuscha) column;
Moles et al.
1985)
ANS crude No Pink salmon Embryo Histology Chronic 10 d WAF (glass Seawater Lab LOEC (histo- Morphological and Brand et al.
(Oncorhynchus (hepatic, head extraction pathology) = 25- stress-induced le- 2001
gorbuscha) kidney, gill column; 54 µg HC/L; 96 sions in their hepat-
tissues); OR Moles et al. h LC50 = 1-3 ic, head kidney and
released and 1985) mg/L gill tissue; most
recaptured 1 effects are consistent
or 2 years over 3 years
later
ANS crude Yes; 70 °C Pink salmon Embryo Mortality; Chronic 83 d WAF from Fresh Lab 1500-2250 mg 7.8 - 16.4 µg 50% mortality Brannon et al.
until 20% loss; (Oncorhynchus BSD oiled gravel oil/kg gravel TPAH/L; 8.3 - achieved using arti- 2006a
and natural gorbuscha) columns 12.1 mg ficially weathered oil
weathering TPAH/kg grav- (not naturally weath-
el ered)
ANS crude Yes; 70 °C Pink salmon Embryo Mortality; Chronic 6 mo WAF from Seawater Lab LOEC (mortali- Carls et al.
until 20% loss (Oncorhynchus BSD; CYP1A oiled gravel ty) <16.5 µg 2005
gorbuscha) induction; and columns TPAH/L;
effects on
growth and
LOEC (CYP1A
survival 6 mo
induction;
later
reduced length)
< 3.7 µg
TPAH/L;

LOEC (reduced
mass) < 0.94 µg
TPAH/L
ANS crude Yes; heated to Pink salmon Embryo Mortality; Chronic Hatch WAF from Seawater Lab LOEC (mortality) LOEC (mor- Direct and indirect Heintz et al.
70 °C until (Oncorhynchus visible le- to stranded oil = 281 mg oil/kg tality) = 18 µg exposed embryos 1999
20% loss gorbuscha) sions; size at emerg- columns, gravel (dose) - TPAH/L showed similar ef-
emergence ence either fresh mortality of 35% fects
(up to 7 or 1 y W 3.8 mg TPAH/kg
mo) on gravel
ANS crude Yes; heated to Pink salmon Embryo Mortality and Chronic Up to 8 WAF from Seawater Lab LOEC (mortali- See Heintz et al. Heintz et al.
70 °C until (Oncorhynchus growth 2 y mo oiled gravel ty) = 5.4 µg (1999) for effects of 2000
20% loss gorbuscha) after embryo columns TPAH/L (15% oil exposure on
exposure mortality in embryos. This paper
returning fish) describes delayed
effects (on returning
adults 2 y later)
LOEC (growth)
= 18 µg
TPAH/L (may
account for
mortality)

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 396
Oil type Weathered? Species Life Response Type Time Method Fresh/ Lab/ Nominal Measured Notes Reference
stage Seawater Field
ANS crude No Pink salmon Embryo Mortality; Chronic 6d Contaminated Seawater Lab LC50 = 29 mg Carls et al.
(Oncorhynchus growth; con- food HC/L; LOEC = 1996
gorbuscha) dition; feeding 2.2 mg/g

ANS crude Yes; naturally Pacific herring Embryo Morphological Chronic ?? Oiled beach- Seawater Field No oil-related histo- Hose et al.
(from in field (1 - 3 (Clupea pallasi) deformities; es in Prince pathological lesions 1996
EVOS y) cytogenetic William observed. Also,
oiled areas) abnormalities Sound - oil-related develop-
(anaphase 1989, 1990, mental and genetic
telophase 1991 effects were unde-
aberrations); tectable in 1990
histopathology and 1991
lesions

ANS crude Yes; heated Mummichog Embryo Mortality; Chronic 11 d Direct expo- Seawater Lab LOEC (body LOEC (calcu- Couillard
to 70 °C (Fundulus hetero- BSD sure to con- length) = 12.7 µg lated) = 270 ng 2002
until 20% clitus) taminated oil/g sand TPAH/g sand
loss sediments
(renewed
daily)

Exxon Yes, naturally Pacific herring Embryo Embryo den- Chronic ?? Unknown Seawater Field Oil effects on Growth rates of McGurk and
Valdez oil in field (Clupea pallasi) sity; mortali- growth rates larval herring in Brown 1996
ty; length; PWS were 0.10-0.17
growth mm/d, half the rate
of wild populations
(0.21-0.41 mm/d);
Evidence of oil
damage? See Nor-
cross et al. (1996)
Iranian No Large yellow Embryo Mortality; Chronic ?? WAF Seawater Lab TJ014: EC50 = (Calculated) Shen et al.
crude croaker (Lar- hatch rate; (stirred 0.5 h, 112.4 mg/L; TJ014: EC50 = 2010
(TJ014), michthys crocea) BSD emulsified LC50 = 8.5 mg/L 0.66 mg TPH/L;
No. -20 8 h) LC50 = 3.01
diesel oil mg/L;
-20: EC50 =510.8
-20: EC50 =
mg/L; LC50 =
0.05 mg/L;
156.2 mg/L
LC50 = 0.92
mg/L

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 397
Oil type Weathered? Species Life Response Type Time Method Fresh/ Lab/ Nominal Measured Notes Reference
stage Seawater Field
MESA No Atlantic herring Embryo Mortality; Chronic 1-144 h CEWAF Seawater Lab 1 h EC50 (fertili- 1 h EC50 = 79 Also reported McIntosh et
(Clupea ha- BSD; fertili- zation inhibition) mg TPH/L ET50s using expo- al. 2010
rengus) zation success = 0.19 %v/v sure times 1 - 144 h.
3 h LC50 = 146
3 h LC50 (free mg TPH/L
swimming em-
bryos) = 0.35%v/v 10 h LC50 = 75
mg TPH/L
10 h LC50 (free
swimming em- 24 h LC50 = 32
bryos) = 0.18%v/v mg TPH/L

24 h EC50 (0dpf
embryos - BSD) =
0.08 %v/v
MESA Yes; heated Mummichog Embryo Mortality; Chronic 11 d Direct expo- Seawater Lab LOEC (body LOEC (calcu- Couillard
to 70 °C (Fundulus hetero- BSD sure to con- length) = 4.5 µg lated) = 230 ng 2002
until 20% clitus) taminated oil/g sand TPAH/g sand
loss sediments
(renewed
daily)

No 2 fuel No Sheepshead min- Embryo Mortality; Chronic 7d WAF from Seawater Lab LC50 (WAF) = Adams et al.
oil + Omni- now (Cyprino- fish biomass 20-30 s (20 ppt) 130 mg/L; 1999
Clean dis- don variegatus) electric (CEWAF) = 70
persant stirring mg/L (estimated)
Prudhoe Yes (in field) Pacific herring Embryo Mortality, Chronic Field: FIELD: Seawater Lab LOEC (deform- Embryos exposed in Kocan et al.
Bay crude (Clupea pallasi) hatching suc- 8-12 d Caged in situ and ities) = 0.24 mg field to oiled sites 1996a
cess, growth Lab: at PWS oiled field TPH/L were consistently of
18 d and unoiled lower weight; con-
sites LOEC (hatch) = founding variable
0.24 mg/L effects on mortality
LAB: 36 h and hatching success
exposures to LOEC (geno-
oil-water toxicity - #
dispersions mitotic cells per
(OWD) fin) = 0.24
mg/L
Prudhoe Field - yes, Pacific herring Embryo Histo- Chronic 14-20 d Oil-water Seawater Lab LOEC (ascites) Larvae from oiled Marty et al.
Bay crude lab - no (Clupea pallasi) pathology (at dispersions and = 0.48 mg/L sites were shorter, 1997a
(evidently) hatch - (OWDs), or field had ingested less
differed collected food, had slower
by from "oiled" growth, and had high
dose) and "un- prevalence of cyto-
oiled" sites genetic damage,
ascites.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 398
Oil type Weathered? Species Life Response Type Time Method Fresh/ Lab/ Nominal Measured Notes Reference
stage Seawater Field
Prudhoe Yes; heated Pink salmon Embryo Sampled 4 Chronic Up to 18 WAF from Fresh Lab LOEC (develop- LOEC (devel- Marty et al.
Bay crude to 70 °C (Oncorhynchus weeks pre-, d oiled gravel mental) = 55.2 mg opmental) = 1997b
until 20% gorbuscha) at, and 13 d columns oil/kg gravel 4.42 µg
loss post- TPAH/L (peak)
emergence LOEC (CYP1A
induction; histo-
pathology - epi-
dermal thick-
ness) = 622 mg/kg

Ultralow No Rainbow trout Embryo BSD; mor- Chronic 24 d WAF, Fresh Lab CEWAF BSD WAF BSD EC50 Schein et al.
sulfur (Oncorhynchus tality CEWAF EC50: 0.04 % v/v and mortality LC50 2009
Diesel No. mykiss) > 1 % v/v
2
CEWAF mortality
LC50: 0.05 % v/v
ANS crude Yes; heated Pacific herring Embryo PAH uptake; Chronic 6-7 d WAF from Seawater Lab Significant brady- Incardona et
to 70 °C (Clupea pallasi) immuno- oiled gravel cardia at 12.5% al. 2009
until 20% fluorescence; columns oiled gravel dose
loss heart rate
ANS crude No Zebrafish (Danio Embryo Pericardial Chronic 2d Mechanically Fresh Lab (Estimated) Overlapping toxicity Carls et al.
rerio) edema; ab- dispersed oil EC50 (ab- curves for embryos 2008
normal heart (WAF); normal heart exposed to mechan-
looping; intra- direct and looping) = 25 ically dispersed oil
cranial hemor- indirect µg TPAH/L; and for embryos
rhaging contact EC50 (peri- embedded in aga-
cardial edema) rose and only ex-
= 45 µg posed to dissolved
TPAH/L fraction
Norman No Rainbow trout Embryo Mortality; Chronic 55 d WAF; Fresh Lab 30 uL oil/L: LT50 After 55 d exposure, Lockhart et
Wells crude (Oncorhynchus water content; CEWAF (estimated): fry from oil, disp, al. 1996
mykiss) growth oil alone = 55 d and oil+disp mix-
oil+disp = 20-45 d tures had greater
water content than
controls (edema?)
Corexit 7664 was
more toxic than
Corexit 9600
Lithuanian No Rainbow trout Embryo, Gill ventila- Acute 25 d Injection of Fresh Lab Reduced GVF; heart Vosyliene et
crude (Oncorhynchus adult tion frequency and (chronic) 1610 mg rate; and body mass al. 2005
mykiss) (GVF); body chronic and 4 d oil/L to tank in oil-exposed fish
mass; heart (acute)
rate

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 399
Oil type Weathered? Species Life Response Type Time Method Fresh/ Lab/ Nominal Measured Notes Reference
stage Seawater Field
Lithuanian No Rainbow trout Embryo, Mortality; Acute Embryo WAF Fresh Lab Embryo LOEC Kazlauskiene
crude (Oncorhynchus juvenile physiological and (7-8d); (stirred and (mortality) = 3.46 et al. 2008
mykiss) (heart rate, chronic larvae aerated for g oil/
gill ventila- (42-44d); 24 h)
tion) juvenile Larvae LOEC
(4d and (mortality) = 0.43
14d) g oil/L

Embryo and lar-


vae LOEC (de-
velopmental and
heart rate) =
0.87 g/L
Heavy fuel No Rainbow trout Embryo, Mortality Acute 96h WAF Fresh Lab Embryo LC50 = Kazlauskiene
oil (Lithu- (Oncorhynchus juvenile, and LC50; (stirred and 1.55 g HFO/L et al. 2004
anian ther- mykiss) adult chronic 60d LC50 aerated for
mal plant) (embryos, 24 h) Larvae LC50 =
larvae); 0.33 g HFO/L
28d LC50
(adult)
Adult LC50 =
2.97 g HFO/L

ANS crude Yes; 70 °C Pink salmon Gamete Mortality; Chronic 8h WAF Fresh Lab > 2250 mg oil/kg > 16.4 µg Observed mortali- Brannon et al.
until 20% (Oncorhynchus BSD (effluent) - gravel TPAH/L ty in gametes not 2006a
loss; and gorbuscha) exposure of statistically great-
natural weath- eggs during er than controls
ering sperm
activation,
fertilization,
and water
hardening
ANS crude No Gulf killifish Juvenile Mortality Acute 96 h Flow- Seawater Lab At least 83% sur- One time addition of Liu et al.
(Fundulus gran- through vival @ 30 mg oil+disp. in a contin- 2006
dis) canal with oil/L dose (HC uous flow system.
caged organ- concentration at 0 HC concentrations
isms - simu- = 14-24 mg/L declined to back-
lated oil water) ground after 3 h
spill
ANS crude No Pink salmon Juvenile Mortality Acute 96 h WAF (glass Seawater Lab 96-h LC50 = Birtwell et al.
(Oncorhynchus extraction 2.2 mg/L 1999
gorbuscha) column; (1990); 2.8
Moles et al. (1991); 1.0
1985) (1992)
Arabian Yes; evapo- Golden grey mul- Juvenile Biliary PAH Acute 48 h CEWAF Seawater Lab PAH metabolites and Milinkovitch
crude oil rated for 24 let (Liza aurata) metabolites; total glutathione et al. 2011b
(54% h using natural glutathione content of liver were
saturates, UV live content; found with CEWAF
36% EROD exposures (44 ug/L
aromatics) TPAH)

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 400
Oil type Weathered? Species Life Response Type Time Method Fresh/ Lab/ Nominal Measured Notes Reference
stage Seawater Field
Australian Yes, artificial Clownfish (Am- Juvenile Mortality Acute 96 h WAF (28 Seawater Lab LC50 (Clown- Overall, acute toxicity Neff et al.
crude oils: weathering by phiprion g/L stock - 0, fish): Wonnich F = is attributed to MAHs; 2000
condensate, distillation clarkii), silverside 8, 16, 64, 35% 100%, light toxicity of crude oils
light, (topping) at minnows 100%) crudes = >100% also attributed to
medium; 150-250 °C to (Menidia berylli- PAHs (about 58%),
Australian simulate 1-15 na) LC50 (Minnow): more so with weather-
diesel from d at sea Wonnich F = 32% ing - where MAH are
Apache 79% W, overall lost and PAH persist
more sensitive
than clownfish
Bombay high Yes; sparged Tilapia (Tilapia Juvenile PAH uptake Acute 96 h WAF Fresh and Lab Increased salinity Shukla et al.
oil field with air for mossambica) (Singer) seawater reduced PAH uptake 2007
(India) crude 130 h at room (15, 30 to liver, gills, muscle
T, until 14% ppt) and gonad; due to
loss reduced solubility or
uptake rates? Only
measured 5 parent
PAH in water, tissue

Brut No Thinlip grey mul- Juvenile Mortality Acute 24 h WSF; Seawater Lab CEWAF was more Milinkovitch
Arabian let (Liza ramada) WAF; toxic than MDO and et al. 2011a
Light oil CEWAF WSF (mortality),
(0-40%) correlating with high
PAH, TPH concentra-
tion
Bunker C No Red sea bream Juvenile Mortality Acute 48 h Static WAF Seawater Lab 0.325 mg total Koyama and
(Pagrus major) (1-2 min (34.5 ppt) hydrocarbons Kakuno 2004
mixing, 5 (?)/L
min settling;
no renewal)

Prestige- Fresh and Thicklip grey mul- Juvenile Gene expres- Acute 2 d and Oil mixed Seawater Lab cyp1a and gst tran- Bilbao et al.
like heavy weathered let (Chelon sion and 16 d with sedi- (35 ppt) scription values of 2010
fuel oil (2.5 h water labrosus) chronic ments and biotransformation
(IFO 380) washing + placed in enzymes elevated in
aeration) static system F and W treatments
(no renewal)
+ fish

ANS crude Yes; heated to Pacific halibut Juvenile LOEC: Chronic 90 d 1 and 2 % Seawater Lab 1600 ug/g in sedi- Moles and
80 °C for 8 h (Hippoglossoides growth; v/v oil ment Norcross
stenolepis) histopathology mixed in 1998
sediments
ANS crude Yes; heated to Rock sole (P. bi- Juvenile LOEC: Chronic 90 d 1 and 2 % Seawater Lab 1600 ug/g in sedi- Moles and
80 °C for 8 h lineatus) growth; v/v oil ment Norcross
histopathology mixed in 1998
sediments

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 401
Oil type Weathered? Species Life Response Type Time Method Fresh/ Lab/ Nominal Measured Notes Reference
stage Seawater Field
ANS crude Yes; heated to Yellowfin sole Juvenile LOEC: growth, Chronic 90 d 1 and 2 % v/v Seawater Lab 1600 ug/g in 34-56% reduced Moles and
80 °C for 8 h (Pleuronectes histopathology oil mixed in sediment growth in all flat- Norcross
asper) sediments fishes, occurring at 1998
sediment concen-
trations of 1600
ug/g
Bunker C Yes, natural- Winter flounder Juvenile Hepatic MFO activity Chronic 1-14 d Contaminated Seawater Lab LOET (time Overall, MFO Lee et al. 2003
ly in field (Pleuronectes sediments for EROD induction was
over 30 y americanus) collected induction) = limited; non-toxic
after Arrow 14 d by Microtox; very
oil spill toxic by amphi-
pod survival tests.
California Yes Tidewater silver- Juvenile Mortality Chronic 7d WAF + UV Seawater Lab 7d LC50 = 0.5 Little et al.
crude side (Menidia - 2.8 mg 2000
beryllina) TPH/L

4d LC50 =
1.09 - 1.77
mg/L

LOEC (4d or
7d) = 1.50
mg/L
Erika tank- Yes (mini- Sole (Solea Juvenile DNA adducts Chronic Unknown Field expo- Seawater Field DNA adducts Fish collected Amat et al.
er fuel mum 2 solea) sure to present two along French Brit- 2006
months at spilled oil months after tany coasts (oiled
sea) spill in juve- beaches from 1999
nile sole from Erika oil spill)
contaminated
sites
Heavy Unknown Goldfish Juvenile LOEC (calculated) Chronic 20 d Field collect- Fresh Lab SOD activity: SOD, CAT, GST Wang et al.
crude from (Carassius au- from SOD, CAT, GST, ed oil- LOEC = 10 g in the hepatic 2009
Victory Oil ratus) MDA contaminated oiled sediment/L system are sensi-
Field soils (stirred tive to crude oil
(Shandong daily) and may serve as
CAT activity:
Province, a monitoring
LOEC
China) index
=0.5g/L(?)

GST activity:
LOEC = 0.5 g/L

MDA content:
LOEC =50 g/L
Nigeria No Guinean tilapia Juvenile Mortality Chronic 96 h WAF Fresh Lab crude oil LC50 = Ndimele et al.
crude oil (Tilapia guineen- 125.89 mg oil/L 2010
(Bonny sis)
Light)

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 402
MESA Yes; sparged Mummichog Juvenile EROD induction Acute 48 h WAF, Seawater Lab 15 ppt. - WAF: EC50 WAF/EC50 Ramachandra
with air for (Fundulus heter- CEWAF (15 and 7.56 % v/v; CEWAF (induc- n et al. 2006
130 h at room oclitus) 30 ppt.) CEWAF: tion potential) = 15
T, until 14% 0.022%v/v ppt.: 343; 30 ppt.:
loss 9.56
30 ppt. - WAF:
1.10 % v/v;
CEWAF:
0.115%v/v
Oil type Weathered? Species Life Response Type Time Method Fresh/ Lab/ Nominal Measured Notes Reference
stage Seawater Field
MESA Yes; sparged Rainbow trout Juvenile EROD induction Acute 48 h WAF, Fresh Lab WAF: 0.106 % WAF: 0.72 µg EC50 Ramachandra
with air for (Oncorhynchus CEWAF v/v, CEWAF: TPAH/L, WAF/EC50 n et al. 2004
130 h at room mykiss) 0.001 % v/v CEWAF: 0.60 CEWAF (induc-
T, until 14% µg TPAH/L tion potential) =
loss 106
MESA Yes; sparged Rainbow trout Juvenile EROD induction Acute 48 h WAF, Fresh and Lab Fresh - WAF: EC50 WAF/EC50 Ramachandra
with air for (Oncorhynchus CEWAF seawater 0.088 % v/v; CEWAF (induction n et al. 2006
130 h at room mykiss) (15 ppt) CEWAF: potential) = Fresh:
T, until 14% 0.00034 % v/v 258; 15 ppt: 276
loss 15 ppt - WAF:
4.98 % v/v;
CEWAF: 0.018
% v/v

Scotian No Rainbow trout Juvenile EROD induction Acute 48 h WAF, Fresh Lab WAF: 0.390 % WAF: 1.56 µg EC50 Ramachandran
Light (Oncorhynchus CEWAF v/v, TPAH/L, WAF/EC50 et al. 2004
mykiss) CEWAF: 2.00 CEWAF (induc-
CEWAF: 0.066 µg TPAH/L tion potential) =
% v/v 5.91
Terra Nova No Rainbow trout Juvenile EROD induction Acute 48 h WAF, Fresh Lab WAF: 3.350 % WAF: 1.80 µg EC50 Ramachandran
(Oncorhynchus CEWAF v/v, TPAH/L, WAF/EC50 et al. 2004
mykiss) CEWAF: 1.50 CEWAF (induc-
CEWAF: 0.003 µg TPAH/L tion potential) =
% v/v 1116
Ultralow No Rainbow trout Juvenile EROD induction Acute 24 h WAF, Fresh Lab CEWAF EC50: Schein et al.
sulfur (Oncorhynchus CEWAF <0.01 % v/v 2009
Diesel No. mykiss)
2
WAF EC50: 6.17
% v/v
Diesel No Atlantic croaker Juvenile Histology of pitui- Chronic 5 to 8 w WAF Seawater Lab 30-56% of oil- Thomas and
(Micropogonias tary/hypothalamus and (Anderson exposed fish failed Budiantara
undulatus) ovarian tissues 1974; dosed to reach sexual 1995
every two maturation; ovarian
days) growth was im-
paired, slowed
development of
ovarian follicles
and widespread
atresia; few oo-
cytes.
Grand No Winter flounder Juvenile, Mortality, histo- Chronic 8w Artificially Seawater Lab 300 mg/kg sedi- Khan 1995
Banks (Pleuronectes adult pathology of gills, contaminated ment
crude americanus) reproduction sediments

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 403
Oil type Weathered? Species Life stage Response Type Time Method Fresh/ Lab/ Nominal Measured Notes Reference
Seawater Field
Prudhoe Bay Unknown High cockscomb Juvenile, CYP1A Chronic 28 d Caged in Seawater Lab 6-fold greater induc- Woodin et al.
crude (Anoplarchus adult induction field at oiled and tion in field, 49-fold 1997
purpurescens) sites, or field greater induction in lab
exposed in
lab to field-
collected oil-
contaminat-
ed sediments
and/or oiled
food

Scotian Light No Rainbow trout Embryo Embryo Chronic 25 d WAF, Fresh Lab 0.03 %v/v (Calculated) Wu et al.
(Oncorhynchus deformities CEWAF 2.1 µg 2012
mykiss) TPAH/L
Federated No Rainbow trout Embryo Embryo Chronic 25 d WAF, Fresh Lab 0.015 %v/v (Calculated) Wu et al.
crude (Oncorhynchus deformities CEWAF 2.7 µg 2012
mykiss) TPAH/L
ANS crude No Rainbow trout Embryo Embryo Chronic 25 d WAF, Fresh Lab 0.02 %v/v (Calculated) Wu et al.
(Oncorhynchus deformities CEWAF 3.4 µg 2012
mykiss) TPAH/L
MESA No Rainbow trout Embryo Embryo Chronic 25 d WAF, Fresh Lab 0.015 %v/v (Calculated) Wu et al.
(Oncorhynchus deformities CEWAF 2.1 µg 2012
mykiss) TPAH/L
Louisiana Yes, mixed Gulf killifish Adult Mortality Acute 7 day WAF 1:10 Brackish Lab Range finding experi- Pilcher et al.
crude with clean (Fundulus gran- males oil:water, ment 2014
brackish dis) 100%
water for
30days, water
fraction iso-
lated
Louisiana Yes, mixed Gulf killifish Adult Mortality Acute 7 day WAF, 75% Brackish Lab Range finding experi- Pilcher et al.
crude with clean (Fundulus gran- males ment 2014
brackish dis)
water for 30
days, water
fraction iso-
lated
Louisiana Ye s, mixed Gulf killifish Adult Mortality Acute 7 day WAF, 50% Brackish Lab Range finding experi- Pilcher et al.
crude with clean (Fundulus gran- males ment 2014
brackish dis)
water for 30
days, water
fraction iso-
lated
Louisiana Yes, mixed Gulf killifish Adult No Mortality Acute 7 day WAF, 25% Brackish Lab Range finding experi- Pilcher et al.
crude with clean (Fundulus gran- males ment 2014
brackish dis)
water for 30-
40 days,
water fraction

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 404
Oil type Weathered? Species Life stage Response Type Time Method Fresh/ Lab/ Nominal Measured Notes Reference
Seawater Field
isolated

Louisiana Yes, mixed Gulf killifish Adult DNA strand Acute 7 day WAF, 25% Brackish Lab Definitive exposure Pilcher et al.
crude with clean (Fundulus gran- males breakage experiment 2014
brackish dis) elevated for,
water for 40 high and
days, water moderate-low
fraction iso- damage
lated categories
Louisiana Yes, mixed Gulf killifish Adult Non statisti- Acute 7 day WAF, 2.5% Brackish Lab Definitive exposure Pilcher et al.
crude with clean (Fundulus gran- males cally signifi- experiment 2014
brackish dis) cant effects
water for 40
days, water
fraction iso-
lated
Phenanthrene Used Phenan- Harpacticoid Adult Sub-lethal Acute 10 day Phenan- 25 ‰ Lab LC50 345 Lotufo and
threne (98% copepod (Schiz- female effects: sig- bioassay threne was artificial µg g-1 dry Fleeger 1997
purity, Al- opera knabeni) nificantly added to seawater wet (291 -
drich Chemi- fewer real- sediment 407)
cal Co.) ized offspring slurry
Phenanthrene Used Phenan- Harpacticoid Adult Male Acute 10 day Phenan- 25 ‰ Lab LC50 349 Lotufo and
threne (98% copepod (Schiz- bioassay threne was artificial µg g-1 dry Fleeger 1997
purity, Al- opera knabeni) added to seawater wet (291 -
drich Chemi- sediment 417)
cal Co.) slurry
Phenanthrene Used Phenan- Harpacticoid Copepodite Sub-lethal Acute 10 day Phenan- 25 ‰ Lab LC50 172 Lotufo and
threne (98% copepod (Schiz- effects: re- bioassay threne was artificial µg g-1 dry Fleeger 1997
purity, Al- opera knabeni) duced mean added to seawater wet (155 –
drich Chemi- fraction of sediment 190)
cal Co.) realized slurry
offspring
composed of
copepodites
Phenanthrene Used Phenan- Harpacticoid Nauplius Reduced Acute 10 day Phenan- 25 ‰ Lab LC50 84 µg Lotufo and
threne (98% copepod (Schiz- mean hatch- bioassay threne was artificial g-1 dry wet Fleeger 1997
purity, Al- opera knabeni) ing number at added to seawater (74 – 96)
drich Chemi- 90 µg g-1 dry sediment
cal Co.) wt slurry
Phenanthrene Used Phenan- Harpacticoid Adult Survival Acute 10 day Phenan- 25 ‰ Lab LC50 105 Lotufo and
threne (98% copepod (Nitocra female lower at >177 bioassay threne was artificial µg g-1 dry Fleeger 1997
purity, Al- lacustris) µg g-1 dry added to seawater wet (95 –
drich Chemi- weight con- sediment 116)
cal Co.) centration, slurry
reduced
proportional
realized
offspring and
fraction of
realized
offspring
comprised of
copepodites

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 405
Oil type Weathered? Species Life stage Response Type Time Method Fresh/ Lab/ Nominal Measured Notes Reference
Seawater Field
at 45 and 90
µg g-1 dry wt

Phenanthrene Used Phenan- Harpacticoid Adult male Survival Acute 10 day Phenan- 25 ‰ Lab LC50 72 µg Lotufo and
threne (98% copepod (Nitocra lower at > 90 bioassay threne was artificial g-1 dry wet Fleeger 1997
purity, Al- lacustris) µg g-1 dry added to seawater (62 – 83)
drich Chemi- weight con- sediment
cal Co.) centration slurry
Phenanthrene Used Phenan- Harpacticoid Copepodite Survival Acute 10 day Phenan- 25 ‰ Lab LC50 43 µg Lotufo and
threne (98% copepod (Nitocra lower at >22 bioassay threne was artificial g-1 dry wet Fleeger 1997
purity, Al- lacustris) µg g-1 dry added to seawater (36-52)
drich Chemi- weight con- sediment
cal Co.) centration slurry
Phenanthrene Used Phenan- Harpacticoid Nauplius Survival Acute 10 day Phenan- 25 ‰ Lab LC50 71 µg Lotufo and
threne (98% copepod (Nitocra lower at >45 bioassay threne was artificial g-1 dry wet Fleeger 1997
purity, Al- lacustris) µg g-1 dry added to seawater (65-77)
drich Chemi- weight con- sediment
cal Co.) centration slurry
Hibernia 200mL of Atlantic cod Adult male Significant Chronic 89 days WAF July – Seawater Lab Mean total hydrocar- Khan 2012
crude oil crude in an (Gadus morhua) changes in October, 9- bon (ppb) concentra-
80L tank with gonadal so- 14 degrees tion (THC) = 49 ± 16
inflow matic index Celsius
2L/min, WAF
drawn off
bottom
Hibernia 200mL of Atlantic cod Adult Significantly Chronic 89 days WAF July – Seawater Lab Mean total hydrocar- Khan 2012
crude oil crude in an (Gadus morhua) female less weight October, 9- bon (ppb) concentra-
80L tank with gain, signifi- 14 degrees tion (THC) = 49 ± 16
inflow cant changes Celsius
2L/min, WAF in gonadal
drawn off somatic index
bottom
Hibernia 200mL of Atlantic cod Adult male Significant Chronic 86 days WAF Au- Seawater Lab Mean total hydrocar- Khan 2012
crude oil crude in an (Gadus morhua) changes in gust – No- bon (ppb) concentra-
80L tank with gonadal so- vember 10- tion (THC) = 42 ± 19
inflow matic index 14 degrees
2L/min, WAF Celsius
drawn off
bottom
Hibernia 200mL of Atlantic cod Adult Significantly Chronic 86 days WAF Au- Seawater Lab Mean total hydrocar- Khan 2012
crude oil crude in an (Gadus morhua) female less weight gust – No- bon (ppb) concentra-
80L tank with gain, signifi- vember 10- tion (THC) = 42 ± 19
inflow cant changes 14 degrees
2L/min, WAF in gonadal Celsius
drawn off somatic index
bottom
Hibernia 200mL of Atlantic cod Adult male Significant Chronic 92 days WAF, Sep- Seawater Lab Mean total hydrocar- Khan 2012
crude oil crude in an (Gadus morhua) changes in tember – bon (ppb) concentra-
80L tank with gonadal so- November tion (THC) = 40 ± 12
inflow matic index, 10-14 de-
2L/min, WAF Immature grees Celsi-
drawn off testes small us
bottom spermato-
cytes, 90%

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 406
Oil type Weathered? Species Life stage Response Type Time Method Fresh/ Lab/ Nominal Measured Notes Reference
Seawater Field
external le-
sions

Hibernia 200mL of Atlantic cod Adult Immature Chronic 92 days WAF, Sep- Seawater Lab Mean total hydrocar- Khan 2012
crude oil crude in an (Gadus morhua) female ovaries, small tember – bon (ppb) concentra-
80L tank with oocytes, 60% November tion (THC) = 40 ± 12
inflow external le- 10-14 de-
2L/min, WAF sions grees Celsi-
drawn off us
bottom
Hibernia 200mL of Atlantic cod Adult male Significant Chronic 86 days WAF, No- Seawater Lab Mean total hydrocar- Khan 2012
crude oil crude in an (Gadus morhua) changes in vember – bon (ppb) concentra-
80L tank with gonadal so- February 0-2 tion (THC) = 31 ± 14
inflow matic index, degrees
2L/min, WAF small sper- Celsius
drawn off matocytes,
bottom disrupted
development,
40% external
lesions
Hibernia 200mL of Atlantic cod Adult Significant Chronic 86 days WAF No- Seawater Lab Mean total hydrocar- Khan 2012
crude oil crude in an (Gadus morhua) female changes in vember – bon (ppb) concentra-
80L tank with gonadal so- February 0-2 tion (THC) = 31 ± 14
inflow matic index, degrees
2L/min, WAF small oocytes Celsius
drawn off disrupted
bottom development,
33% external
lesions
Hibernia 200mL of Atlantic cod Adult male Significant Chronic 86 days WAF De- Seawater Lab 5 Mean total hydrocar- Khan 2012
crude oil crude in an (Gadus morhua) changes in cember – bon (ppb) concentra-
80L tank with gonadal so- February 0-2 tion (THC) = 15 ± 10
inflow matic index, degrees
2L/min, WAF small sper- Celsius
drawn off matocytes,
bottom disrupted
development,
53% external
lesions
Hibernia 200mL of Atlantic cod Adult Small oocytes Chronic 86days WAF De- Seawater Lab Mean total hydrocar- Khan 2012
crude oil crude in an (Gadus morhua) female disrupted cember – bon (ppb) concentra-
80L tank with development, February 0-2 tion (THC) = 15 ± 10
inflow 37 % external degrees
2L/min, WAF lesions Celsius
drawn off
bottom
Hibernia 200mL of Atlantic cod Adult male 50 % external Chronic 38 days WAF, April Seawater Lab Mean total hydrocar- Khan 2012
crude oil crude in an (Gadus morhua) lesions – May 1-3 bon (ppb) concentra-
80L tank with degrees tion (THC) = 22 ± 12
inflow Celsius
2L/min, WAF
drawn off
bottom
Hibernia 200mL of Atlantic cod Adult 40 % external Chronic 38 days WAF April Seawater Lab Mean total hydrocar- Khan 2012

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 407
Oil type Weathered? Species Life stage Response Type Time Method Fresh/ Lab/ Nominal Measured Notes Reference
Seawater Field
crude oil crude in an (Gadus morhua) female lesions – May 1-3 bon (ppb) concentra-
80L tank with degrees tion (THC) = 22 ± 12
inflow Celsius
2L/min, WAF
drawn off
bottom
Hibernia 200mL of Atlantic cod Adult Atretic oo- Chronic 90 days WAF , June Seawater Lab Mean total hydrocar- Khan 2012
crude oil crude in an (Gadus morhua) female cytes, post- – June 6-12 bon (ppb) concentra-
80L tank with ovulatory degrees tion (THC) = 28 ± 9
inflow follicles, Celsius
2L/min, WAF disrupted
drawn off development
bottom 57 % external
lesions
Hibernia 200mL of Atlantic cod Adult Significant Chronic 4 days WAF , Seawater Lab Mean total hydrocar- Khan 2012
crude oil crude in an (Gadus morhua) female changes in October – bon (ppb) concentra-
80L tank with gonadal so- March 0-10 tion (THC) = 21 ± 8
inflow matic index, degrees
2L/min, WAF disrupted Celsius
drawn off development,
bottom small oocytes
56% external
lesions
Heavy fuel WAF Rainbow trout Embryo PAH concen- Chronic 25 days WAF 1:9 Fresh Lab WAF: >0.01% Martin et al.
oil 6303 (Oncorhynchus tration to water: oil v/v, LC 50 >45 2014
mykiss) cause toxicity ratio µg /L TPH-F,
>5 µg /L
TPAH-F
Heavy fuel Chemically Rainbow trout Embryo PAH concen- Chronic 25 days Chemically Fresh Lab CEWAF: Martin et al.
oil 6303 enhanced (Oncorhynchus tration to enhanced 0.0037 % v/v, 2014
mykiss) cause toxicity WAF (1:9 LC 50 473 µg
water:oil /L TPH-F,
ratio then EC50 338 µg
surface /L
application
of dispersant
at disper-
sant: oil ratio
1:10)
Heavy fuel Oil poured on Rainbow trout Embryo Chronic 25 days Contaminat- Fresh Lab LC50 505 µg Martin et al.
oil 6303 gravel, (Oncorhynchus ed gravel in oil/g gravel, 2014
mykiss) columns EC50 49 µg/L
from which
effluent
taken as
WAF
Heavy fuel Artificially Rainbow trout Embryo Chronic 25 days Contaminat- Fresh Lab LC50 591 µg Martin et al.
oil 6303 weathered- (Oncorhynchus ed gravel in oil/g gravel, 2014
stranded on mykiss) columns EC50 66 µg/L
gravel from which
effluent
taken as
WAF

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 408
Oil type Weathered? Species Life stage Response Type Time Method Fresh/ Lab/ Nominal Measured Notes Reference
Seawater Field
Heavy fuel Oil stranded Rainbow trout Embryo Chronic 25 days Contaminat- Fresh Lab LC 50 1031 µg Martin et al.
oil 7102 on gravel (Oncorhynchus ed gravel in oil/g gravel, 2014
mykiss) columns EC50 120 µg/L
from which
effluent
taken as
WAF
MESA crude Oil stranded Rainbow trout Embryo Chronic 25 days Contaminat- Fresh Lab LC 50 4607 µg Martin et al.
on gravel (Oncorhynchus ed gravel in oil/g gravel, 2014
mykiss) columns EC50 1076
from which µg/L
effluent
taken as
WAF
Heavy fuel WAF pro- Rainbow trout Embryo Induced liver Chronic 24 days High energy, Fresh Lab LC 115 µg /L, Adams et al.
oil 7102 duced from (Oncorhynchus EROD activi- chemically 48-h EC50 2014b
flow of water mykiss) ty to a level enhanced 6.03 µg TPH-
through oiled equivalent to WAF F/L
gravel that of β-
naphtho-
flavone
Pyrolytic No Rainbow trout Embryo Significantly Acute 22 days WAF: Em- Freeh Lab Le Bihanic et
PAH fraction (Oncorhynchus more half bryos placed al. 2014
from contam- mykiss) hatched in gravel
inated sedi- embryos and spiked with
ment abnormal PAH frac-
larvae, DNA tion for first
damage ten days
Erika heavy No Rainbow trout Embryo Embryo Acute 22 days WAF: Em- Fresh Lab Le Bihanic et
oil (Oncorhynchus mortality and bryos placed al. 2014
mykiss) decreased in gravel
hatching spiked with
success, PAH frac-
abnormal tion for first
larvae, DNA ten days
damage
Arabian No Rainbow trout Embryo Reduced Acute 22 days WAF: Em- Fresh Lab Le Bihanic et
Light oil (Oncorhynchus larval size, bryos placed al. 2014
mykiss) abnormal in gravel
larvae, hem- spiked with
orrhages, PAH frac-
DNA damage tion for first
ten days
PAH No Haddock (Mela- Adult Elevated Chronic Fish sam- Seawater Field – Balk et al
nogrammus ae- concentration pled from North 2011
glefinus) of bile PAH natural pop- Sea
metabolites ulations in
with 2,3 and areas with
4 rings, and 2 extensive oil
hydroxy- production
naphthalene.
Altered En-
zyme activi-
ty, increased

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 409
Oil type Weathered? Species Life stage Response Type Time Method Fresh/ Lab/ Nominal Measured Notes Reference
Seawater Field
hepatic DNA
adducts

PAH No Atlantic cod Adult Elevated Chronic Fish sam- Seawater Field – Significantly Balk et al
(Gadus morhua) concentration pled from North elevated 2011
of bile PAH natural pop- Sea EROD ac-
metabolites ulations in tivity
Enzyme areas with
activity in- extensive oil
duced by production
oxidative
stress caused
by petroleum
pollutants
Orimulsion- No Atlantic herring Embryo Survival Acute 24h WAF Artificial Lab LC50 0.00259- Boudreau et
400 (Clupea ha- reduced at seawater 0.0375% al. 2009
(PDVSA- rengus) 0.0032%
BITOR) concentration
Orimulsion- No Mummichog Embryo Survival Acute 24h WAF Artificial Lab LC50 0.0421- Boudreau et
400 (Fundulus heter- reduced at seawater 0.0478, EC50 al. 2009
(PDVSA- oclitus) 0.1 and .0082-0.0157%
BITOR) 0.032% con-
centration,
significant
increase in
development
abnormali-
ties, smaller
larvae
No. 6 fuel oil No Mummichog Embryo Survival Acute 24h WAF Artificial Lab LC50 2.81% - Boudreau et
(Fundulus heter- reduced at seawater 6.12% al. 2009
oclitus) 3.2, 10 and
18% concen-
tration in
assay, ab-
normal larvae
PAHs Spill of bun- Lake Whitefish Larvae Significantly Chronic Wabamun Fresh Field – Comparing oil exposed Debruyn et al.
(naphtha- ker C oil and (Coregonus clu- increased Lake water in situ vs. reference sites, 2007
Imperial pole peaformis) level of ky- PAH’s detected at oil-
lenephenan-
treating oil, phosis and exposed sites
threne);
plus constant skeletal and
Wabamun
inputs from fin deformi-
Lake
coal plant ties
Pyrene No Daphnia magna Highly toxic Acute 24h in 1 and 10 Fresh Lab EC50 0.06 uM Accumulation of Huovinen et
in presence solution, µg/L chemical prior to expo- al. 2001
of UV radia- 15 min sure important to tox-
tion UV icity
exposure
Anthracene No Daphnia magna Photo in- Acute 24h in 10 µg /L Fresh Lab EC50 0.22uM Accumulation of Huovinen et
duced toxici- solution, chemical prior to expo- al. 2001
ty 15 min sure important to tox-
UV icity
exposure

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 410
Oil type Weathered? Species Life stage Response Type Time Method Fresh/ Lab/ Nominal Measured Notes Reference
Seawater Field
Retene No Daphnia magna Photo in- Acute 24h in 10 and 32 Fresh Lab EC50 1.54 uM Accumulation of Huovinen et
duced toxici- solution, µg/L chemical prior to expo- al. 2001
ty 15 min sure important to tox-
UV icity
exposure
Phenanthrene No Daphnia magna Not photo- Acute 24h in 10, 32, 100, Fresh Lab EC50 >1.80 Accumulation of Huovinen et
toxic solution, 320 µg /L uM chemical prior to expo- al. 2001
15 min sure important to tox-
UV icity
exposure
Dehydro- No Daphnia magna Embryo Not photo- Acute 24h in 100 and 320 Fresh Lab Huovinen et
abietic acid toxic solution, µg /L al. 2001
15 min
UV
exposure
Napthalene No Zebrafish (Danio Embryo 100% mortal- Acute 24 – 48h Passive Fresh Lab PAC 27193.2 ± Seilor et al.
rerio) ity 24h expo- dosing via 936 2014
sure PDMS sili-
cone
Acenaph- No Zebrafish (Danio Embryo 100% mortal- Acute 24 – 48h Passive Fresh Lab Seilor et al.
thene rerio) ity 24h expo- dosing via 2014
sure PDMS sili-
cone
Fluorene No Zebrafish (Danio Embryo 100% mortal- Acute 24 – 48h Passive Fresh Lab Seilor et al.
rerio) ity 24h expo- dosing via 2014
sure PDMS sili-
cone
Phenanthrene No Zebrafish (Danio Embryo 100% mortal- Acute 24 – 48h Passive Fresh Lab Seilor et al.
rerio) ity 24h expo- dosing via 2014
sure PDMS sili-
cone
Fluoranthene No Zebrafish (Danio Embryo 100% mortal- Acute 24 – 48h Passive Fresh Lab Seilor et al.
rerio) ity 24h expo- dosing via 2014
sure PDMS sili-
cone
Pyrene No Zebrafish (Danio Embryo 50% mortali- Acute 24 – 48h Passive Fresh Lab Seilor et al.
rerio) ty 24h expo- dosing via 2014
sure PDMS sili-
cone
Cosco Busan Artificial Pacific herring Embryo Pericardial Chronic 8 days WAF: Seawater Lab LC50 11 – 138, High losses of herring Incardona et
bunker oil (Clupea pallasi) edema after Embryos mix 22 EC20 for car- following spill in San al 2012
exposure to placed in psu salin- diotoxicity Francisco Bay, CBBO
1.0 g/kg oiled gravel ity 300-1000 ng/L interacts with sunlight
gravel efflu- weathered PACs to produce acutely
ent, photo- with con- lethal necrosis at low
toxicity in 0.3 tinuous water and tissue PAC
and 1.0g/kg seawater concentrations
UV-t
Alaskan Artificial Pacific herring Embryo Pericardial Chronic 8 days WAF: Seawater Lab Incardona et
North Slope (Clupea pallasi) edema after Embryos mix 22 al 2012
crude exposure to placed in psu salin-
0.3 and 1.0 oiled gravel ity
g/kg gravel weathered
effluent with con-

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 411
Oil type Weathered? Species Life stage Response Type Time Method Fresh/ Lab/ Nominal Measured Notes Reference
Seawater Field
treatments tinuous
phototoxicity seawater
in 1.0g/kg
UV-t
Cosco Busan Artificial Zebrafish (Danio Embryo Cytotoxicity, Chronic 24 h WAF: Seawater Lab Pyrene, fluoranthene, Incardona et
bunker oil rerio) cellular Embryos mix 22 and chrysene contrib- al 2012
membrane placed in psu salin- ute to toxicity
damage, oiled gravel ity
mortality weathered
with con-
tinuous
seawater
Alaska North Artificial: Zebrafish (Danio Embryo - Reduced Chronic 48h WAF: Oiled Artificial Lab Exposed as embryos Hicken et al.
Slope crude water passed rerio) Adult larval surviv- gravel efflu- seawater then reared for 10-11 2011
oil through col- al, (5 to 17% ent months to detect ef-
umns con- increase fects on adults, small
taining 6 g/kg mortality), group exposed again at
of oil and changes in day 97 and found
gravel ventricular CYP1A induction
shape corre- played protective role.
lated with
reduced adult
swimming
performance,
reduced
cardiac out-
put
Pyrolitic No Zebrafish (Danio Embryo - Slower Chronic 9 months Food spiked Fresh Lab Vignet et al.
fraction rerio) Adults growth in with com- 2014a, b
mass and plex mix-
length; fe- ture of
males with PAHs
lower body
mass; disrup-
tion of jaw
growth in
larvae and
malformation
in adults;
specific
activities of
digestive
enzymes
reduced;
adult behav-
iour disrupt-
ed.
Erika heavy No Zebrafish (Danio Embryo – Slower Chronic 9 months Food spiked Fresh Lab Vignet et al.
oil rerio) Adults growth in with com- 2014a, b
mass and plex mix-
length; males ture of
with smaller PAHs
body mass;
disruption of
jaw growth in
282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 412
Oil type Weathered? Species Life stage Response Type Time Method Fresh/ Lab/ Nominal Measured Notes Reference
Seawater Field
larvae and
malformation
in adults;
specific
activities of
digestive
enzymes
reduced;
adult behav-
iour disrupt-
ed; most
affected
relative to
light oil and
pyrolytic
fraction.
Arabian light No Zebrafish (Danio Embryo - Slower Chronic 9 months Food spiked Fresh Lab Vignet et al.
crude oil rerio) Adults growth in with com- 2014a,b
mass and plex mix-
length; males ture of
with reduced PAHs
body mass;
disruption of
jaw growth in
larvae and
malformation
in adults;
digestive
enzyme
alkaline
phosphatase
recovery
affected.
MESA Yes Atlantic herring Embryo Blue sac Chronic 13 – 19 WAF/CEW Salt (15 Lab TPH - 1.02 The toxicities of WAF Adams et al.
(Clupea ha- disease days AF ppt) mg/L and CEWAF were the 2014a
rengus) same when expressed
as measured concentra-
tions of TPH

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 413
REFERENCES

Chapter 1

AER (Alberta Energy Regulator). 2013. Pipeline Performance in Alberta 1990-2012. Report 2013-B. August, 2013.
https://www.aer.ca/documents/reports/R2013-B.pdf.
AER. 2014a. Pace Oil and Gas Ltd. Wellhead Piping Failure Licence No. W0057420. May 19, 2012. Alberta
Energy Regulator Investigation Report. https://www.aer.ca/documents/reports/IR_20140303-
PaceOilandGas.pdf
AER. 2014b. Plains Midstream Canada ULC NPS 12 Rangeland South. Pipeline Failures and Release into the Red
Deer River. Licence No. 5844, Line 1. June 7, 2012. Alberta Energy Regulator Investigation Report.
https://www.aer.ca/documents/reports/IR_20140304-PlainsRangeland.pdf.
Anderson, C.M., Mayes, M. and R. Labelle. 2012. Update of Occurrence Rates for Offshore Oil Spills. USDI
BOEM and BSEE, Herndon, VA.
API (American Petroleum Institute). 2009. Analysis of U.S. Spillage. API Publication No. 356. American Petroleum
Institute, Washington, D.C.
Boudreau, P.R., Gordon, D.C. Jr., Harding, G.C., Loder, J.W., Black, J., Bowen, W.D., Campana, S., Cranford, P.J.,
Drinkwater, K.F., Van Eeckhaute, L., Gavaris, S., Hannah, C.G., Harrison, G., Hunt, J.J. McMillan, J.,
Melvin, G.D., Milligan, T.G., Muschenheim, D.K., Neilson, J.D., Page, F.H., Pezzack, D.S., Robert, G.,
Sameoto, D. and H. Stone. 1999. The Possible Environmental Impacts of Petroleum Exploration Activities
on the Georges Bank Ecosystem. Canadian Technical Report of Fisheries and Aquatic Sciences 2259.
CBC. 2013. Tanker truck crashes into BC creek. CBC News July 27, 2013. http://www.cbc.ca/news/canada/british-
columbia/jet-fuel-spill-evacuation-order-lifted-in-b-c-1.1306998.
Christopherson, S. and K. Dave. 2014. A New Era of Crude Oil Transport: Risks and Impacts in the Great Lakes
Basin, Community and Regional Development Institute Reports. Issue Number 15. Department of
Development Sociology. Cornell University.
ERCB (Energy Resources Conservation Board). 2011. Penn West Petroleum Ltd. Well Blowout 14-20-064-
10W5M. Energy Resources Conservation Board Investigation Report.
https://www.aer.ca/documents/reports/IR_20110823_PennWest.pdf
ERCB. 2012. Midway Energy Ltd. Hydraulic Fracturing Incident: Interwellbore Communication. January 13, 2012.
Energy Resources Conservation Board Investigation Report.
https://www.aer.ca/documents/reports/IR_20121212_Midway.pdf.
ERCB. 2013. Plains Midstream Canada ULC NPS 20 Rainbow Pipeline Failure. License No. 5592, Line No. 1.
April 28, 2011. Energy Resources Conservation Board Investigation Report. Calgary, AB.
https://www.aer.ca/documents/reports/IR_20130226-PlainsMidstream.pdf.
EPA. 2015. Galena Train Derailment. U.S. Environmental Protection Agency. EPA in Illinois.
http://www2.epa.gov/il/galena-train-derailment.
Executive Flight Centre. 2014. Lemon Creek Environmental Monitoring Plan Update. August 14, 2014.
http://www.lemoncreekresponse.ca/index.php.
Frittelli, J., Andrews, A., Parkomak, P., Pirog, R., Ramseur, J/L. and M. Ratner. 2014. U.S. Rail Transportation of
Crude Oil: Background and Issues for Congress. Congressional Research Service 7-5700. R43390.
Gillis, L. 2015. Water advisory finally lifted in Gogama derailment area. Timmins Press June 10, 2015.
http://www.timminspress.com/2015/06/10/water-advisory-finally-lifted-in-gogama-derailment-area.
Government of Saskatchewan. 2015. Petroleum and Natural Gas Spill Report Directory.
http://www.qp.gov.sk.ca/Spills/Spill%20Report%20Directory.pdf
Hurley, G.V. and J. Ellis. 2004. Environmental Effects of Exploratory Drilling Offshore Canada: EEM Data and
Literature Review. Final Report Prepared for the Canadian Environmental Assessment Agency Regulatory
Advisory Committee (RAC).
Independent Panel Review. 2014. Independent Panel Review of the Canadian Natural Primrose Flow to Surface
Causation Report. http://www.aer.ca/documents/reports/CNRLPrimrose_PanelReport_201407.pdf
ITOPF. 2014. Oil Tanker Spill Statistics 2014. International Tanker Owners Pollution Federation.
http://www.itopf.com/knowledge-resources/documents-guides/document/oil-tanker-spill-statistics-2013/.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 414
ITOPF. 2015a. Atlantic Empress Case Study. International Tanker Owners Pollution Federation.
http://www.itopf.com/in-action/case-studies/case-study/atlantic-empress-west-indies-1979/).
ITOPF. 2015b. Amoco-Cadiz Case Study. International Tanker Owners Pollution Federation.
http://www.itopf.com/in-action/case-studies/case-study/amoco-cadiz-france-1978/).
ITOPF. 2015c. Odyssey Case Study. International Tanker Owners Pollution Federation. http://www.itopf.com/in-
action/case-studies/case-study/odyssey-off-canada-1988/).
ITOPF. 2015d. Tanker incident statistics. International Tanker Owners Pollution Federation.
http://www.itopf.com/knowledge-resources/data-statistics/statistics/
Lee, K., Prince, R.C., Greer, C.W., Doe, K.G., Wilson, J.E.H., Cobanli, S.E., Wohlgeschaffen, G.D., Alroumi, D.,
King. T. and G.H. Tremblay. 2003. Composition and toxicity of residual Bunker C fuel oil in intertidal
sediments after 30 years. Spill Science and Technology Bulletin 8: 187-199.
LeFever, J. 2008. What's happening at Parshall, North Dakota. North Dakota Department of Mineral Resources
Newsletter 35 (1): 1-2.
Mackenzie, D. 2011. Operation clean-up. New Scientist 212 (2836): 46-49.
Mangione, K. 2015. CN train derails south of Timmins, Ont. CTV News February 15, 2015.
http://www.ctvnews.ca/canada/cn-train-derails-south-of-timmins-ont-1.2237139.
Martin, J.D., Adam, J., Hollebone, B., King, T., Brown, R.S. and P.V. Hodson. 2014. Chronic toxicity of heavy fuel
oils to fish embryos using multiple exposure scenarios. Environmental Toxicology and Chemistry 33: 677-
687.
National Energy Board. 2015. Canadian Crude Oil Exports by Rail. https://www.neb-
one.gc.ca/nrg/sttstc/crdlndptrlmprdct/stt/cndncrdlxprtsrl-eng.html.
NRC (National Research Council). 2003. Oil in the Sea III: Inputs, Fates and Effects. National Academies Press,
Washington, DC. http://www.nap.edu/openbook.php?isbn=0309084385.
NTSB (US National Transportation Safety Board). 2013. Accident Involving Two Freight Trains Casselton, North
Dakota. United States National Transportation Safety Board Preliminary Report.
http://www.ntsb.gov/investigations/AccidentReports/Reports/Casselton_ND_Preliminary.pdf
NTSB. 2014. National Transportation Safety Board Safety Recommendation. January 23, 2014.
http://www.ntsb.gov/safety/safety-recs/RecLetters/R-14-001-003.pdf.
NTSB. 2015. National Transportation Safety Board Docket. NTSB Accident ID DSA14MR004, Dec 30, 2013,
Casselton, ND. Public release date April 27, 2015.
http://dms.ntsb.gov/pubdms/search/hitlist.cfm?docketID=55926&CurrentPage=1&EndRow=15&StartRow
=1&order=1&sort=0&TXTSEARCHT=) .
Owens, E.H., Taylor, E., Marty, R. and D.I. Little. 1993. An inland oil spill response manual to minimize adverse
environmental impacts. In: Proceedings of 1993 International Oil Spill Conference. American Petroleum
Institute, Washington, DC. Pp.105-109.
Owens, E.H., Prince, R.C. and R.B. Taylor. 2008. Natural attenuation of heavy oil on a coarse sediment beach:
results from Black Duck Cove, Nova Scotia, Canada over 35 years following the Arrow oil spill. In:
Proceedings of the Thirty-first AMOP Technical Seminar on Environmental Contamination and Response,
June 3-5, 2008, Calgary, Alberta. pp. 585-599.
Raby, J. 2015. CSX provides update on W.Va. oil train derailment cleanup. Washington Post July 21, 2015.
http://www.washingtontimes.com/news/2015/jul/21/csx-to-discuss-west-virginia-oil-train-derailment-/#!
Riordan Seville, L., Federico-O'Murchuand, S. And T. Connor. 2015. Heimdal, North Dakota, evacuated after fiery
oil train crash. NBC News May 6, 2015. http://www.nbcnews.com/news/us-news/heimdal-north-dakota-
evacuated-after-fiery-oil-train-crash-n354686
SL Ross Environmental Research Ltd. 2014. Guideline Document for Measuring the Biological Effects of
Accidental Oil Spills. Report to Fisheries and Oceans Canada, National Contaminants Advisory Group,
Ottawa, ON.
Spies, R.B., Rice, S.D., Wolfe, D.A. and B.A. Wright. 1996. The effect of the Exxon Valdez oil spill on Alaskan
coastal environment. In: Proceedings of the 1993 Exxon Valdez Oil Spill Symposium. American Fisheries
Society, Bethesda, MD. Pp. 1-16.
Thormann, M.N. and S.E. Bayley. 2008. Impacts of the CN Rail Oil Spill on Softstem Bulrush-Dominated
Lacustrine Marshes in Wabamun Lake. Report submitted to Alberta Environment, Edmonton, AB.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 415
TSB (Transportation Safety Board of Canada). 2001. Crude Oil Pipeline Rupture. Enbridge Pipelines Inc. 864-
Millimetre Line ¾, Mile Post 109.42 Near Hardisty, Alberta. 17 January 2001. Transportation Safety Board
of Canada Pipeline Investigation Report P01H0004. http://www.tsb.gc.ca/eng/rapports-
reports/pipeline/2001/p01h0004/p01h0004.asp.
TSB. 2013a. Runaway and Main-Track Derailment. Montreal, Maine and Atlantic Railway Freight Train MMA-
002 Mile 0.23, Sherbrooke Subdivision. Lac Megantic, Quebec. 06 July 2013. Transportation Safety Board
of Canada Railway Investigation Report R13D0054. http://www.tsb.gc.ca/eng/rapports-
reports/rail/2013/r13d0054/r13d0054.pdf.
TSB. 2013b. Statistical Summary Pipeline Occurrences 2013. Transportation Safety Board of Canada.
http://www.bst-tsb.gc.ca/eng/stats/pipeline/2013/sspo-2013.pdf.
TSB. 2015a. Statistical Summary Railway Occurrences 2014. Transportation Safety Board of Canada.
http://www.tsb.gc.ca/eng/stats/rail/2014/sser-ssro-2014.pdf
TSB 2015b Derailment and Fire of Second Canadian National Crude Oil Train Near Gogama, Ontario.
http://www.tsb.gc.ca/eng/medias-media/communiques/rail/2015/r15h0021-20150317.asp
TSPS (Tanker Safety Panel Secretariat). 2013. A Review of Canada’s Ship-Source Oil Spill Preparedness and
Response Regime: Setting the Course for the Future. Ottawa.
https://www.tc.gc.ca/media/documents/mosprr/transport_canada_tanker_report_accessible_eng.pdf
Van Pelt, K. 2015. Update 5 – CSX train hauling North Dakota oil derails, cars ablaze in W. Virginia. Reuters
February 17, 2015. http://uk.reuters.com/article/2015/02/17/usa-train-derailment-csx-
idUKL1N0VQ0P820150217.
Wilhelm, S.I., Robertson, G.J., Ryan, P.C. and D.C. Schneider. 2007. Comparing an estimate of seabirds at risk to a
mortality estimate from the November 2004 Terra Nova FPSO oil spill. Marine Pollution Bulletin 54: 537-
544.
WSP. 2014. Risk Assessment for Marine Spills in Canadian Water: Phase 1, Oil Spills South of the 60th Parallel.
Report to Transport Canada. http://files.wspdigital.com/risk/oil/english/131-17593-00_ERA_Oil-Spill-
South_150116.pdf.
Young, L. 2014. Crude oil spills are bigger from trains than pipelines. Global News January 8, 2014.
http://globalnews.ca/news/1069624/how-do-crude-spills-compare-by-rail-truck-pipeline-you-may-be-
surprised/

Chapter 2

Abbasnezhad, H., Foght, J.M. and M.R. Gray. 2011a. Adhesion to the hydrocarbon phase increases phenanthrene
degradation by Pseudomonas fluorescens LP6a. Biodegradation 22: 485-496.
Abbasnezhad, H., Gray, M. and J.M. Foght. 2011b. Influence of adhesion on aerobic biodegradation and
bioremediation of liquid hydrocarbons. Applied Microbiology and Biotechnology 92: 653-675.
Advanced Technology, Inc. and Continental Shelf Associates, Inc. 1990. The Physical Persistence of Spilled Oil: An
Analysis of Oil Spills Previous to Exxon Valdez. National Oceanic and Atmospheric Administration
(NOAA). http://www.oil-spill-info.com/Publications/1990_Persistence_Spilled_Oil_NOAA_rpt.pdf
(accessed June 2015)
Allen, E. W. 2008. Process water treatment in Canada’s oil sands industry: I. Target pollutants and treatment
objectives. Journal of Environmental Engineering and Science 7: 123–138.
AMAP (Arctic Monitoring and Assessment Programme). 1998. Chapter 10: Petroleum Hydrocarbons. AMAP
Assessment Report: Arctic Pollution Issues. Oslo, Norway. http://www.amap.no/documents/doc/amap-
assessment-report-arctic-pollution-issues/68
American Fuel & Petrochemical Manufacturers. 2014. A Survey of Bakken Crude Oil Characteristics Assembled for
the U.S. Department of Transportation. http://www.ourenergypolicy.org/wp-
content/uploads/2015/06/Survey-of-Crude-Oil-Characteristics_FINAL-1.pdf (accessed June, 2015)
Andersen, S.I. and J.G. Speight. 2001. Petroleum resins: Separation, character and role in petroleum. Petroleum
Science and Technology 19:1-34
AOSRT (Arctic Oil Spill Response Technology – Joint Industry Programme). 2014. Environmental Impacts of

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 416
Arctic Oil Spills and Arctic Oil Spill Response Technologies. Literature Review and Recommendations.
Arctic Oil Spill Response Joint Industry Programme. http://www.arcticresponsetechnology.org/research-
projects.
Atlas, R.M. and T.C. Hazen. 2011. Oil biodegradation and bioremediation: A tale of the two worst spills in U.S.
History. Environmental Science & Technology 45: 6709-6715.
Azam, F. and F. Malfatti. 2007. Microbial structuring of marine ecosystems. Nature Reviews Microbiology 5:782-
792.
Bertoncini, F., Courtiade Tholance, M. and D. Thiebaut. 2013. Gas Chromatography and 2D-Gas Chromatography
for Petroleum Industry: The Race for Selectivity. IFP Energies Nouvelles Publications. Editions Technip,
Paris.
Blenkinsopp, S., Wang, A., Foght, J., Westlake, D.W.S., Sergy, G., Fingas, M., Landriault, M., Sigouin, L. and K.
Semple. 1996. Assessment of the Freshwater Biodegradation Potential of Oils Commonly Transported in
Alaska. Environment Canada Report. ASPS# 95:0065.
Bobra, M. and E.J. Tennyson. 1989. Photooxidation of petroleum. In: Boehm, P.D. and J.W. Farrington (eds.)
Proceedings of the 12th Arctic and Marine Oilspill Program (AMOP) Technical Seminar. Environmental
Protection Service, Ottawa, Canada, pp. 129-147.
Boek, E.S., Headen, T.F. and J.T. Padding. 2010. Multi-scale simulation of asphaltene aggregation and deposition in
capillary flow. Faraday Discussions 144: 271-284
Boufadel, M.C., Sharifi, Y., Van Aken, B., Wrenn, B.A. and K. Lee. 2010. Nutrient and oxygen concentrations
within the sediments of an Alaskan beach polluted with the Exxon Valdez oil spill. Environmental Science
& Technology 44: 7418-7424.
BP (British Petroleum). 2013. Gulf Science Data MC-252 Oil Characterization Data File. Reference No. O-01v01-
01. http://gulfsciencedata.bp.com/go/doctype/6145/178706 (accessed June 2015)
Brakstad, O.G., Nordtug, T. and M. Throne-Holst. 2015. Biodegradation of dispersed Macondo oil in seawater at
low temperature and different oil droplet sizes. Marine Pollution Bulletin 93 (1-2): 144-152.
Brandvik, P.J., and L.-G. Faksness. 2009. Weathering processes in Arctic oil spills: Meso-scale experiments with
different ice conditions. Cold Regions Science and Technology 55: 160-166.
Brooks, G.R., Larson, R.A., Schwing, P.T., Romero, I., Moore, C., Reichart, G.-J., Jilbert, T., Chanton, J.P.,
Hastings, D.W., Overholt, W.A., Marks, K.P. Kostka, J.E., Holmes, C.W. and D. Hollander. 2015.
Sedimentation pulse in the NE Gulf of Mexico following the 2010 DWH blowout. PLOS One 10(7):
e0132341.
Buskey, N. 2013. 40,000-pound tar mat unearthed on Grand Terre. Houma Today, Houma, LA. http://www.
houmatoday.com/article/20130703/ARTICLES/130709834. (accessed Sept, 2015)
Caumette, G., Lienemann, C.-P., Merdrignac, I., Bouyssiere, B., and R. Lobinski. 2009. Element speciation analysis
of petroleum and related materials. Journal of Analytical Atomic Spectrometry 24: 263-276.
Chanton, J., Zhao, T., Rosenheim, B.E., Joye, S., Bosman, S., Brunner, C., Yeager, K.M., Diercks, A.R. and D.
Hollander. 2014. Using natural abundance radiocarbon to trace the flux of petrocarbon to the seafloor
following the Deepwater Horizon oil spill. Environmental Science & Technology 49: 847-854.
Chronopoulou, P.M., Sanni, G.O., Silas-Olu, D.I., van der Meer, J.R., Timmis, K.N., Brussard, C.P.D. and T.J.
McGenity. 2015. Generalist hydrocarbon-degrading bacterial communities in the oil-polluted water
column of the North Sea. Microbial Biotechnology 8: 434-447.
Clemente, J.S. and P.M. Fedorak. 2005. A review of the occurrence, analyses, toxicity, and biodegradation of
naphthenic acids. Chemosphere 60: 585–600.
Crosby, S., Fay, R., Groark, C., Kani, A., Smith, J. R., Sullivan, T. and R. Pavia. 2013. Transporting Alberta Oil
Sands Products: Defining the Issues and Assessing the Risks. U.S. Dept. of Commerce, NOAA Technical
Memorandum NOS OR&R 43. Emergency Response Division, NOAA, Seattle, WA.
http://permanent.access.gpo.gov/gpo45878/noaa_oil_sands_report_09.2013.pdf
Crude Quality Inc. 2015. crudemonitor.ca. http://crudemonitor.ca. (accessed June 2015)
Daling, P.S., Moldestad, M.Ø., Johansen, Ø., Lewis, A. and R. Rødal. 2003. Norwegian testing of emulsion

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 417
properties at sea – the importance of oil type and release conditions. Spill Science & Technology Bulletin
8(2): 123-136.
Davenport, C. 2015. U.S. will allow drilling for oil in Arctic Ocean. New York Times, May 11, 2015.
http://www.nytimes.com/2015/05/12/us/white-house-gives-conditional-approval-for-shell-to-drill-in-
arctic.html?_r=0 (Accessed August 2015)
de Araujo, P.L.B., Mansoori, G.A. and E.S. de Araujo. 2012. Diamondoids: occurrence in fossil fuels, applications
in petroleum exploration and fouling in petroleum production. A review paper. International Journal of
Oil, Gas and Coal Technology 5(4): 316–367.
Deppe, U., Richnow, H.H. Michaelis, W. and G. Antranikian. 2005. Degradation of crude oil by an arctic microbial
consortium. Extremophiles 9:461-470.
Dettman, H., Inman, A. and S. Salmon. 2005. Chemical characterization of GPC fractions of Athabasca bitumen
asphaltenes isolated before and after thermal treatment. Energy & Fuels 19: 1399-1404.
Dettman, H. and G. Irvine. 2015. Comparison of oil-in-water emulsion stability of diluted bitumen, light and heavy
crude oils. Abstract, 38th Arctic and Marine Oilspill Program (AMOP) Technical Seminar. Environment
Canada, Ottawa.
Dew, W.A., Hontela, A., Rood, S.B. and G.G. Pyle. 2015. Biological effects and toxicity of diluted bitumen and its
constituents in freshwater systems. Journal of Applied Toxicology. (in press) doi: 10.1002/jat.3196.
Dickins, D. 2011. Behavior of oil spills in ice and implications for Arctic spill response. Paper OTC 22126. Arctic
Technology Conference Houston, Texas, USA, 7–9 February 2011.
http://www.arcticresponsetechnology.org/wp-content/uploads/2012/11/Dickins-Review-Paper.pdf.
(accessed June, 2015)
Dorobantu, L.S., Yeung, A.K.C., Foght, J.M. and M.R. Gray. 2004.Stabilization of oil-water emulsions by
hydrophobic bacteria. Applied and Environmental Microbiology 70(10): 6333-6336.
Ehrenhauser, F.S., Avij, P., Shu, X., Dugas, V., Woodson, I., Liyana-Arachchi, T., Zhang, A., Hung, F.R. and K.T.
Valsaraj. 2014. Bubble bursting as an aerosol generation mechanism during an oil spill in the deep-sea
environment: laboratory experimental demonstration of the transport pathway. Environmental Science:
Processes and Impacts 16: 65-73
Environmental Technology Centre, Environment Canada. 2015. Oil Properties Database. http://www.etc-
cte.ec.gc.ca/databases/oilproperties/ (accessed June 2015)
Fingas, M.F. 2004. Modeling evaporation using models that are not boundary-layer regulated. Journal of Hazardous
Materials 107: 27-36.
Fingas, M.F. 2014-2015. Bitumens and diluted bitumens From Western Canadian Oil Sands: Chapters 1-9.
International Spill Control Organization (ISCO) Newsletters 457-465. http://www.spillcontrol.org/
(accessed May, 2015)
Fingas, M.F. 2014. Water-in-oil emulsions: Formation and prediction. Journal of Petroleum Science Research 3(1):
38-49.
Fingas, M.F. 2015a. Appendix C. Ice nomenclature. In: Fingas, M.F. (ed.) Handbook of Oil Spill Science and
Technology. Wiley & Sons, Inc., Hoboken, NJ.
Fingas, M.F. 2015b. Chapter 1. Introduction to oil chemistry and properties. In: Fingas, M.F. (ed.) Handbook of Oil
Spill Science and Technology. Wiley & Sons, Inc., Hoboken, NJ.
Fingas, M.F. 2015c. Chapter 7: Oil and petroleum evaporation. In: Fingas, M.F. (ed.) Handbook of Oil Spill Science
and Technology. Wiley & Sons, Inc., Hoboken, NJ.
Fingas, M.F. 2015d. Diluted bitumen (Dilbit): A future high risk spilled material. Interspill 2015 Conference
Proceedings, Amsterdam, March 24-26, 2015.
Fingas, M.F. 2015e. Review of the properties and behaviour of diluted bitumens. In: Proceedings of the 38th Arctic
and Marine Oilspill Program (AMOP) Technical Seminar. Environment Canada, Ottawa. pp. 470-494.
Fingas, M.F. and B. Fieldhouse. 2015. Chapter 8: Water-in-oil emulsions: Formation and prediction. In: Fingas,
M.F.( ed.) Handbook of Oil Spill Science and Technology. Wiley & Sons, Inc., Hoboken, NJ.
FISG (Federal Interagency Solutions Group, Oil Budget Calculator Science and Engineering Team). 2010. Oil
Budget Calculator Deepwater Horizon. Technical Documentation. A report to the National Incident

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 418
Command. http://www.restorethegulf.gov/sites/default/files/documents/pdf/OilBudgetCalc_Full_HQ-
Print_111110.pdf. (accessed August 2015)
Fitzpatrick, F.A., Boufadel, M.C., Johnson, R., Lee, K., Graan, T.P., Bejarano, A.C., Zhu, Z., Waterman, D.,
Capone, D.M., Hayter, E., Hamilton, S.K., Dekker, T., Garcia, M.H. and J.S. Hassan. 2015. Oil-Particle
Interactions and Submergence from Crude Oil Spills in Marine and Freshwater Environments – Review of
the Science and Future Science Needs. U.S. Geological Survey Open File Report 2015-1076.
http://dx.doi.org/10.3133/ofr20151076.
Foght, J. 2008. Anaerobic biodegradation of aromatic hydrocarbons: Pathways and prospects. Journal of Molecular
Microbiology and Biotechnology 15: 93-120.
Fowler, S.W., Readman, J.W., Oregioni, B., Villeneuve, J.-P. and K. McKay. 1993. Petroleum hydrocarbons and
trace metals in nearshore Gulf sediments and biota before and after the 1991 war: An assessment of
temporal and spatial trends. Marine Pollution Bulletin 27: 171-182.
Fustic, M., Bennett, B., Huang, H. and S. Larter. 2012. Differential entrapment of charged oil – New insights on
McMurray Formation oil trapping mechanisms. Marine and Petroleum Geology 36: 50-69.
González-Macías, C., Schifter, I., Lluch-Cota, D.B., Méndez-Rodríguez, L., and S. Hernández-Vázquez. 2006.
Distribution, enrichment and accumulation of heavy metals in coastal sediments of Salina Cruz Bay,
Mexico. Environmental Monitoring and Assessment 118: 211-230.
Gosselin, P., Hrudey, S.E., Naeth, M.A., Plourde, A., Therrien, R., Van Der Kraak, G. and Z. Xu. 2010.
Environmental and Health Impacts of Canada’s Oil Sands Industry. Royal Society of Canada Expert Panel
Report. http://www.rsc.ca/en/expert-panels/rsc-reports. (Accessed May 2015).
Government of Canada. 2013. Properties, Composition and Marine Spill Behaviour, Fate and Transport of Two
Diluted Bitumen Products from the Canadian Oil Sands. Federal Government Technical Report En84-
96/2013E-PDF ISBN 978-1-100-230004-7. http://publications.gc.ca/site/eng/457066/publication.html.
Gray, M.R. 2015. Upgrading Oilsands Bitumen and Heavy Oil. University of Alberta Press, Edmonton, AB.
Greer, C.W., Fortin, N., Sanschagrin, S., Yergeau, E., Whyte, L.G., King, T.L. and K. Lee. 2015. Natural
degradation potential for crude oil at sub-zero temperatures in the Canadian Arctic marine environment.
Abstract, 38th Arctic and Marine Oilspill Program (AMOP) Technical Seminar. Environment Canada,
Ottawa.
Grewer, D.M., Young, R.F., Whittal, R.M. and P.M. Fedorak. 2010. Naphthenic acids and other acid-extractables in
water samples from Alberta: What is being measured? Science of the Total Environment 408: 5997–6010.
Gros, J., Nabi, D., Würz, B., Wick, L.Y., Brussaard, C.P.D., Huisman, J., van der Meer, J.R., Reddy, C.M. and J.S.
Arey. 2014. First day of an oil spill on the open sea: Early mass transfers of hydrocarbons to air and water.
Environmental Science & Technology 48: 9400-9411.
Harper, J.R., Laforest, S. and G. Sergey. 2015. Field investigations of intertidal sediment permeability related to
spilled oil retention in British Columbia shorelines. In: Proceedings of the 38th Arctic and Marine Oilspill
Program (AMOP) Technical Seminar. Environment Canada, Ottawa. Pp. 631-648.
Hazen, T.C., Dubinsky, E.A., DeSantis, T.Z., Andersen, G.L., Piceno, Y.M., Singh, N., Jansson, J.K., Probst, A.,
Borglin, S.E., Fortney, J.L., Stringfellow, W.T., Bill, M., Conrad, M.E., Tom, L.M., Chavarria, K.L., Alusi,
T.R., Lamendella, R., Joyner, D.C., Spier, C., Baelum, J., Auer, M., Zemla, M.L., Chakraborty, R.,
Sonnenthal, E.L., D’Haeseleer, P., Holman, H.Y.N., Osman, S., Lu, Z.M., Van Nostrand., J., Deng, Y.,
Zhou, J.Z. and O.U. Mason. 2010. Deep-sea oil plume enriches indigenous oil-degrading bacteria. Science
330: 204–208.
Head, I.M., Jones, D.M. and W.F.M. Röling. 2006. Marine microorganisms make a meal of oil. Nature Reviews
Microbiology 4: 1730.
Hess Corporation. 2012. Safety Data Sheet No. 6608. http://www.hess.com/docs/us-safety-data-sheets/crude-oil-
(sour).pdf?sfvrsn=2. (accessed June, 2015).
Hoff, A. and H.D. Dettman. 2012. Global asphaltene content and structure related to process behaviour. Paper
WHOC12-275. In: Proceedings, World Heavy Oil Congress. Aberdeen, Scotland.
Hollebone, B.P. 2015. Appendix A. The oil properties data appendix. In: M.F. Fingas, M.F. (ed.) Handbook of Oil
Spill Science and Technology. Wiley & Sons, Inc., Hoboken, NJ.
Hollebone, B.P., Fieldhouse, B., Sergey, G., Lamter, P., Wang, Z., Yang, C. and M. Landriaut. 2011. The behaviour

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 419
of heavy oil in fresh water lakes. In: Proceedings of the 34th Arctic and Marine Oilspill Program (AMOP)
Technical Seminar on Environmental Contamination and Response. Environment Canada, Ottawa. Pp.
668-709.
Indian and Northern Affairs Canada. 1995. Petroleum Exploration in Canada. Chapter 4 – Mackenzie Delta and
Beauford Sea. http://www.aadnc-aandc.gc.ca/eng/1321378288752/1321378450630 (accessed August,
2015)
Islam, A., Kim, D., Yim, U.H., Shim, W.J. and S. Kim. 2015. Structure-dependent degradation of polar compounds
in weathered oils observed by atmospheric pressure photo-ionization hydrogen/deuterium exchange
ultrahigh resolution mass spectrometry. Journal of Hazardous Materials 296: 93-100.
ITOPF (International Tanker Owners Pollution Federation Ltd.,). 2014. Fate of Oil Spills: Weathering.
http://www.itopf.com/knowledge-resources/documents-guides/fate-of-oil-spills/weathering/ (accessed July,
2015)
Jones, D., West, C.E., Scarlett, A.G., Frank, R.A. and S.J. Rowland. 2012. Isolation and estimation of the ‘aromatic’
naphthenic acid content of an oil sands process-affected water extract. Journal of Chromatography A 1247:
171-175.
Johnson, R.J., Smith, B.E., Sutton, P.A., McGenity, T.J., Rowland, S.J. and C. Whitby. 2011. Microbial
biodegradation of aromatic alkanoic naphthenic acids is affected by the degree of alkyl side chain
branching. ISME Journal 5: 486–496.
Jokuty, P., Whiticar, S., Fingas, M., Meyer, E. and C. Knobel. 1995. Hydrocarbon groups and their relationships to
oil properties and behaviour. In: Proceedings of the 18th Arctic and Marine Oilspill Program (AMOP)
Technical Seminar. Environment Canada, Ottawa. Pp. 1-19.
Joung, DJ. and A.M. Shiller. 2013. Trace element distribution in the water column near the Deepwater Horizon well
blowout. Environmental Science & Technology 47: 2161-2168.
Joye, S.B. 2015. Deepwater Horizon, 5 years on. Science 349: 592-593.
Kang, Z., Yeung, A., Foght, J.M. and M.R. Gray. 2008a. Hydrophobic bacteria at the hexadecane-water interface:
examination of micrometer-scale interfacial properties. Colloids and Surfaces B: Biointerfaces 67: 59-66.
Kang, Z., Yeung, A., Foght, J.M. and M.R. Gray. 2008b. Mechanical properties of hexadecane-water interfaces with
adsorbed hydrophobic bacteria. Colloids and Surfaces B: Biointerfaces 62: 273-279.
Katakey, R. and W. Zhu. 2015. Shell halts Alaska offshore exploration after failing to find enough oil. Bloomberg
Business, September 27, 2015. http://www.bloomberg.com/news/articles/2015-09-28/shell-to-stop-
exploring-offshore-alaska-on-regulations-costs (accessed October 2015)
Kaufman, M. 2015. US Follows Royal Dutch Shell plc. Backs away from Arctic drilling.
http://royaldutchshellplc.com/2015/10/19/us-follows-royal-dutch-shell-plc-backs-away-from-arctic-
drilling/ (accessed October 19, 2015)
King, T.L., Robinson, B., Boufadel, M. and K. Lee. 2014. Flume tank studies to elucidate the fate and behavior of
diluted bitumen spilled at sea. Marine Pollution Bulletin 83: 32-37.
King, T., Robinson, B., Ryan, S. and Y. Lu. 2015a. Fate of Chinese and Canadian oil treated with chemical
dispersants in a wave tank. In: Proceedings of the 38th Arctic and Marine Oilspill Program (AMOP)
Technical Seminar. Environment Canada, Ottawa. pp. 798-811.
King, T.L., Robinson, B., McIntyre, C., Toole, P., Ryan, S., Saleh, F., Boufadel, M.C. and K. Lee. 2015b. Fate of
surface spills of Cold Lake blend diluted bitumen treated with dispersant and mineral fines in a wave tank.
Environmental Engineering Science 32: 250-261.
Kleindienst, S., M. Seidel, K. Ziervogel, S. Grim, K. Loftis, S. Harrison, S.Y. Malkin, M.J. Perkins, J. Field, M.L.
Sogin, T. Dittmar, U. Passow, P.M. Medeiros and S.B. Joye. 2015. Chemical dispersants can suppress the
activity of natural oil-degrading microorganisms. Proceedings National Academy of Sciences (USA) 112:
14900-14905. (doi:10.1073/pnas.1507380112).
Kokal, S. 2002. Crude oil emulsions: A state-of-the-art review. SPE 77497. Society of Petroleum Engineers.
www.onepetro.org/journal-paper/SPE-77497-PA
Kropp, K.G., Saftic, S., Andersson, J.T. and P.M. Fedorak. 1996. Transformations of six isomers of
dimethylbenzothiophene by three Pseudomonas strains. Biodegradation 7: 203-221.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 420
Laforest, S., Lambert, P. and M. Goldthorp. 2015. The Shoreline Studies Program: An update on shoreline surveys
and initial studies on the penetration and retention of diluted bitumen into marine shoreline substrates. In:
Proceedings of the 38th Arctic and Marine Oilspill Program (AMOP) Technical Seminar. Environment
Canada, Ottawa. Pp. 649-663.
Lea-Smith, D.J., Biller, S.J., Davey, M.P., Cotton, C.A.R., Perez Sepulveda, B.M., Turchyn, A.B., Scanlan, D.J.,
Smith, A.G., Chisholm, S.W. and C.J. Howe. 2015. Contribution of cyanobacterial alkane production to the
ocean hydrocarbon cycle. Proceedings National Academy of Sciences (USA) in press. doi:
10.1073/pnas.1507274112
Lee, K. 2002. Oil–particle interactions in aquatic environments: Influence on the transport, fate, effect and
remediation of oil spills. Spill Science & Technology Bulletin 8(1): 3-8.
Lee, K. and E.M. Levy. 1989. The enhanced biodegradation of condensate and crude oil on beaches of Atlantic
Canada. In: Proceedings 1989 Oil Spill Conference, February 13 - 16, 1989, San Antonio, Texas. American
Petroleum Institute, Environmental Protection Agency, and United States Coast Guard. pp. 479-486.
Lee, K., M. Boudreau, J. Bugden, L. Burridge, S.E. Cobanli, S. Courtenay, S. Grenon, B. Hollebone, P. Kepkay, Z.
Li, M. Lyons, H. Niu, T.L. King, S. MacDonald, E.C. McIntyre, B. Robinson, S.A. Ryan and G.
Wohlgeschaffen. 2011a. State of Knowledge Review of Fate and Effect of Oil in the Arctic Marine
Environment 2011. National Energy Board of Canada, Ottawa, ON. 267 pp. https://docs.neb-one.gc.ca/ll-
eng/llisapi.dll/fetch/2000/90463/621169/700096/704342/A2A8R2_-
_NEB_State_of_Knowledge_Review_of_Fate_and_Effect_of_Oil_DFO_COOGER.pdf?nodeid=704343&
vernum=-2
Lee, K., Li, Z., Robinson, B., Kepkay, P.E., Blouin, M. and B. Doyon. 2011b. Field trials of in-situ oil spill
countermeasures in ice-infested waters. In: Proceedings of the International Oil Spill Conference (IOSC).
http://dx.doi.org/10.7901/2169-3358-2011-1-160 (accessed July, 2015)
Lee, K., Bugden, J., Cobanli, S., King, T., McIntyre, C., Robinson, B., Ryan, S. and G. Wohlgeschaffen. 2012. UV-
epifluorescence microscopy analysis of sediments recovered from the Kalamazoo River. U.S. EPA
Kalamazoo Administrative Record, document #1277. http://www.epa.gov/enbridgespill/ar/enbridge-AR-
1277.pdf (accessed July, 2015)
Madison, B.N., Hodson, P.V. and V.S. Langlois. 2015. Diluted bitumen causes deformities and molecular responses
indicative of oxidative stress in Japanese medaka embryos. Aquatic Toxicology 165:222-230.
Martin, J.D., Adams, A., Hollebone, B.P., King, T., Brown, R.S. and P.V. Hodson. 2014. Chronic toxicity of heavy
fuel oils to fish embryos using multiple exposure scenarios. Environmental Toxicology and Chemistry
33:677-687.
Mbadinga, S.M., Wang, L.-Y., Zhou, L, Liu, J.-F., Gu, J.-D., and B.-Z. Mu. 2011. Microbial communities involved
in anaerobic degradation of alkanes. International Biodegradation and Biodeterioration 65: 1-13.
McFarlin, K.M., Prince, R.C., Perkins, R. and M.B. Leigh. 2014. Biodegradation of dispersed oil in Arctic seawater
at -1°C. PLoS ONE 9: e84297.
McGenity, T.J., Folwell, B.D., McKew, B.A. and G.O. Sanni. 2012. Marine crude-oil biodegradation: a central role
for interspecies interactions. Aquatic Biosystems 8: 10.
McGowan, F. 1990. The development of Orimulsion and Venezuelan oil strategy. Energy Policy 18(10): 913-926.
Melbye, A.G., Brakstad, O.G., Hokstad, J.N., Gregersen, I.K., Hansen, B.H., Booth, A.M., Rowland, S.J. and K.E.
Tollefsen. 2009. Chemical and toxicological characterization of an unresolved complex mixture-rich
biodegraded oil. Environmental Toxicology and Chemistry 28(9): 1815-1824.
Murphy, D.W., Li, C., d’Albignac, V., Morra, D. and J. Katz. 2015. Splash behavior and oily marine aerosol
production by raindrops impacting oil slicks. Journal of Fluid Mechanics. 780: 536-577.
Nmegbu, G.C.J. 2014. Analysis of effects of temperature on petrophysical properties of petroleum reservoir rocks
and fluids. International Journal of Engineering Research and Development 10(6): 53-61.
North Dakota Petroleum Council. No date. Bakken Crude Properties. http://www.ndoil.org/resources/bkn/.
(accessed June 2015)
NRC (National Research Council). 1999. Spills of Nonfloating Oils: Risk and Response. National Academy of
Science, National Research Council, Washington, D.C.
http://www.nap.edu/openbook.php?isbn=0309065909. (accessed June 2015).

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 421
NRC. 2003. Oil in the Sea III: Inputs, Fates and Effects. National Academies Press.
http://www.nap.edu/catalog/10388.html. (accessed June, 2015).
NRC. 2013. Effects of Diluted Bitumen on Crude Oil Transmission Pipelines. National Academy of Sciences,
Washington, D.C. http://www.nap.edu/read/18381/ (accessed Oct 2015)
NRC. 2014. Responding to Oil Spills in the U.S. Arctic Marine Environment. National Academies Press,
Washington, D.C. http://www.nap.edu/catalog/18625/responding-tooil-spills-in-the-us-arctic-marine-
environment. (Accessed June, 2015)
NTSB (National Transportation Safety Board). 2012. Enbridge Incorporated Hazardous Liquid Pipeline Rupture and
Release. Accident Report NTSB/PAR-12/01, PB2012-916501.
http://www.ntsb.gov/investigations/AccidentReports/Reports/PAR1201.pdf. (Accessed July 2015).
Omotoso, O.E., Munoz, V.A. and R.J. Mikula. 2002. Mechanisms of crude oil–mineral interactions. Spill Science &
Technology Bulletin 8(1): 45-54
Owens, E.H., Harper, J.R. Robson, W. and P.D. Boehm. 1987. Fate and persistence of crude oil stranded on a
sheltered beach. Arctic 40 (Supp. 1): 109-123.
Passow, U. 2014. Formation of rapidly-sinking, oil-associated marine snow. Deep-Sea Research II:
doi:10.1016/j.dsr2.2014.10.001 (in press)
Passow, U., Ziervogel, K., Asper, V. and A. Diercks. 2012. Marine snow formation in the aftermath of the
Deepwater Horizon oil spill in the Gulf of Mexico. Environmental Research Letters 7: 035301.
Payne, J.R., McNabb Jr., G.D. and J.R. Clayton Jr. 1990. Oil weathering behavior in Arctic environments. Polar
Research 10: 632-662.
Peng, P., Morales-Izquierdo, A., Hogg, A. and O.P. Strausz. 1997. Molecular structure of Athabasca asphaltene:
Sulfide, ether and ester linkages. Energy & Fuels 11: 1171-1187.
Peters, K.E., Walters, C.C. and J.M. Moldowan. 2005. The Biomarker Guide. Vol. 2. Biomarkers and Isotopes in
Petroleum Exploration and Earth History. 2nd Edition. Cambridge University Press.
Petrich, C., Karlsson, J. and H. Eicken. 2013. Porosity of growing sea ice and potential for oil entrainment. Cold
Regions Science and Technology 87: 27-32.
Petrova, L.M. Abbakumova, N.A., Foss, T.R. and G.V. Romanov. 2011. Structural features of asphaltene and
petroleum resin fractions. Petroleum Chemistry 51(4): 252-256.
Prince, R.C. 2015. Oil spill dispersants: Boon or bane? Environmental Science & Technology 49: 6376-6384.
Prince, R.C., Elmendorf, D.L., Lute, J.R., Hsu, C.S., Haith, C.E., Senius, J.D., Dechert, G.J.,Douglas, G.S. and E.L.
Butler. 1994. 17α(H),21β(H)-Hopane as a conserved internal marker for estimating the biodegradation of
crude oil. Environmental Science and Technology 28:142–145.
Prince, R.C. and C.C. Walters. 2007. Chapter 11. Biodegradation of oil hydrocarbons and its implications for source
identification. In: Wang, Z. and S.A. Stout (eds.) Oil Spill Environmental Forensics: Fingerprinting and
Source Identification. Academic Press, Burlington MA. pp. 349-380.
Romero, I.C., Schwing, P.T., Brooks, G.R., Larson, R.A., Hastings, D.W., Ellis, G., Goddard, E.A. and D.J.
Hollander. 2015. Hydrocarbons in deep-sea sediments following the 2010 Deepwater Horizon blowout in
the northeast Gulf of Mexico. PLOS ONE, 10(5): e0128371
Rowland, S.J., Scarlett, A.G., Jones, D., West, C.E. and R.A. Frank. 2011. Diamonds in the rough: Identification of
individual naphthenic acids in oil sands process water. Environmental Science & Technology 45: 3154-
3159.
Santos-Echeandía, J., Prego, R. and A. Cobelo-García. 2008. Influence of the heavy fuel spill from the Prestige
tanker wreckage in the overlying seawater column levels of copper, nickel and vanadium (NE Atlantic
Ocean). Journal of Marine Systems 72: 350-357.
Segato, R. no date. Quality Guidelines for Western Canadian Condensate. Canadian Association of Petroleum
Producers (CAPP). http://www.coqa-inc.org/docs/default-source/meeting-presentations/Segato0608.pdf.
(accessed June, 2015)
Sergy, G.A. and P.J. Blackall. 1987. Design and conclusions of the Baffin Island Oil Spill project. Arctic 40
(supplement): 1-9. http://arctic.journalhosting.ucalgary.ca/arctic/index.php/arctic/index
Sheremata, J.M., Gray, M.R., Dettman, H.D. and W.C. McCaffrey. 2004. Quantitative molecular representation and

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 422
sequential optimization of Athabasca asphaltenes. Energy & Fuels 18: 1377-1384.
Short, J.W. 2013. Susceptibility of Diluted Bitumen Products from the Alberta Tar Sands to Sinking in Water.
Report to the National Energy Board, # D72-80-2. 21 pp. http://docs.neb-one.gc.ca/ll-
eng/llisapi.dll?func=ll&objId=941166&objAction=browse (accessed July, 2015)
Sikkema, J., de Bont, J.A.M. and B. Poolman. 1995. Mechanisms of membrane toxicity of hydrocarbons.
Microbiology Reviews 59: 201-222.
Simecek-Beatty, D. and W.J. Lehr. 2007. Chapter 13. Trajectory modeling of marine oil spills. In: Wang, Z. and
S.A. Stout (eds.) Oil Spill Environmental Forensics: Fingerprinting and Source Identification. Academic
Press, Burlington MA. pp. 405-418
Speight, J.G. 2004. Petroleum asphaltenes: Part 1, Asphaltenes, resins and the structure of petroleum. Oil & Gas
Science and Technology – Revue d’IFP 59(5): 467-477.
Speight, J.G. 2014. The Chemistry and Technology of Petroleum. 5th Edition. CRC Press, Hoboken, NJ.
Stoffyn-Egli, P. and K. Lee. 2002. Formation and characterization of oil-mineral aggregates. Spill Science &
Technology Bulletin 8(1): 31-44.
Strain, P. M. 1986. The persistence and mobility of a light crude oil in a sandy beach. Marine Environmental
Research 19: 49-76.
Strausz, O.P., Mohelsky, T.W. and E.M. Lown. 1992. The molecular structure of asphaltene: an unfolding story.
Fuel 71: 1355-1363.
Thibodeaux, L.J., Valsarag, K.T., John, V.T., Papadopoulos, K.D., Pratt, L.R. and N.S. Pesika. 2011. Marine oil
fate: Knowledge gaps, basic research, and development needs; a perspective based on the Deepwater
Horizon spill. Environmental Engineering Science 28(2): 87-93.
Tonina, D. and J.M. Buffington. 2007. Hyporheic exchange in gravel bed rivers with pool-riffle morphology:
Laboratory experiments and three-dimensional modeling. Water Resources Research 43(W01421).
Tonina, D. and J.M. Buffington. 2009. Hyporheic exchange in mountain rivers I: Mechanics and environmental
effects. Geography Compass 3(3): 1063-1086.
Tsaprailis, H. 2014. Properties of Dilbit and Conventional Crude Oils. Report to Alberta Innovates–Energy and
Environmental Solutions. Alberta Innovates–Technology Futures. http://www.ai-
ees.ca/media/10927/properties_of_dilbit_and_conventional_crude_oils_-_aitf_-_final_report_revised.pdf
(accessed June, 2015).
US-EPA (US Environmental Protection Agency). 1994. Chemical summary for cyclohexane.
http://www.epa.gov/chemfact/s_cycloh.txt (accessed June, 2015).
US-EPA. 2013a. Attachment 2: U.S. EPA Volume Estimate for Submerged Line 6B Oil in the Kalamazoo River.
http://www.epa.gov/enbridgespill/pdfs/20130625/enbridge_epareview_20130508_qsoreport_attachment2.p
df (accessed July 4, 2015).
US-EPA. 2013b. Fact Sheet August 2013. Dredging begins on Kalamazoo River, Enbridge oil spill, Marshall,
Michigan. http://epa.gov/enbridgespill/pdfs/enbridge_fs_201308.pdf (accessed August, 2015)
Valentine, D.L., Fisher, G.B., Bagby, S.C., Nelson, R.K., Reddy, C.M., Sylva, S.P. and M.A. Woo. 2014. Fallout
plume of submerged oil from Deepwater Horizon. Proceedings of the National Academy of Sciences (USA)
111(45): 15906-15911.
Van Hamme, J.D., Singh, A. and O.P. Ward. 2003. Recent advances in petroleum microbiology. Microbiology and
Molecular Biology Reviews 67: 503-549.
Venosa, A.D and E.L. Holder. 2007. Biodegradability of dispersed crude oil and two different temperatures. Marine
Pollution Bulletin 54(5): 545-553.
Warnock, A.M., Hagen, S.C. and D.L. Passeri. 2015. Marine tar residues: a review. Water, Air and Soil Pollution
226: 68. DOI 10.1007/s11270-015-2298-5. (in press)
Wang, Z., Fingas, M., Blenkinsopp, S., Sergy, G., Landriault, M., Sigouin, L., Foght, J., Semple, K., and D.W.S.
Westlake. 1998. Comparison of oil composition changes due to biodegradation and physical weathering in
different oils. Journal of Chromatography A 809: 89-107.
Waterman, D.M. and M.H. García. 2015. Laboratory Tests of Oil-Particle Interactions in a Freshwater Riverine
Environment with Cold Lake Blend Weathered Bitumen. Civil Engineering Studies; Hydraulic Engineering
Series No. 106. Ven Te Chow Hydrosystems Laboratory, University of Illinois, Urbana IL.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 423
Wei, A., Moldowan, J.M, Peters, K.E., Wang, Y. and W. Xiang. 2007. The abundance and distribution of
diamondoids in biodegraded oils from the San Joaquin Valley: Implications for biodegradation of
diamondoids in petroleum reservoirs. Organic Geochemistry 38: 1910-1926.
Welch, M.A. 2014. Northern energy hub? Pipeline to Churchill explored -- deepwater port ice-free longer. Winnipeg
Free Press, August 1, 2014. http://www.winnipegfreepress.com/local/northern-energy-hub-
239208921.html (accessed October 2015)
Whitby, C. 2010. Microbial naphthenic acid biodegradation. Advances in Applied Microbiology 70: 93-125.
Wikipedia. 2015. Western Canadian Select. http://en.wikipedia.org/wiki/Western_Canadian_Select. (accessed June
2015)
Winter, J. and R. Haddad. 2014. Ecological impacts of dilbit spills: Considerations for natural resource damage
assessment. 37th Arctic and Marine Oil Spill Program (AMOP) Technical Seminar. Environment Canada,
Ottawa. https://usresponserestoration.files.wordpress.com/2014/06/winter-and-haddad-dilbit-nrda-
submitted-4-11-14.pdf.
Wise Jr., J.P., Wise, J.T.F., Wise, C.F., Wise, S.S., Gianios Jr., C., Xie, H., Thomson, W.D., Perkins, C., Falank, C.
and J.P. Wise Sr. 2014. Concentrations of the genotoxic metals, chromium and nickel, in whales, tar balls,
oil slicks and released oil from the Gulf of Mexico in the immediate aftermath of the Deepwater Horizon
oil crisis: Is genotoxic metal exposure part of the Deepwater Horizon legacy? Environmental Science &
Technology 48: 2997-3006.
WSP. 2014. Risk Assessment for Marine Spills in Canadian Water: Phase 1, Oil Spills South of the 60th Parallel.
Report to Transport Canada. http://files.wspdigital.com/risk/oil/english/131-17593-00_ERA_Oil-Spill-
South_150116.pdf.
Yang, C., Wang, Z., Yang, Z., Hollebone, B., Brown, C.E., Landriault, M., and B. Fieldhouse. 2011. Chemical
fingerprints of Alberta oil sands and related petroleum products. Environmental Forensics 12: 173–188.
Yang, F., Tchoukov, P., Pensini, E., Dabros, T., Czarnecki, J., Masliyah, J. and Z. Xu. 2014. Asphaltene sub-
fractions responsible for stabilizing water-in-oil emulsions: Part 1: Interfacial behaviours. Energy & Fuels
28(11): 6897-6904.
Yang, F., Tchoukov, P., Dettman, H., Tecklebrham, R.B., Liu, L, Dabros, T., Czarnecki, J. and Z. Xu. 2015.
Asphaltene sub-fractions responsible for stabilizing water-in-oil emulsions: Part II: Molecular
representations and molecular dynamic situations. Energy & Fuels doi: 10.1021/acs.energyfuels.5b00657
(in press).
Yarranton, H.W., Motahhari, H., Schoeggl, F.F. and Z.J. Zhou. 2015. Evaporative weathering of diluted bitumen
films. 174557-PA SPE. Journal of Canadian Petroleum Technology 54: 233-244.
Yim, U.H., Kim, M., Ha, S.Y., Kim, S. and W.J. Shim. 2012. Oil spill environmental forensics: The Heibei Spirit oil
spill case. Environmental Science & Technology 46: 6431-6437.
Zhang, X., Wiseman, S., Yu, H., Liu, H., Giesy, J.P. and M. Hecker. 2011. Assessing the toxicity of naphthenic
acids using a microbial genome wide living cell reported array system. Environmental Science &
Technology 45(5): 1984-1991.
Zhao, L., Boufadel, M.C., Socolofsky, S.A., Adams, E., King, T. and K. Lee. 2014. Evolution of droplets in subsea
oil and gas blowouts: Development and validation of the numerical model VDROP. Journal of Marine
Pollution Bulletin 83: 58-69.
Zhou, J., Dettman, H. and M. Bundred. 2015. A comparative analysis of environmental behavior of diluted bitumen
and conventional crudes. In: Proceedings of the 38th Arctic and Marine Oilspill Program (AMOP)
Technical Seminar. Environment Canada, Ottawa. pp. 495-516.

Chapter 3

Abed, R.M.M., Al-Thukair, A. and D. De Beer. 2006. Bacterial diversity of a cyanobacterial mat degrading
petroleum compounds at elevated salinities and temperatures. FEMS Microbiology Ecology 57: 290–301.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 424
Abu Laban, N., Dao, A., Semple, K. and J. Foght. 2014. Biodegradation of C7 and C8 isoalkanes under
methanogenic conditions. Environmental Microbiology: (in press) doi:10.1111/1462-2920.12643.
Arrigo, K.R. 2005. Marine microorganisms and global nutrient cycles. Nature 437: 349-355.
Atlas, R.M. 1981. Microbial degradation of petroleum hydrocarbons: An environmental perspective.Microbiology
Reviews 45: 180-209.
Atlas, R. M. (ed.). 1984. Petroleum Microbiology. Macmillan Co., New York.
Atlas, R. and T. Hazen. 2011. Oil biodegradation and bioremediation: A tale of the two worst spills in U.S.
history.Environmental Science & Technology 45: 6709–6715.
Bartha, R. and I. Bossert. 1984. The treatment and disposal of petroleum wastes. In: Atlas, R.M. (ed.) Petroleum
Microbiology. Macmillan, New York, NY. pp. 553–578.
Beal, R. and W.B. Betts. 2000. Role of rhamnolipid biosurfactants in the uptake and mineralization of hexadecane in
Pseudomonas aeruginosa. Journal of Applied Microbiology 89: 158-168.
Belore, R. C. and I. Buist. 1988. Modelling of oil spills in snow. In: Proceedings of the 11th Arctic Marine Oil Spill
Program Technical Seminar, Vancouver, BC, 7-9 June. Environment Canada, Ottawa, ON. Pp. 9-29.
Bosma, T.N.P., Middeldorp, P.J.M., Schraa, G. and A.J.B. Zender. 1997. Mass transfer limitation of
biotransformation: quantifying bioavailability. Environmental Science and Technology 31: 248-252.
Bossert, I. and R. Bartha. 1984. The fate of petroleum in soil ecosystems. In: Atlas, R.M. (ed.) Petroleum
Microbiology. Macmillan Publishing Company, New York. Pp. 435-476.
Boufadel, M.C., Suidan, M.T., Rauch, C.H.,Ahn, C.H. and A.D. Venosa. 1999. Nutrient transport in beaches
subjected to freshwater input and tides. In: Proceedings of 1999 International Oil Spill Conference.
American Petroleum Institute, Washington, DC. Pp. 471-476.
Bragg, J.R. and E.H. Owens. 1995. Shoreline cleansing by interactions between oil and fine mineral particles. In:
Proceedings of 1995 International Oil Spill Conference. American Petroleum Institute, Washington, DC.
Pp. 216-227.
Button, D. K. 1985. Kinetics of nutrient-limited transport and microbial-growth. Microbiology Review 49 (3): 270-
297.
Caldwell, M.E., Garrett, R.M., Prince, R.C. and J.M. Suflita. 1998. Anaerobic biodegradation of long-chain n-
alkanes under sulfate-reducing conditions. Environmental Science and Technology 32: 2191-2195.
Campo, P., Venosa, A.D. and M.T. Suidan. 2013. Biodegradability of Corexit 9500 and Dispersed South Louisiana
Crude Oil at 5 and 25 °C. Environmental Science and Technology 47: 1960-1967.
Carls, M.G., Thomas, R.E., Lilly, M.R. and S.D. Rice. 2003. Mechanism for transport of oil-contaminated
groundwater into pink salmon redds. Marine Ecology Progress Series 248: 245–255.
Cerniglia, C.E. 1992. Biodegradation of polycyclic aromatic hydrocarbons. Biodegradation 3: 351-368.
Chronopoulou, P.M., Gbemisola, O.S., Silas-Olu, D.I., van der Meer, J.R., Timmis, K.M., Brussaard, C.P.D. and
T.J. McGenity. 2015. Generalist hydrocarbon-degrading bacterial communities in the oil-polluted water
column of the North Sea. Microbial Biotechnology 8(3): 434-447.
Coates, J.D., Woodward, J., Allen, J., Philip, P. and D.R. Lovley. 1997. Anaerobic degradation of polycyclic
hydrocarbons and alkanes in petroleum-contaminated marine harbour sediments. Applied Environmental
Microbiology 63: 3589-3593.
Colwell, R.R., Mills, A.L., Walker, J.D. Garcia-Tolle, P. and P.V. Campos. 1978. Microbial ecology studies of the
Metula spill in the Straits of Magellan. Journal of the Fisheries Research Board of Canada 35: 573-580.
Cooney, J.J. 1984. The fate of petroleum pollutants in freshwater ecosystems. In: Atlas (ed) Petroleum
Microbiology. Macmillan Publishing Company, New York. Pp. 355-398.
Das, N. and P. Chandran. 2011. Microbial Degradation of Petroleum Hydrocarbon Contaminants: An Overview.
Biotechnology Research International 2011: 13 pp. http://dx.doi.org/10.4061/2011/941810.
Diaz, M.P., Boyd, K.G., Grigson, S.G.W. and J.G. Burgess. 2002. Biodegradation of crude oil across a wide range
of salinities by an extremely halotolerant bacterial consortium MPD-M, immobilized onto polypropylene
fibers. Biotechnology and Bioengineering 79(2):145–153.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 425
Dickins, D.F. 2011.Behaviour of oil spills in ice and implications for spill response.In Proceedings of the Arctic
Technology Conference, Houston TX, 7-9 February 2011, Offshore Technology Conference (OTC) Paper #
22126, pp.15.
Dutta, T. K., and S. Harayama. 2000. Fate of crude oil by the combination of photooxidation and biodegradation.
Environ. Sci. Technol. 34:1500–1505.
Faksness, L.-G. and P. J. Brandvik. 2008a. Distribution of water soluble components from Arctic marine oil spills –
A combined laboratory and field study. Cold Regions Science and Technology 54: 97-105.
Faksness, L.-G. and P. J. Brandvik. 2008b. Distribution of water soluble components from oil encapsulated in Arctic
sea ice: Summary of three field seasons. Cold Regions Science and Technology 54: 106-114.
Fingas, M. 2015. Diluted bitumen (dilbit): a future high risk spilled material. In: Proceedings of the 2015 Interspill
Conference, March 24-26, 2015, Amsterdam.
Floodgate, G. 1984. The fate of petroleum in marine ecosystems. In: Atlas, R.M. (ed.) Petroleum Microbiology.
Macmillan Publishing Company, New York. Pp. 355-398.
Foght, J. 2006. Potential for biodegradation of sub-littoral residual oil by naturally occurring microorganisms
following the Lake Wabamun train derailment. http://environment.gov.ab.ca/info/library/7746.pdf
Foght, J.M. 2008. Anaerobic biodegradation of aromatic hydrocarbons: Pathways and prospects. Journal of
Molecular Microbiology and Biotechnology 15: 93-120.
Foght, J.M., Fedorak, P.M., and D.W.S. Westlake. 1990. Mineralization of 14C-hexadecane and 14C-phenanthrene
in crude oil: specificity among bacterial isolates. Canadian Journal of Microbiology 35: 169-175.
Fuchs, G., Boll, M. and J. Heider. 2011. Microbial degradation of aromatic compounds - from one strategy to four.
Nature Reviews Microbiology 9: 803-816.
Garrett, R. M., Pickering, I.J., Haith, C.E. and R. C. Prince. 1998. Photooxidation of crude oils. Environmental
Science & Technology 32: 3719-3723.
Gittel A., Donhauser, J., Røy, H., Girguis, P.R., Jørgensen, B.B. and K.U. Kjeldsen. 2015. Ubiquitous presence and
novel diversity of anaerobic alkane degraders in cold marine sediments. Frontiers in Microbiology 6:1-15.
Golyshin, P.N, Chernikova, T.N., Abraham, W.R., Lunsdorf, H., Timmis, K.N. and M.M. Yakimov. 2012.
Oleiphilaceae fam. nov., to include Oleiphilus messinensis gen. nov., sp nov., a novel marine bacterium
that obligately utilizes hydrocarbons. International Journal of Systematic Evolutionary Microbiology 52:
901-911.
Goodman, R. 2009. Wabamun: A Major Inland Spill. CiteSeerX.
http://citeseerx.ist.psu.edu/viewdoc/summary?doi=10.1.1.541.9388. (accessed August 2015).
Government of Canada. 2013. Properties, Composition and Marine Spill Behaviour, Fate and Transport of Two
Diluted Bitumen Products from the Canadian Oil Sands. Federal Government Technical Report.
Environment Canada.
Greer, C.W., Fortin, N., Sanschagrin, S., Yergeau, E., Whyte, L.G., King, T.L. and K. Lee. 2015. Natural
degradation potential for crude oil at sub-zero temperatures in the Canadian Arctic marine environment.
Abstract, 38th Arctic and Marine Oilspill Program (AMOP) Technical Seminar. Environment Canada,
Ottawa.
Gros, J., D. Nabi, B. Wurz, L.Y. Wick, C.P.D. Brussaard, J. Huisman, J.R. van der Meer, C.M. Reddy, and J.S.
Arey. 2015. First day of an oil spill on the open sea: early mass transfers of hydrocarbons to air and water,
Environmental Science and Technology 48 (16): 9400-9411.
Gutierrez, T., Berry, D., Yang, T., Mishamandani, S., Mckay, L. Teske, A. and M.D. Aitken. 2013. Role of bacterial
exopolysaccharides (EPS) in the fate of the oil released during the Deepwater Horizon Oil Spill. PLoS ONE
8(6): e67717. DOI: 10.1371/journal.pone.0067717.
Haas, C. and S.E.L. Howell. 2015. Ice thickness in the Northwest Passage. Geophysical Research Letters 42 (18):
7673-7680.
Harms, H. and T.N.P. Bosma. 1997. Mass transfer limitation of microbial growth and pollutant degradation. Journal
of Industrial Microbiology 18: 97-105.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 426
Hazen, T.C., Dubinsky, E.A., Desantis, T.Z., Andersen, G.L., Piceno, Y.M., Singh, N., Jansson, J.K., Probst, A.,
Borglin, S.E., Fortney, J.L., Stringfellow, W.T., Bill, M., Conrad, M.E., Tom, L.M., Chavarria KL, Alusi
TR, Lamendella R, Joyner DC, Spier C, Baelum J, Auer M, Zemla ML, Chakraborty, R., Sonnenthal, E.L.,
D'haeseleer, P., Holman, H.Y., Osman, S., Lu, Z., Van Nostrand, J.D., Deng, Y., Zhou, J. and O.U. Mason.
2010. Deep-sea oil plume enriches indigenous oil- degrading bacteria. Science 330: 204-208.
Head, I.M. and R.P.J. Swannell. 1999. Bioremediation of petroleum hydrocarbon contaminants in marine habitats.
Current Opinion in Biotechnology 10: 234-239.
Head, I.M., D.M. Jones, and W.F.M. Roling. 2006. Marine microorganisms make a meal of oil. Nature Reviews
Microbiology 4: 173-182.
IPIECA (International Petroleum Industry Environmental Conservation Association). 1995. Biological Impacts of
Oil Pollution: Rocky shores. IPIECA Report Series 7:22.
Itoh, S. and T. Suzuki. 1972. Effect of rhamnolipids on growth of a Pseudomonas aeruginosa mutant deficient in n-
paraffin-utilizing ability. Agricultural Biology and Chemistry 36: 2233-2235.
Ivanov, B., Bezgreshnov, A., Kubyshkin, N. and A. Kursheva. 2005. Spreading of oil products in sea ice and their
influence on the radiation properties of the snow-ice cover. In: Proceedings of the 18th International Port
and Oceans Engineering Under Arctic Conditions. pp 853-862.
Johnsen, A. R., Wick, L.Y. and H. Harms. 2005. Principles of microbial PAH-degradation in soil Environmental
Pollution 133: 71-84.
Jordan, R.E. and J.R. Payne. 1980. Fate and Weathering of Petroleum Spills in the Marine Environment. Ann Arbor
Science Publishers, Inc., Ann Arbor, MI.
Jones, T.J., Neustadter, E.L. and K.P. Whittingham. 1978. Water-in-crude oil emulsion stability and emulsion
destabilization by chemical demulsifiers. Journal of Canadian Petroleum Technology 17: 2-8.
Kanaly, R.A. and S. Harayama. 2010. Advances in the field of high-molecular-weight polycyclic aromatic
hydrocarbon biodegradation by bacteria. Microbial Biotechnology 3: 136-164.
Karl, T.R., Melillo, J.M.and T.C. Peterson. (eds.). 2009. Global Climate Change Impacts in the United States. A
report of the U.S. Global Change Research Program (USGCRP). Cambridge University Press, Cambridge,
UK.
Karrick, N.L. 1977. Alteration in petroleum resulting from physical-chemical and microbiological factors. In:
Malins, D.C. (Ed) Effects of Petroleum on Arctic and Subarctic Environments and Organisms Vol. 1.
Nature and Fate of Petroleum. Academic Press, Inc., New York. Pp. 225-299.
Kostka, J.E., Teske, A.P., Joye, S.B. and I.M. Head. 2014. The metabolic pathways and environmental controls of
hydrocarbon biodegradation in marine ecosystems. Frontiers in Microbiology: doi:
10.3389/fmicb.2014.00471.
Leahy, J.G. and R.R. Colwell. 1990. Microbial degradation of hydrocarbons in the environment. Microbiological
Reviews 54: 305-315.
Lea-Smith, D.J., S.J. Biller, M.P. Davey, C.A.R. Cotton, B.M. Perez Sepulveda, A.V. Turchyn, D.J. Scanlan, A.G.
Smith, S.M. Chisholm and C.J. Howe. 2015. Proceedings National Academy of Science: doi:
10.1073/pnas.1507274112.
Lee, K., Tremblay, G.H., and Gauthier, J., Cobanli, S.E. and M. Griffin. 1997a Bioaugmentation and biostimulation:
a paradox between laboratory and field results. In: Proceedings of 1997 International Oil Spill Conference.
American Petroleum Institute, Washington DC. Pp. 697-705.
Lee, K., Weise, A.M. and S. St. Pierre. 1997b. Enhanced oil biodegradation with mineral fine interaction. Spill
Science & Technology Bulletin 3(4): 263-267.
Lee, K., Li, Z., Robinson, B., Kepkay, P., Blouin, M. and B. Doyon. 2011. Field trials of in-situ oil spill
countermeasures in ice-infested waters.In:Proceedings of the 2011 International Oil Spill Conference,
Portland, Oregon, USA.May 23 – 26, 2011.pp. abs160.
Liao, Y., Geng, A. and H. Huang. 2009. The influence of biodegradation on resins and asphaltenes in the Liaohe
Basin. Organic Geochemistry 40: 312-320.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 427
Margesin, R. and F. Schinner. 2001. Biodegradation and bioremediation of hydrocarbons in extreme environments.
Applied Microbiology Biotechnology 56(5-6): 650-63.
Margesin R, Moertelmaier, C. and J. Mair. 2013. Low-temperature biodegradation of petroleum hydrocarbons (n-
alkanes, phenol, anthracene, pyrene) by four actinobacterial strains. International Biodeterioration and
Biodegradation 84:185–191.
Martins, L.M. and R.S. Peixoto. 2012. Biodegradation of petroleum hydrocarbons in hypersaline environments.
Brazilian Journal of Microbiology 43(3): 865-872.
Mbadinga, M., Wang, L-Y., Zhou, L., Liu, J-F., Gu, J-D. and B-Z. Mu. 2011. Microbial communities involved in
anaerobic degradation of alkanes. International Biodeterioration & Biodegradation 65: 1-13.
McFarlin, K.M., Prince, R.C., Perkins, R. and M.B. Leigh. 2014. Biodegradation of dispersed oil in arctic seawater
at -1°C.PLoS One9(1): e84297.doi: 10.1371/journal.pone.0084297.
McGenity, T.J., Folwell, B.D., McKew, B.A. and G.O. Sanni. 2012. Marine crude-oil biodegradation: a central role
for interspecies interactions. Aquatic Biosystems 8: 10.
Minai-Tehrani, D., Minoui, S. and A. Herfatmanesh. 2009. Effect of salinity on biodegradation of polycyclic
aromatic hydrocarbons (PAHs) of heavy crude oil in soil. Bulletin of Environmental Contamination and
Toxicology 82(2): 179-84.
Mitsch, W.J. and J.G. Gosselink. 1993. Wetlands. Van Nostrand Reinhold, New York.
Muskoka Watershed Council. 2007. Muskoka Watersheds Report Card.
http://www.muskokawatershed.org/programs/report-card/
Naveenkumar, S., Manoharan, N., Ganesan, S. Manivannan, S.P. and G. Velsamy. 2010. Isolation, screening and in
vitro mutational assessment of indigenous soil bacteria for enhanced capability in petroleum degradation.
International Journal of Environmental Science 1(4): 498.
Noordman, W. H. and D.B. Janssen. 2002. Rhamnolipid stimulates uptake of hydrophobic compounds by
Pseudomonas aeruginosa. Journal of Applied Microbiology 68: 4502-4508.
NRC (National Research Council). 1985. Oil in the Sea: Inputs, Fates and Effects. National Academies Press,
Washington DC.
NRC. 2003. National Research Council: Oil in the Sea III: Inputs, Fates and Effects. National Academies Press,
Washington, D.C.
NRC. 2014. Responding to Oil Spills in the U.S. Arctic Marine Environment. National Academies Press,
Washington, D.C.
Oberbremer, A. and R. Muller-Hurtig. 1989. Aerobic stepwise hydrocarbon degradation and formation of
biosurfactants by an original soil population in a stirred reactor. Applied Microbiology and Biotechnology
31: 582-586.
Pérez-Pantoja, D., González, B, and D.H. Pieper. 2010. Aerobic degradation of aromatic hydrocarbons. In: Timmis,
K.N., McGenity, T.J., Meer, J.R. and V. Lorenzo (eds.) Handbook of Hydrocarbon and Lipid
Microbiology. Springer, Berlin Heidelberg. pp. 799-837.
Peters, K.E., C.C. Walters, and J.M. Moldowan. 2005. The biomarker guide. II. Biomarkers and isotopes in
petroleum systems and earth history. Cambridge University Press, UK.
Pineda-Flores, G., Bollarguello, G. and A. Mestahoward. 2004. A microbial mixed culture isolated from a crude oil
sample that uses asphaltenes as a carbon and energy source. Journal of Biodegradation 15: 145–151.
Polak, J. and B. C-Y. Lu. 1973. Mutual solubility of hydrocarbons and water at 0 and 25°C. Canadian Journal of
Chemistry 51(12): 4018-4023.
Prince, R. C. 1993. Petroleum spill bioremediation in marine environments. Critical Reviews in Microbiology 19:
217-242.
Prince, R.C. 2010. Eukaryotic hydrocarbon degraders. In: Timmis, K.N., McGenity, T.J., van der Meer, J.R. and V.
Lorenzo (eds.) Handbook of Hydrocarbon and Lipid Microbiology. Springer, Berlin. Pp. 2065-2078.
Prince, R. C. and J. R. Clark. 2004. Bioremediation of marine oil spills. Studies in Surface Science and
Catalysis151: 495-512.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 428
Pritchard, P.H., Mueller, J.G., Rogers, J.C., Kremer, F.V. and J.A. Glaser. 1992. Oil spill bioremediation:
experiences, lessons and results from the Exxon Valdex oil spill in Alaska. Biodegradation 3: 315-335.
Prosser, J. I., Bohannan, B.J. M., Curtis, T.P., Ellis, R.J., Firestone, M.K. and R.P. Freckleton. 2007. The role of
ecological theory in microbial ecology. Nature Reviews Microbiology 5: 384–392.
Redfield, A. C. 1934. In: Daniel, R. J. (ed.) James Johnstone Memorial Volume. Liverpool University Press,
Liverpool. pp. 176–192.
Rojo, F. 2009. Degradation of alkanes by bacteria. Environmental Microbiology 11: 2477-2490.
Sabirova, J.S., Ferrer, M., Regenhardt, D., Timmis, K.N. and P.N. Golyshin. 2006. Proteomic insights into
metabolic adaptations in Alcanivorax borkumensis induced by alkane utilization. Journal of Bacteriology
188: 3763-3773.
Sabirova, J.S., Chernikova, T.N., Timmis, K.N. and P.N. Golyshin. 2008. Niche specificity factors of a marine oil-
degrading bacterium Alkanivorax borkumensis SK2. FEMS Microbiological Letters 285: 89-96.
Schneiker, S.V.A., dos Santos, P.M., Bartels, D., Bekel, T., Brecht, M., Buhrmester, J., Chernikova, T.N., Denaro,
R., Ferrer, M., Gertler, C., Goesmann, A., Golyshina, O.V., Kaminski, F., Khachane, A.N., Lang, S., Linke,
B., McHardy, A.C., Meyer, F., Nechitaylo, T., Pühler, A., Regenhardt, D., Rupp, O., Sabirova, J.S.,
Selbitschka, W., Yakimov, M.M., Timmis, K.N., Vorhölter, F.J., Weidner, S., Kaiser, O. and P.N.
Golyshin. 2006. Genome sequence of the ubiquitous hydrocarbon-degrading marine bacterium Alcanivorax
borkumensis. Nature Biotechnology 24: 997-1004.
Semanov, G., Volkov, V., Somkin, V. and D. Iljushenko-Krylov. 1997. Working Paper 76, Coastal Pollution
Emergency Plan. Part I. International Sea Route Programme (INSROP).
Serova, I. 1992. Behaviour of oil in ice and water at low temperatures. In: Combatting Marine Oil Spills in Ice and
Cold Climates, HELLCOM Seminar, Helsinki, Finland.
Singsaas, I., Brandvik, P.J., Daling, P.S., Reed, M. and A. Lewis. 1994. Fate and behaviour of oil spilled in the
presence of ice - A comparison of the results from recent laboratory, meso-scale flume and field tests. In:
Proceedings of the 17th Arctic and Marine Oilspill Program (AMOP) Technical Seminar. Environment
Canada, Ottawa, Ontario. pp. 355-370.
Short, J.W. and P.M. Harris. 1997. Pristane Monitoring in Mussels and Predators of Juvenile Pink Salmon and
Herring, Exxon Valdez Oil Spill Restoration Project Annual Report,
http://www.evostc.state.ak.us/Store/AnnualReports/1996-96195-Annual.pdf
Srinivasan, R., Lu, Q., Sorial, G.A. Venosa, A.D. and J. Mullin. 2007. Dispersant effectiveness of heavy fuel oils
using the Baffled Flask Test. Environmental Engineering Science 24 (9): 1307-1320.
Strassner, J.E. 1968. Effect of pH on Interfacial Films and Stability of Crude Oil-Water Emulsions. Journal of
Petroleum Technology 20 (3): 303-312.
Teramoto, M., Ohuchi, M., Hatmanti, A., Darmayati, Y., Widyastuti, Y., Harayama, S. and Y. Fukunaga. 2011.
Oleibacter marinus gen. nov., sp. nov., a bacterium that degrades petroleum aliphatic hydrocarbons in a
tropical marine environment. International Journal of Systematic and Evolutionary Microbiology 61: 375-
380.
Tissot, B.P. and D.H. Welte. 1984. Petroleum Formation and Occurrence. SpringerVerlag, Berlin, New York.
Ulrich, A.C., Guigard, S.E., Foght, J.M., Semple, K.M., Pooley K., Armstrong, J.E. and K.W. Biggar. 2009. Effect
of salt on aerobic biodegradation of petroleum hydrocarbons in contaminated groundwater. Biodegradation
20(1): 27-38.
Uraizee, F.A., Venosa, A.D. and M.T. Suidan. 1998. A model for diffusion controlled bioavailability of crude oil
components. Biodegradation 8: 287-296.
U.S. EPA. 2013. Climate Impacts on Alaska. http://www.epa.gov/climatechange/impacts-adaptation/alaska.html.
Valentine, D.L., Burch Fisher, G., Bagby, S.C., Nelson, R.K., Reddy, C.M., Sylva, S.P. and M.A. Wood. 2014.
Fallout plume of submerged oil from Deepwater Horizon. Proceedings of the National Academy of
Sciences 111 (145): 15906-15911.
Venkatesh, S., El-Tahan, H., Comfort, G. and R. Abdelnour. 1990. Modeling the behavior of oil spills in Ice-
infested waters. Atmosphere-Ocean 28: 303-329.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 429
Vinas, M, Grifoli, M., Sabate, J. and A.M. Solanas. 2002. Biodegradation of a crude oil by three microbial consortia
of different origins and metabolic capabilities. Journal of Industrial Microbiology & Biotechnology 28:
252-260.
Volkering, F., Breure, A. M., Sterkenburg, A and J.G. van Andel. 1992. Microbial degradation of polycyclic
aromatic hydrocarbons: Effect of substrate availability on bacterial growth kinetics. Applied Microbiology
and Biotechnology 36: 548-522.
Volkering, F., Breure, A.M. and W.H. Rulkens. 1998. Microbiological aspects of surfactant use for biological soil
remediation. Biodegradation 8: 401-417.
Wang, Z., Fingas, M., Blenkinsopp, S., Sergy, G., Landriault, M., Sigouin, L., Foght, J. and D.W.S. Westlake. 1998.
Comparison of oil composition changes due to biodegradation and physical weathering in different oils.
Journal of Chromatography A 809: 89-107.
Wang, Z., Yang, C., Hollebone, B., Brown, C.E., Landriault, M., Yang, Z. and J. Sun. 2011. Characterization of
biodegraded Wabamun spill oil samples incubated with lake fresh sediment or enrichment culture. In:
Proceedings of the 2011 International Oil Spill Conference. American Petroleum Institute, Washington
DC. pp. abs66.http://ioscproceedings.org/doi/pdf/10.7901/2169-3358-2011-1-66
Ward, D.M. and T.D. Brock. 1978. Hydrocarbon biodegradation in hypersaline environments. Applied
Environmental Microbiology 35: 353-359.
Wenger, L.M., C.L. Davis, and G.H. Isaksen. 2002. Multiple controls on petroleum biodegradation and impact on
oil quality. SPE Reservoir Evaluation and Engineering 5: 375-383.
Westlake, D.W.S., Jobson, A., Phillipee, R., and F.D. Cook. 1974. Biodegradability and crude oil composition.
Canadian Journal of Microbiology 20: 915-928.
Whitehouse, B.G. 1984. The effects of temperature and salinity on the aqueous solubility of polynuclear aromatic
hydrocarbons. Marine Chemistry 14: 319–332.
Wrenn, B.A., Boufadel, M.C., Suidan, M.T. and A.D. Venosa. 1997a. Nutrient transport during bioremediation of
crude oil contaminated beaches.In: In-Situ and On-Site Bioremediation: Volume 4. Battelle Press,
Columbus, OH. pp. 267-272.
Wrenn, B.A., Suidan, M.T., Strohmeier, K.L., Eberhart, B.L., Wilson, G.J. and A.D. Venosa.1997b. Nutrient
transport during bioremediation of contaminated beaches: Evaluation with lithium as a conservative
tracer.Water Research 31: 515-524.
WSP Canada, Inc. 2014. Risk assessment for marine spills in Canadian waters. Phase 1. Oil spills south of the 60th
parallel. Report to Transport Canada. 172 pp. and appendices. http://wcel.org/sites/default/files/file-
downloads/131-17593-00_ERA_Oil-Spill-South_150116_pp1-124.pdf
Yakimov, M.M., Timmis, K.N. and P.N. Golyshin. 2007. Obligate oil-degrading marine bacteria. Current Opinion
in Biotechnology 18 (3): 257-266.
Yang, S., Wen, X., Zhao, L., Shi, Y, and H. Jin. 2014. Crude oil treatement leads to shift of bacterial communities in
soils from the deep active layer and upper permafrost along the China-Russion crude oil pipeline route.
PLoS One 9(5) e96552.
Zhu, X., Venosa, A.D., Suidan M.T. and K. Lee. 2001. Guidelines for the Bioremediation of Marine Shorelines and
Freshwater Wetlands. U.S. Environmental Protection Agency.
Zhu, X., Venosa, A.D., Suidan, M.T. and K. Lee. 2004. Guidelines for the Bioremediation of Oil-Contaminated Salt
Marshes. EPA/600/R-04-074. U.S. Environmental Protection Agency.
Zobell, C. E. 1946. Action of microorganisms on hydrocarbons. Bacteriological Reviews 10:1–49.
Zobell, C.E. 1973. Microbial degradation of oil: Present state, problems, and perspectives. In: Ahearn, D.G. and S.P.
Meyers (eds.) The Microbial Degradation of Oil Pollutants. Publication No. LSU-SG-73-01. Louisiana
State University, Baton Rouge, LA. Pp. 3-16.

Chapter 4

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 430
Ackleh, A.S., Ioup, G.E., Ioup, J.W., Ma, B.L., Newcomb, J.J., Pal, N., Sidorovskaia, N.A. and C. Tiemann. 2012.
Assessing the Deepwater Horizon oil spill impact on marine mammal population through acoustics:
Endangered sperm whales. Journal of the Acoustical Society of America 131: 2306-2314.
Adams, G.G., Klerks, P.L., Belanger, S.E. and D. Dantin. 1999. The effect of the oil dispersant Omni-Clean® on the
toxicity of fuel oil no. 2 in two bioassays with the sheepshead minnow Cyprinodon variegatus.
Chemosphere 39: 2141-2157.
Adams, J., Sweezey, M. and P.V. Hodson. 2014a. Oil and oil dispersant do not cause synergistic toxicity to fish
embryos. Environmental Toxicology and Chemistry 33: 825-835.
Adams, J.E., Munno, K., Bornstein, J., King, T., Brown, R.S., Hollebone, B.P. and P.V. Hodson. 2014b.
Identification of compounds in heavy fuel oil that are chronically toxic to rainbow trout embryos through
effects-driven chemical fractionation. Environmental Toxicology and Chemistry 33: 825-835.
ADNR. 2014. Chapter Four: Habitat, Fish, and Wildlife. Alaska Peninsula Areawide Oil and Gas Lease Sales
(corrected 12/01/14). Written finding of the director. November 26, 2014. Alaska Department of Natural
Resources, Anchorage, AK.
Aeppli, C., Carmichael, C.A., Nelson, R.K., Lemkau, K.L., Graham, W.M., Redmond, M.C., Valentine, D.L. and
C.M. Reddy. 2012. Oil weathering after the Deepwater Horizon disaster led to the formation of oxygenated
residues. Environmental Science & Technology 46(16): 8799-8807.
Alloy, M.M., Boube, I., Griffitt, R.J., Oris, J.T. and A.P. Roberts. 2015. Photo-induced toxicity of Deepwater
Horizon slick oil to blue crab (Callinectes sapidus) larvae. Environmental Toxicology and Chemistry 34:
2061–2066.
Anderson, J.A., Kuhl, A.J. and A.N. Anderson. 2014. Toxicity of oil and dispersed oil on juvenile mud crabs,
Rhithropanopeus harrisii. Bulletin of Environmental Contamination and Toxicology 92: 375-380.
AOSRT (Arctic Oil Spill Response Technology – Joint Industry Programme). 2014. Environmental Impacts of
Arctic Oil Spills and Arctic Oil Spill Response Technologies. Literature Review and Recommendations.
Arctic Oil Spill Response Joint Industry Programme. http://www.arcticresponsetechnology.org/research-
projects.
Arias, C.R., Koenders, K. and A.M. Larsen. 2013. Predominant bacteria associated with red snapper from the
northern Gulf of Mexico. Journal of Aquatic Animal Health 25: 281-289.
Babcock, M. M., Irvine, G. V., Harris, P. M., Cusick, J. A. and S.D. Rice. 1996. Persistence of oiling in mussel beds
three and four years after the Exxon Valdez oil spill. In: Proceedings of the “Exxon Valdez Oil Spill
Symposium” held at Anchorage, Alaska, February 2-5, 1993. American Fisheries Society Symposium.
1996. American Fisheries Society, Bethesda MD.
Balk, L., Hylland, K., Hansson, T., Berntssen, M.H.G., Beyer, J., Jonsson, G., Melbye, A., Grung, M., Torstensen,
B.E, Borseth, J.F., Skarphedinsdottir, H. and J. Klungsoyr. 2011. Biomarkers in natural fish populations
indicate adverse biological effects of offshore oil production. PLoS ONE 6: 1-10.
Ballachey, B.E., Bodkin, J.L., Esler, D. and S.D. Rice. 2014. Lessons from the 1989 Exxon Valdez oil spill: A
biological perspective. Chap. 9. In: Alford, J.B., Peterson, M.S. and C.C. Green (eds) Impacts of Oil Spill
Disasters on Marine Habitats and Fisheries in North America.. CRC Press. pp. 181-197.
Barron, M., Carls, M., Short, J. and S. Rice. 2003. Photoenhanced toxicity of aqueous phase and chemically
dispersed weathered Alaska North Slope crude oil to Pacific herring eggs and larvae. Environmental
Toxicology and Chemistry 22: 650-660.
Barron, M.G, and L. Ka’aihue. 2001. Potential for photoenhanced toxicity of spilled oil in Prince William Sound
and Gulf of Alaska waters. Marine Pollution Bulletin 43: 86-92.
Barron, M.G., Carls, M.G., Heintz, R. and S.D. Rice. 2004. Evaluation of fish early life-stage toxicity models of
chronic embryonic exposures to complex polycyclic aromatic hydrocarbon mixtures. Toxicological Science
78: 60-67.
Barron, M.G., Carls, M.G., Short, J.W., Rice, S.D., Heintz, R.A., Rau, M. and R.D. Giulio. 2005. Assessment of the
phototoxicity of weathered Alaska North Slope crude oil to juvenile pink salmon. Chemosphere 60: 105-
110.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 431
Barron, M.G., Hemmer, M.J. and C.R. Jackson. 2013. Development of aquatic toxicity benchmarks for oil products
using species sensitivity distributions. Integrated Environmental Assessment and Management 9: 610-615.
Basu, N., Billiard, S., Fragoso, N., Omoike, A., Tabash, S., Brown, S. and P. Hodson. 2001 Ethoxyresorufin-o-de-
ethylase induction in trout exposed to mixtures of polycyclic aromatic hydrocarbons. Environmental
Toxicology and Chemistry 20: 1244-1251.
Bauder, M.B., Palace, V.P. and P.V. Hodson. 2005. Is oxidative stress the mechanism of toxicity in retene-exposed
trout larvae? Environmental Toxicology and Chemistry 24: 694-702.
Baumann, P.C., Smith, I.R. and C.D. Metcalfe. 1996. Linkages between chemical contaminants and tumors in
benthic Great Lakes fish. Journal of Great Lakes Research 22: 131-152.
Bell, B., Spotila, J.R. and J. Congdon. 2006. High incidence of deformity in aquatic turtles in the John Heinz
National Wildlife Refuge. Environmental Pollution 142: 457-465.
Beyer, J., Trannum, H.C., Bakke, T. and P.V. Hodson. 2015. Environmental effects of the Deepwater Horizon oil
spill: a review. Marine Pollution Bulletin (submitted).
Billiard, S.M., Querbach, K. and P.V. Hodson. 1999. Toxicity of retene to early life stages of two freshwater fish
species. Environmental Toxicology and Chemistry 18: 2070-2077.
Billiard, S.M., Hahn, M.E., Franks, D.G., Peterson, R.E., Bols, N.C. and P.V. Hodson. 2002. Binding of polycyclic
aromatic hydrocarbons (PAHs) to teleost arylhydrocarbon receptors (AHRs). Comparative Biochemistry
and Physiology B 133: 55-68.
Billiard, S.M., Meyer, J.N., Wassenberg, D.M., Hodson, P.V. and R.T. Di Giulio. 2008. Non-additive effects of
PAHs on early vertebrate development: mechanisms and implications for risk assessment. Toxicological
Sciences 105: 5-23.
Birtwell, I.K. 2008. Comments on the effects of oil spillage on fish and their habitat - Lake Wabamun, Alberta.
Report submitted to the Department of Fisheries and Oceans, Edmonton, AB.
Birtwell, I.K., Fink, R., Brand, D., Alexander, R. and C.D. McAllister. 1999. Survival of pink salmon
(Oncorhynchus gorbuscha) fry to adulthood following a 10-day exposure to the aromatic hydrocarbon
water-soluble fraction of crude oil and release to the Pacific Ocean. Canadian Journal of Fisheries and
Aquatic Sciences 56: 2087-2098.
Bolognesi, C., Perrone, E., Roggieri, P. and A. Sciutto. 2006. Bioindicators in monitoring long term genotoxic
impact of oil spill: Haven case study. Marine Environmental Research 62 (Supplement 1): S287-S291.
Bornstein, J.M., Adams, J.E., Hollebone, B., King, T., Hodson, P.V. and R.S. Brown. 2014. Effects-driven chemical
fractionation of heavy fuel oil to isolate compounds toxic to trout embryos. Environmental Toxicology and
Chemistry 33: 814-824.
Boudreau, M., Sweezey, M., Lee, K., Hodson, P.V. and S.C. Courtenay. 2009. Toxicity of Orimulsion-400 to early
life stages of Atlantic herring (Clupea harengus) and mummichog (Fundulus heteroclitus). Environmental
Toxicology and Chemistry 28: 1206-1217.
Brand, D.G., Fink, R., Bengeyfield, W., Birtwell, I.K. and C.D. McAllister. 2001. Salt water-acclimated pink
salmon fry (Oncorhynchus gorbuscha) develop stress-related visceral lesions after 10-day exposure to
sublethal concentrations of the water-soluble fraction of North Slope crude oil. Toxicologic Pathology 29:
574-584.
Brannon, E.L., Quinn, T.P., Whitman, R.P., Nevissi, A.E., Nakatani, R.E. and C.D. McAuliffe. 1986. Homing of
adult chinook salmon after brief exposure to whole and dispersed crude oil. Transactions of the American
Fisheries Society 115: 823-827.
Brannon, E.L., Maki, A.W., Moulton, L.L. and P.R. Parker. 2006. Results from a sixteen year study on the effects of
oiling from the Exxon Valdez on adult pink salmon returns. Marine Pollution Bulletin 52(8): 892-899.
Brannon, E.L., Collins, K., Cronin, M.A., Moulton, L.L., Maki, A.L. and K.R. Parker. 2012. Review of the Exxon
Valdez Oil spill effects on pink salmon in Prince William Sound, Alaska. Reviews in Fisheries Science
20(1): 20-60.
Brette, F., Machado, B., Cros, C., Incardona, J.P., Scholz, N.L. and B.A. Block. 2014. Crude oil impairs cardiac
excitation-contraction coupling in fish. Science 343: 772-776.
Brewton, R.A., Fulford, R. and R.J. Griffitt. 2013. Gene expression and growth as indicators of effects of the BP
Deepwater Horizon oil spill on spotted seatrout (Cynoscion nebulosus). Journal of Toxicology and
Environmental Health, Part A 76: 1198-1209.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 432
Brinkworth, L.C., Hodson, P.V., Tabash, S. and P. Lee. 2003. CYP1A induction and blue sac disease in early
developmental stages of rainbow trout (Oncorhynchus mykiss) exposed to retene. Journal of Toxicology
and Environmental Health - Part A 66: 47-66.
Brown-Peterson, N.J., Brewton, R.A., Griffitt, R.J. and R.S. Fulford. 2014. Impacts of the Deepwater Horizon oil
spill on the reproductive biology of spotted seatrout (Cynoscion nebulosus). Chap. 5. In: Alford, J.B.,
Peterson, M.S. and C.C. Green (eds) Impacts of Oil Spill Disasters on Marine Habitats and Fisheries in
North America. CRC Press. pp. 237-252.
Campagna, J.A., Miller, K.W. and S.A. Forman. 2003. Mechanisms of actions of inhaled anesthetics. New England
Journal of Medicine 348: 2110-2124.
Camus, L. and G.H. Olsen. 2008. Embryo aberrations in sea ice amphipod Gammarus wilkitzkii exposed to water
soluble fraction of oil. Marine Environmental Research 66: 221-222.
Carls, M.G., Rice, S. and J. Hose. 1999. Sensitivity of fish embryos to weathered crude oil: Part I. low-level
exposure during incubation causes malformations, genetic damage, and mortality in larval Pacific herring
(Clupea pallasi). Environmental Toxicology and Chemistry 18: 481-493.
Carls, M.G., Holland, L., Larsen, M., Collier, T.K., Scholz, N.L. and J.P. Incardona. 2008. Fish embryos are
damaged by dissolved PAHs, not oil particles. Aquatic Toxicology 88(2): 121-127.
Carmack, E.C. and R.W. Macdonald. 2002. Oceanography of the Canadian Shelf of the Beaufort Sea: A setting for
marine life. Arctic 55(Supplement 1): 29-45.
Carmichael, R.H., Graham, W.M., Aven, A., Worthy, G. and S. Howden. 2012a. Were multiple stressors a ‘perfect
storm’ for northern Gulf of Mexico bottlenose dolphins (Tursiops truncatus) in 2011? Plos One 7.
Carmichael, R.H., Jones, A.L., Patterson, H.K., Walton, W.C., Perez-Huerta, A., Overton, E.B., Dailey, M. and K.L.
Willett. 2012b. Assimilation of oil-derived elements by oysters due to the Deepwater Horizon Oil spill.
Environmental Science & Technology 46: 12787-12795.
Chase, D.A., Edwards, D.S., Qin, G., Wages, M.R., Willming, M.M., Anderson, T.A. and J.D. Maul. 2013.
Bioaccumulation of petroleum hydrocarbons in fiddler crabs (Uca minax) exposed to weathered MC-252
crude oil alone and in mixture with an oil dispersant. Science of the Total Environment 444: 121–127.
Cleveland, C.J. 2013. Deepwater Horizon oil spill. In: The Encyclopedia of Earth.
http://www.eoearth.org/view/article/161185/
Clumpner, C. 2015. Oiled Wildlife Preparedness for the Arctic: One Model from Alaska. International Bird Rescue,
Astoria, Oregon, USA. Platform Presentation at the 38th Arctic and Marine Oil Pollution Technical
Seminar, June 2-4, 2015, Vancouver, B.C.
Coelho, G., Clark, J. and D. Aurand. 2013. Toxicity testing of dispersed oil requires adherence to standardized
protocols to assess potential real world effects. Environmental Pollution 177: 185–188.
Colavecchia, M.V., Backus, S.M., Hodson, P.V. and J.L. Parrott. 2004. Toxicity of oil sands to early life stages of
fathead minnows (Pimephales promelas). Environmental Toxicology Chemistry 23: 1709-1718.
Colavecchia, M.V., Hodson, P.V. and J.L. Parrott. 2006. CYP1A induction and blue sac disease on early life stages
of white suckers (Catostomus commersoni) exposed to oil sands. Journal of Toxicology and Environmental
Health Part A 69: 1-28.
Colavecchia, M.V., Hodson, P.V. and J.L. Parrott. 2007. The relationships among CYP1A induction, toxicity and
eye pathology in early life stages of fish exposed to oil sands. Journal of Toxicology and Environmental
Health 70: 1542-1555.
Collier, T.K., Singh, S.V., Awasthi, Y.C. and U. Varanasi. 1992. Hepatic xenobiotic metabolizing enzymes in two
species of benthic fish showing different prevalences of contaminant- associated liver neoplasms.
Toxicology and Applied Pharmacology 113: 319-324.
Couillard, C.M., Lee, K., Legare, B. and T.L. King. 2005. Effect of dispersant on the composition of the water-
accommodated fraction of crude oil and its toxicity to larval marine fish. Environmental Toxicology and
Chemistry 24(6): 1496-1504.
Crowe, K.M., Newton, J.C., Kaltenboeck, B. and C. Johnson. 2014. Oxidative stress responses of gulf killifish
exposed to hydrocarbons from the Deepwater Horizon oil spill: Potential implications for aquatic food
resources. Environmental Toxicology and Chemistry 33: 370-374.
de Hoop, L., Schipper, A.M., Leuven, R.S.E.W., Huijbregts, M.A.J., Olsen, G.H., Smit, M.G.D. and A.J. Hendriks.
2011. Sensitivity of polar and temperate marine organisms to oil components. Environmental Science &
Technology 45(20): 9017-9023.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 433
de Soysa, T.Y., Ulrich, A., Friedrich, T., Pite, D., Compton, S.L., Ok, D., Bernardos, R.L., Downes, G.B., Hsieh, S.,
Stein, R., Lagdameo, M.C., Halvorsen, K., Kesich, L-R. and M.J.F. Barresi. 2012. Macondo crude oil from
the Deepwater Horizon oil spill disrupts specific developmental processes during zebrafish embryogenesis.
BMC Biology 10: doi:10.1186/1741-7007-10-40.
Debruyn, A.M.H., Wernick, B.G., Stefura, C., McDonald, B.G., Rudolph, B.-L., Patterson, L. and P.M. Chapman.
2007. In situ experimental assessment of lake whitefish development following a freshwater oil spill.
Environmental Science and Technology 41: 6983 -6989.
de Perre, C., Le Menach, K., Ibalot, F., Parlanti, E. and H. Budzinski. 2014. Development of solid-phase
microextraction to study dissolved organic matter-polycyclic aromatic hydrocarbon interactions in aquatic
environment. Analytica Chimica Acta 807: 51-60.
Dew, W.A., Hontela, A., Rood, S.B. and G.G. Pyle. 2015. Biological effects and toxicity of diluted bitumen and its
constituents in freshwater systems. Journal of Applied Toxicology 2015: doi: 10.1002/jat.3196.
Di Toro, D.M., McGrath, J.A. and D.J. Hansen. 2000. Technical basis for narcotic chemicals and polycyclic
aromatic hydrocarbon criteria. I. Water and tissue. Environmental Toxicology and Chemistry 19(8): 1951-
1970.
Dubansky, B., Whitehead, A., Miller, J.T., Rice, C.D. and F. Galvez. 2013. Multitissue molecular, genomic, and
developmental effects of the Deepwater Horizon Oil Spill on resident Gulf killifish (Fundulus grandis).
Environmental Science & Technology 47: 5074-5082.
Dupuis, A. and F. Ucan-Marin. 2014. A Literature Review on the Aquatic Toxicology of Petroleum Oil: An
Overview of Oil Properties and Effects to Aquatic Biota. Fisheries and Oceans Canada, Science Advisory
Secretariat, Ottawa, ON.
Dussauze, M., Pichavant-Rafini, K., Le Floch, S., Lemaire, P. and M. Theron. 2015. Acute toxicity of chemically
and mechanically dispersed crude oil on juvenile sea bass (Dicentrarchus labrax): Absence of synergistic
effects between oil and dispersants. Environmental Toxicology and Chemistry 34(7): 1543-1551.
Duesterloh, S., Short, J.W. and M.G. Barron. 2002. Photoenhanced toxicity of weathered Alaska North Slope crude
oil to the calanoid copepods Calanus marshallae and Metridia okhotensis. Environmental Science &
Technology 36: 3953-3959.
Engelhardt, R.F. 1983. Petroleum effects on marine mammals. Aquatic Toxicology 4: 199-217.
Fallahtafti, S., Rantanen, T., Brown, R.S., Snieckus, V. and P.V. Hodson. 2012. Toxicity of hydroxylated alkyl-
phenanthrenes to the early life stages of Japanese medaka (Oryzias latipes). Aquatic Toxicology 106-107:
56-64.
Finch, B.E., Wooten, K.J. and P.N. Smith. 2011. Embryotoxicity of weathered crude oil from the Gulf of Mexico in
mallard ducks (Anas platyrhynchos). Environmental Toxicology and Chemistry 30: 1885-1891.
Fodrie, F.J. and K.L. Heck, Jr. 2011. Response of coastal fishes to the Gulf of Mexico Oil disaster. PLoS ONE 6:
e21609.
Fodrie, F.J., Able, K.W., Galvez, F., Heck, K.L., Jensen, O.P., Lopez-Duarte, P.C., Martin, C.W., Turner, R.E. and
A. Whitehead. 2014. Integrating organismal and population responses of estuarine fishes in Macondo spill
research. Bioscience 64: 778-788.
Ford, R.G., Bonnell, M.L., Baroujean, D.H., Page, G.W., Carter, H.W., Sharp, B.E., Heinemann, D. and J.L. Casey.
1996. Total direct mortality of seabirds from the Exxon Valdez oil spill. American Fisheries Society
Symposium 18: 684-711
Fossi, M.C. and L. Marsili. 1997. The use of non-destructive biomarkers in the study of marine mammals.
Biomarkers 2: 205-216.
Fragoso, N., Hodson, P., Kozin, I., Brown, R. and J. Parrott. 1999. Kinetics of mixed function oxygenase induction
and retene excretion in retene-exposed rainbow trout. Environmental Toxicology and Chemistry 18: 2268-
2274.
Fremling, C.R. 1981. Impacts of a spill of No. 6 fuel oil on Lake Winona. In: Proceedings of the 1981 Oil Spill
Conference. American Petroleum Institute, Washington, DC. pp. 419-421.
French-McCay, D. P. 2002. Development and application of an oil toxicity and exposure model, OilToxEx.
Environmental Toxicology and Chemistry 21(10): 2080-2094.
Fry, B. and L.C. Anderson. 2014. Minimal incorporation of Deepwater Horizon oil by estuarine filter feeders.
Marine Pollution Bulletin 80: 282-287.
Garcia, T.I., Shen, Y.J., Crawford, D., Oleksiak, M.F., Whitehead, A.,and R.B. Walter. 2012. RNA-Seq reveals
complex genetic response to deepwater horizon oil release in Fundulus grandis. BMC Genomics 13.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 434
Garcia-Blanco, S., Venosa, A.D., Suidan, M.T., Lee, K., Cobanli, S. and J.R. Haines. 2007. Biostimulation for the
treatment of an oil-contaminated coastal salt marsh. Biodegradation 18(1): 1-15.
Gardiner, W.W., Word, J.Q., Word, J.D., Perkins, R.A., McFarlin, K.M., Hester, B.W., Word, L.S.,and C.M. Ray.
2013. The acute toxicity of chemically and physically dispersed crude oil to key arctic species under arctic
conditions during the open water season. Environmental Toxicology and Chemistry 32: 2284-2300.
Garr, A., Laramore, S.,and W. Krebs. 2014. Toxic effects of oil and dispersant on marine microalgae. Bulletin of
Environmental Contamination and Toxicology 93: 654-659.
Geiger, H. J., B. G. Bue, B.G., Sharr, S.,Wertheimer, A.C. and T. M. Willette. 1996. A life history approach to
estimating damage to Prince William Sound pink salmon caused by the Exxon Valdez oil spill. American
Fisheries Society Symposium 18: 487–498.
Gilfillan, E.S. 1992. Toxic Effects of Oil and Chemically Dispersed Oil on Marine Animals and Plants. Prepared for
the State of Maine, Department of Environmental Protection.
Gong, Y., Zhao, X., Cai, Z., O'Reilly, S.E., Hao, X. and D. Zhao. 2014. A review of oil, dispersed oil and sediment
interactions in the aquatic environment: Influence on the fate, transport and remediation of oil spills.
Marine Pollution Bulletin 79: 16-33.
Greenberg, B.M., Huang, X.D., Dixon, D.G., Ren, L., McConkey, B.J. and C.L. Duxbury. 1993. Quantitative
structure activity relationships for the photo-induced toxicity of polycyclic aromatic hydrocarbons to
duckweed - a preliminary model. In: Gorsuch, J.W., Dwyer, F.J., Ingersoll, C.G. and T.W.L. Point (eds.)
Environmental Toxicology and Risk Assessment. American Society for Testing and Materials, Philadelphia.
pp. 369-378.
Greer, C.D., Hodson, P.V., Li, Z., King, T. and L. Lee. 2012. Toxicity of crude oil chemically dispersed in a wave
tank to Atlantic herring (Clupea harengus) embryos. Environmental Toxicology and Chemistry 31: 1324-
1333.
Gulec, I. and D. Holdway. 2000. Toxicity of crude oil and dispersed crude oil to ghost shrimp Palaemon serenus
and larvae of Australian bass Macquaria novemaculeata. Environmental Toxicology 15: 91-98.
Häkkinen, J., Vehniäinen, E. and A. Oikari. 2004. High sensitivity of northern pike larvae to UV-B but no UV-
photoinduced toxicity of retene. Aquatic Toxicology 66: 393-404.
Haney, J.C., Geiger, H.J. and J.W. Short. 2014a. Bird mortality from the Deepwater Horizon oil spill. I. Exposure
probability in the offshore Gulf of Mexico. Marine Ecology Progress Series 513: 225-237.
Haney, J.C., Geiger, H.J. and J.W. Short. 2014b. Bird mortality from the Deepwater Horizon oil spill. II. Carcass
sampling and exposure probability in the coastal Gulf of Mexico. Marine Ecology Progress Series 513:
239-252.
Hatlen, K., Sloan, C.A., Burrows, D.G., Collier, T.K., Scholz, N.L. and J.P. Incardona. 2011. Natural sunlight and
residual fuel oils are an acutely lethal combination for fish embryos. Aquatic Toxicology 99: 56-64.
Hawkins, S.A., Billiard, S.M., Tabash, S.P., Depew, D.C., Brown, R.S. and P.V. Hodson. 2002. Altering
cytochrome P4501A activity affects polycyclic aromatic hydrocarbon metabolism and toxicity in rainbow
trout (Oncorhynchus mykiss). Environmental Toxicology and Chemistry 21: 1845-1853.
Hedgpeth, B.M. and R.J. Griffitt. 2015. Simultaneous exposure to chronic hypoxia and dissolved PAHs results in
reduced egg production and larval survival in the Sheepshead minnow (Cyprinodon variegatus).
Environmental Toxicology and Chemistry: doi: 10.1002/etc.3207.
Heintz, R.A. 2007. Chronic exposure to polynuclear aromatic hydrocarbons in natal habitats leads to decreased
equilibrium size, growth, and stability of pink salmon populations. Integrated Environmental Assessment
and Management 3: 351-363.
Heintz, R.A., Short, J.W. and S.D. Rice. 1999. Sensitivity of fish embryos to weathered crude oil: Part II. Increased
mortality of pink salmon (Oncorhynchus gorbuscha) embryos incubating downstream from weathered
Exxon Valdez crude oil. Environmental Toxicology and Chemistry 18: 494-503.
Heintz, R.A., Rice, S.D., Wertheimer, A.C., Bradshaw, R.F., Thrower, F.P., Joyce, J.E. and J.W. Short. 2000.
Delayed effects on growth and marine survival of pink salmon Oncorhynchus gorbuscha after exposure to
crude oil during embryonic development. Marine Ecology Progress Series 208: 205-216.
Hellebust, A., Hanna, B., Sheath, R.G., Gergis, M. and T.C. Hutchinson. 1975. Experimental crude oil spills on a
small subarctic lake in the Mackenzie valley, N.W.T.: Effects on phytoplankton, periphyton, and attached
aquatic vegetation. In: Proceedings of the 1975 International Oil Spill Conference, March 1975. American
Petroleum Institute, Washington, DC. pp. 509-515. doi.org/10.7901/2169-3358-1975-1-509

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 435
Hemmer, M.J., Barron, M.G. and R.M. Greene. 2011. Comparative toxicity of eight oil dispersants, Louisiana sweet
crude oil (LSC), and chemically dispersed LSC to two aquatic test species. Environmental Toxicology and
Chemistry 30: 2244-2252.
Henkel, J.R., Sigel, B.J. and C.M. Taylor. 2012. Large-scale impacts of the Deepwater Horizon oil spill: Can local
disturbance affect distant ecosystems through migratory shorebirds? Bioscience 62: 676-685.
Henkel, J.R., Sigel, B.J. and C,M. Taylor. 2014. Oiling rates and condition indices of shorebirds on the northern
Gulf of Mexico following the Deepwater Horizon oil spill. Journal of Field Ornithology 85: 408-420
Hicken, C.E., Linbo, T.L., Baldwin, D.H., Willis, M.L., Myers, M.S., Holland, L., Larsen, M., Stekoll, M.S., Rice,
S.D., Collier, T.K., Scholz, N.L. and J.P. Incardona. 2011. Sublethal exposure to crude oil during
embryonic development alters cardiac morphology and reduces aerobic capacity in adult fish. PNAS
108(17): 7086-7090.
Highsmith, R. C., Rucker, T. L., Stekoll, M. S., Saupe, S. M., Lindeberg, M. R., Jenne, R. N. and W.P. Erickson.
1996. Impact of the Exxon Valdez oil spill on intertidal biota. In: Proceedings of the “Exxon Valdez Oil
Spill Symposium” held at Anchorage, Alaska, February 2-5, 1993. American Fisheries Society Symposium.
1996. American Fisheries Society, Bethesda MD.
Hodson, P.V. and L.L. Williams. 2012. Oil spills from pipelines – why we should be concerned? Platform talk 326
at the 33rd Annual Meeting of the Society of Environmental Toxicology and Chemistry, Long Beach CA,
Nov 11-15, 2012.
Hodson, P.V., Ibrahim, I., Zambon, S., Ewert, A. and K. Lee. 2002. Bioavailability to fish of sediment PAH as an
indicator of the success of in situ remediation treatments at an experimental oil spill. Bioremediation
Journal 6: 297-313.
Hodson, P.V., Khan, C.W., Saravanabhavan, G., Clarke, L., Brown, R.S., Hollebone, B., Wang, Z., Short, J.W., Lee,
K. and T. King. 2007a. Alkyl PAH in crude oil cause chronic toxicity to early life stages of fish. In:
Proceedings of the 30th Arctic and Marine Oilspill Program Technical Seminar, Edmonton, AB. pp. 291-
300.
Hodson, P.V., Qureshi, K., Noble, C.A.J., Akhtar, P. and R.S. Brown. 2007b. Inhibition of CYP1A enzymes by
alpha-napthoflavone causes both synergism and antagonism of retene toxicity to rainbow trout
(Oncorhynchus mykiss). Aquatic Toxicology 81: 275-285.
Hodson, P.V., Collier, T.K. and J.D. Martin. 2011. Technical Data Report: Toxicity to Fish –Potential Effects of an
Oil Spill into the Kitimat River from a Northern gateway Pipeline Rupture. Enbridge Northern Gateway
Project.
Hollebone, B.P., Fieldhouse, B., Sergey, G., Lamter, P., Wang, Z., Yang, C. and M. Landriaut. 2011. The behaviour
of heavy oil in fresh water lakes. In: Proceedings of the 34th Arctic and Marine Oilspill Program (AMOP)
Technical Seminar on Environmental Contamination and Response. Environment Canada, Ottawa. Pp.
668-709.
Hooten, A. J. and R.C. Highsmith. 1996. Impacts on selected intertidal invertebrates in Herring Bay, Prince William
Sound, after the Exxon Valdez oil spill. In: Proceedings of the “Exxon Valdez Oil Spill Symposium” held
at Anchorage, Alaska, February 2-5, 1993. American Fisheries Society Symposium. 1996. American
Fisheries Society, Bethesda MD.
Incardona, J.P., Collier, T.K. and N.L. Scholz. 2004. Defects in cardiac function precede morphological
abnormalities in fish embryos exposed to polycyclic aromatic hydrocarbons. Toxicology and Applied
Pharmacology 196: 191-205.
Incardona, J.P., Carls, M.G., Teraoka, H., Sloan, C., Collier, T. and N. Scholz. 2005. Aryl hydrocarbon receptor-
independent toxicity of weathered crude oil during fish development. Environmental Health Perspectives
113: 1755-1762.
Incardona, J.P., Day, H.L., Collier, T.K. and N.L Scholz. 2006. Developmental toxicity of 4-ring polycyclic
aromatic hydrocarbons in zebrafish is differentially dependent on AH receptor isoforms and hepatic
cytochrome P4501A metabolism. Toxicology and Applied Pharmacology 217: 308-321.
Incardona, J.P., Carls, M.G., Day, H.L., Sloan, C.A., Bolton, J.L., Collier, T.K. and N.L. Scholz. 2009. Cardiac
arrhythmia is the primary response of embryonic Pacific herring (Clupea pallasi) exposed to crude oil
during weathering. Environmental Science &Technology 43: 201-207.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 436
Incardona, J., Ylitalo, G., Myers, M., Scholz, N., Collier, T., Vines, C., Griffin, F., Smith, E. and G. Cherr. 2011.
The 2007 Cosco Busan Oil Spill: Field and Laboratory Assessment of Toxic Injury to Pacific Herring
Embryos and Larvae in the San Francisco Estuary. Final Report. NOAA Fisheries, National Oceanic and
Atmospheric Administration, Northwest Fisheries Science Center, Environmental Conservation Division,
Ecotoxicology and Environmental Assessment Programs, Seattle, WA.
Incardona, J.P., Vines, C.A., Anulacion, B.F., Baldwin, D.H., Day, H.L., French, B.L., Labenia, J.S., Linbo, T.L.,
Myers, M.S., Olson, O.P., Sloan, C.A., Sol, S., Griffin, F.J., Menard, K., Morgan, S.G., West, J.E., Collier,
T.K., Ylitalo, G.M., Cherr, G.N. and N.L. Scholz. 2012a. Unexpectedly high mortality in Pacific herring
embryos exposed to the 2007 Cosco Busan oil spill in San Francisco Bay. Proceedings of the National
Academy of Sciences 109(2): E51-E58.
Incardona, J.P., Vines, C.A., Linbo, T.L., Myers, M.S., Sloan, C.A., Anulacion, B.F., Boyd, D., Collier, T.K.,
Morgan, S., Cherr, G.N. and N.L. Scholz. 2012b. Potent phototoxicity of marine bunker oil to translucent
herring embryos after prolonged weathering. PLoS ONE 7(2): e30116.
Incardona, J.P., Swarts, T.L., Edmunds, R.C., Linbo, T.L., Aquilina-Beck, A., Sloan, C.A., Gardner, L.D., Block,
B.A. and N.L. Scholz. 2013. Exxon Valdez to Deepwater Horizon: Comparable toxicity of both crude oils
to fish early life stages. Aquatic Toxicology 142-143: 303-316.
Incardona, J.P., Gardner, L.D., Linbo, T.L., Brown, T.L., Esbaugh, A.J., Mager, E.M., Stieglitz, J.D., French, B.L.,
Labenia, J.S., Laetz, C.A., Tagal, M., Sloan, C.A., Elizur, A., Benetti, D.D., Grosell, M., Block, B.A. and
N.L. Scholz. 2014. Deepwater Horizon crude oil impacts the developing hearts of large predatory pelagic
fish. Proceedings of the National Academy of Sciences 111(15): E1510-E1518.
Irons, D. B. 1996. Size and productivity of Black-legged Kittiwake colonies in Prince William Sound before and
after the Exxon Valdez oil spill. In: Proceedings of the “Exxon Valdez Oil Spill Symposium” held at
Anchorage, Alaska, February 2-5, 1993. American Fisheries Society Symposium. 1996. American Fisheries
Society, Bethesda MD.
Judson, R.S., Martin, M.T., Reif, D.M., Houck, K.A., Knudsen, T.B., Rotroff, D.M., Xia, M.H., Sakamuru, S.,
Huang, R.L., Shinn, P., Austin, C.P., Kavlock, R.J. and D.J. Dix. 2010. Analysis of eight oil spill
dispersants using rapid, in vitro tests for endocrine and other biological activity. Environmental Science &
Technology 44: 5979-5985.
Jung, J., Kim, M., Yim, U.H., Ha, S.Y., An, J.G., Won, J.H., Han, G.M., Kim, N.S., Addison, R.F. and W.J. Shim.
2011. Biomarker responses in pelagic and benthic fish over 1 year following the Hebei Spirit oil spill
(Taean, Korea). Marine Pollution Bulletin 62: 1859-66.
Jung, J-H., Chae, Y.S., Kim, H.N., Kim, M., Yim, U.H., Ha, S.Y., Han, G.M., An, J.G., Kim, E. and W.J. Shim.
2012. Spatial variability of biochemical responses in resident fish after the M/V Hebei Spirit oil spill
(Taean, Korea). Ocean Science Journal 47: 209-214.
Jung, J.-H., Hicken, C.E., Boyd, D., Anulacion, B.F., Carls, M.G., Shim, W.J. and J.P. Incardona. 2013.
Geologically distinct crude oils cause a common cardiotoxicity syndrome in developing zebrafish.
Chemosphere 91: 1146-1155.
Kelly, E.N., Schindler, D.W., Hodson, P.V., Short, J.W., Radmanovich, R. and C.W. Nielsen. 2010. Oil sands
development contributes elements toxic at low concentrations in the Athabasca River and its tributaries.
Proceedings of the National Academy of Sciences 107: 16178-16183.
Kennedy, C.J. and A.P. Farrell. 2008. Immunological alterations in juvenile Pacific herring, Clupea pallasi, exposed
to aqueous hydrocarbons derived from crude oil. Environmental Pollution 153: 638-648.
Khan, R.A. 1991. Influence of concurrent exposure to crude oil and infection with Trypanosoma murmanensis
(Protozoa, Mastigophora) on mortality in winter flounder, Pseudopleuronectes americanus. Canadian
Journal of Zoology 69: 876-880.
Kim, M., Yim, U.H., Hong, S.H., Jung, J.-H., Choi, H.-W., An, J., Won, J. and W.J. Shim. 2010. Hebei Spirit oil
spill monitored on site by fluorometric detection of residual oil in coastal waters off Taean, Korea. Marine
Pollution Bulletin 60(3): 383-389.
Kiparissis, Y, Akhtar, P., Hodson, P.V. and R.S. Brown. 2003. Partition Controlled Delivery (PCD) of toxicants: A
novel in vivo approach for embryotoxicity testing. Environmental Science & Technology 37: 2262-2266.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 437
Kocan, R.M., Hose, J.E., Brown, E.D. and T.T. Baker. 1996. Pacific herring (Clupea pallasi>) embryo sensitivity to
Prudhoe Bay petroleum hydrocarbons: laboratory evaluation and in situ exposure at oiled and unoiled sites
in Prince William Sound. Canadian Journal of Fisheries and Aquatic Sciences 53: 2366-2375.
Kuhl, A.J., Nyman, J.A., Kaller, M.D. and C.C. Green. 2013. Dispersant and salinity effects on weathering and
acute toxicity of South Louisiana crude oil. Environmental Toxicology and Chemistry 32: 2611-2620.
Kujawinksi, E.B., Soule, M.C.K., Valentine, D.L., Boysen, A.K., Longnecker, K. and M.C. Redmond. 2011. Fate of
dispersants associated with the Deepwater Horizon oil spill. Environmental Science & Technology 45:
1298-1306.
Landrum, P.F., Chapman, P.M., Neff, J. and D.S. Page. 2012. Evaluating the aquatic toxicity of complex organic
chemical mixtures: Lessons learned from polycyclic aromatic hydrocarbon and petroleum hydrocarbon
case studies. Integrated Environmental Assessment and Management 8: 217-230.
Laramore, S., Krebs, W. and A. Garr. 2014. Effects of Macondo Canyon 252 oil (naturally and chemically
dispersed) on larval Crassostrea virginica (Gmelin, 1791). Journal of Shellfish Research 33: 709-718.
Lari, E., Abtahi, B., Hashtroudi, M.S., Mohaddes, E. and K.B. Døving. 2015. The effect of sublethal concentrations
of water soluble fraction of crude oil on the chemosensory function of caspian roach, Rutilus caspicus
(Yakovlev, 1870). Environmental Toxicology and Chemistry 34(8):1826-32.
Lavoie, E.J., Tulley-Freiler, L., Bedenko, V. and D. Hoffmann. 1981. Mutagenicity tumor initiating activity, and
metabolism of methylphenanthrenes. Cancer Research 41: 3441-3447.
Lavoie, E.J., Tulley-Freiler, L., Bedenko, V. and D. Hoffman. 1982. Mutagenicity of substituted phenanthrenes in
Salmonella typhimurium. Mutation Research 116: 91.
Lee, K. 1983. Vanadium in the aquatic ecosystem. In: Nriagu, J.O. (ed.) Aquatic Toxicology: Advances in
Environmental Science and Technology, Volume 13. John Wiley & Sons, Inc., N.Y. pp. 155-187.
Lee, K., Wong, C.S., Cretney, W.J., Whitney, F.A., Parsons, T.R., Lalli, C.M. and J. Wu. 1985. Microbial response
to crude oil and Corexit 9527: SEAFLUXES enclosure study. Microbial Ecology 11: 337-351.
Lee, L.E.J., Stassen, J. and K. Lee. 2000. The mystery snail, Viviparus georgianus, as biomonitors of oil-spill and
bioremediation strategies in fresh-water habitats. In: Proceedings of the 27th Annual Aquatic Toxicity
Workshop, St. John's, Newfoundland, 1-4 October 2000. pp. 74.
Lee, K., Venosa, A.D., Suidan, M.T., Garcia-Blanco, S., Greer, C.W., Wohlgeschaffen, G., Cobanli, S.E., Tremblay,
G.H. and K.G. Doe. 2003a. Habitat recovery in an oil-contaminated salt marsh following biorestoration
treatments. In: Proceedings of the 2003 International Oil Spill Conference, Vancouver, BC. April 7-10,
2003. American Petroleum Institute Publication No. 14730B. pp. 977-982.
Lee, K., R.C. Prince, R.C., Greer, C.W., Doe, K.G., Wilson, J.E.H., Cobanli, S.E., Wohlgeschaffen, G.D., Alroumi,
D., King, T. and G.H. Tremblay. 2003b. Composition and toxicity of residual Bunker C fuel oil in intertidal
sediments after 30 years. Spill Science and Technology Bulletin 8: 187-199.
Lee, K., Boudreau, M., Bugden, J., Burridge, L., Cobanli, S.E., Courtenay, S., Grenon, S., Hollebone, B., Kepkay,
P., Li, Z., Lyons, M., Niu, H., King, T.L., MacDonald, S., McIntyre, E.C., Robinson, B., Ryan, S.S. and G.
Wohlgeschaffen. 2011. State of Knowledge Review of Fate and Effect of Oil in the Arctic Marine
Environment. A report prepared for the National Energy Board of Canada Arctic Roundtable, Fisheries and
Oceans Canada. 01/2011; National Energy Board of Canada. https://www.neb-one.gc.ca/ll-
eng/livelink.exe/fetch/2000/90463/621169/700096/704342/A2A8R2_-
_NEB_State_of_Knowledge_Review_of_Fate_and_Effect_of_Oil_DFO_COOGER.pdf?nodeid=704343&
vernum=0.
Lee, K., Budgen, J., Cobanli, S., King, T., McIntyre, C., Robinson, B., Ryan, S. and G. Wohlgeschaffen. 2012. UV-
Epifluorescence Microscopy Analysis of Sediments Recovered from the Kalamazoo River. Centre for
Offshore Oil, Gas and Energy Research (COOGER), Fisheries and Oceans Canada.
Leighton, F.A. 1993. The toxicity of petroleum oils to birds. Environmental Reviews 1: 92-103.
Lewis, M. and R. Pryor. 2013. Toxicities of oils, dispersants and dispersed oils to algae and aquatic plants: Review
and database value to resource sustainability. Environmental Pollution 180: 345-367.
Lin, H., Morandi, G.D., Brown, R.S., Snieckus, V., Rantanen, T., Jørgensen, K.B. and P.V. Hodson. 2015.
Quantitative structure-activity relationships for chronic toxicity of alkyl-chrysenes and alkyl-
benz[a]anthracenes to Japanese medaka embryos (Oryzias latipes). Aquatic Toxicology 159: 109-118.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 438
Little, E.E., Hurtubise, R. and L. Cleveland. 1998. Photoenhanced Toxicity of Diluent to the Frog, Rana
sphenocephala. Columbia Environmental Research Center, Unites States Geological Survey, Columbia,
MO.
Little, E., Cleveland, L., Calfee, R. and M. Barron. 2000. Assessment of the photoenhanced toxicity of a weathered
oil to the tidewater silverside. Environmental Toxicology and Chemistry 19: 926-932.
Litz, J.A., Baran, M.A., Bowen-Stevens, S.R., Carmichael, R.H., Colegrove, K.M., Garrison, L.P., Fire, S.E.,
Fougeres, E.M., Hardy, R., Holmes, S., Jones , W., Mase-Guthrie, B.E., Odell, D.K., Rosel, P.E., Saliki,
J.T., Shannon, D.K., Shippee, S.F., Smith, S.M., Stratton, E.M., Tumlin, M.C., Whitehead, H.R., Worthy,
G.A.J. and T.K. Rowles. 2014. Review of historical unusual mortality events (UMEs) in the Gulf of
Mexico (1990-2009): providing context for the multi-year northern Gulf of Mexico cetacean UME declared
in 2010. Diseases of Aquatic Organisms 112: 161-175.
Logan, K.A., Scott, D., Rosenberger, A. and M. MacDuffee. 2015. Potential Effects on Fraser River Salmon from an
Oil Spill by the Trans Mountain Expansion Project. Report prepared for the National Energy Board (NEB)
hearings reviewing Kinder Morgan’s proposed Trans Mountain Expansion project. Raincoast Conservation
Foundation, Sidney, BC. http://www.raincoast.org/wp-content/uploads/2015/06/RCF-Fraser-SS-Salmon-
NEB-May-2015.pdf
Longpré, D., Lee, K., Tremblay, G.H. and V. Jarry. 2000. The response of Scirpus pungens to crude oil
contaminated sediments. In: Proceedings of the 27th Annual Aquatic Toxicity Workshop, St. John's,
Newfoundland, 1-4 October 2000. pp. 127.
Loughlin, T.R., Ballachey, B.E. and B.A. Wright. 1996. Overview of studies to determine injury caused by the
Exxon Valdez oil spill to marine mammals. In: Proceedings of the “Exxon Valdez Oil Spill Symposium”
held at Anchorage, Alaska, February 2-5, 1993. American Fisheries Society, Bethesda, MD. pp. 798-808.
Madison, B.N., Hodson, P.V. and V.S. Langlois. 2015. Diluted bitumen causes deformities and molecular responses
indicative of oxidative stress in Japanese medaka embryos
Aquatic Toxicology 165: 222-230.
Mager, E.M., Esbaugh, A.J., Stieglitz, J.D., Hoenig, R., Bodinier, C., Incardona, J.P., Scholz, N.L., Benetti, D.D.
and M. Grosell. 2014. Acute embryonic or juvenile exposure to Deepwater Horizon Crude Oil impairs the
swimming performance of mahi-mahi (Coryphaena hippurus). Environmental Science & Technology 48:
7053-7061.
Martin, J.D., Adams, A., Hollebone, B.P., King, T., Brown, R.S. and P.V. Hodson. 2014. Chronic toxicity of heavy
fuel oils to fish embryos using multiple exposure scenarios. Environmental Toxicology and Chemistry 33:
677-687.
Marty, G.D., Short, J.W., Dambach, D.M., Willits, N.H., Heintz, R.A., Rice, S.D., Stegeman, J.J. and D.E. Hinton.
1997. Ascites, premature emergence, increased gonadal cell apoptosis, and cytochrome P4501A induction
in pink salmon larvae continuously exposed to oil-contaminated gravel during development. Canadian
Journal of Zoology 75: 989-1007.
Marty, G.D., Quinn, T.J., Carpenter, G., Meyers, T.R. and N.H. Willits. 2003. Role of disease in abundance of a
Pacific herring (Clupea pallasi) population. Canadian Journal of Fisheries and Aquatic Sciences 60: 1258-
1265.
Mathew, R., McGrath, J.A. and D.M. Di Toro. 2008. Modeling polycyclic aromatic hydrocarbon bioaccumulation
and metabolism in time-variable early life state exposures. Environmental Toxicology and Chemistry 27:
1515-1525
McCarty, L.S. and D. Mackay. 1993. Enhancing ecotoxicological modeling and assessment: Body residues and
modes of toxic action. Environmental Science & Technology 27: 1719-1728.
McGrath, J.A. and D.M. Di Toro. 2009. Validation of the target lipid model for toxicity assessment of residual
petroleum constituents: monocyclic and polycyclic aromatic hydrocarbons. Environmental Toxicology and
Chemistry 28: 1130-1148.
McIntosh, S., King, T., Wu, D. and P.V. Hodson. 2010. Toxicity of dispersed crude oil to early life stages of
Atlantic herring (Clupea harengus). Environmental Toxicology and Chemistry 29: 1160-1167.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 439
Melbye, A.G., Brakstad, O.G., Hokstad, J.N., Gregersen, I.K., Hansen, B.H., Booth, A.M., Rowland, S.J. and K.E.
Tollefsen. 2009. Chemical and toxicological characterization of an unresolved complex mixture-rich
biodegraded oil. Environmental Toxicology and Chemistry 28(9): 1815-1824.
Middaugh, D.P., Chapman, P.J., Shelton, M.E., McKenney, C.L. and L.A. Courtney. 2002. Effects of fractions from
biodegraded Alaska North Slope crude oil on embryonic inland silversides, Menidia beryllina. Archives of
Environmental Contamination and Toxicology 42: 236-243.
Misaki, K., Kawami, H., Tanaka, T., Handa, Y., Nakamura, M., Matsui, S., and T. Matsuda. 2007. Aryl hydrocarbon
receptor ligand activity of polycyclic aromatic ketones and polycyclic aromatic quinones. Environmental
Toxicology and Chemistry 26: 1370-1379.
Mitchelmore, C.L. and J.E. Baker. 2010. Acute and chronic effects of oil dipsersant and dispersed oil to sensitive
symbiotic cnidarian species, including corals. Final Project Report to the Coastal Response Research
Center. University of Maryland, Center for Environmental Science, Solomons, MD.
Moody, R.M., Cebrian, J. and K.L. Heck. 2013. Interannual recruitment dynamics for resident and transient marsh
species: Evidence for a lack of impact by the Macondo oil spill. Plos One 8.
Muhling, B.A., Roffer, M.A., Lamkin, J.T., Ingram, G.W., Upton, M.A., Gawlikowski, G., Muller-Karger, F.,
Habtes, S. and W.J. Richards. 2012. Overlap between Atlantic bluefin tuna spawning grounds and observed
Deepwater Horizon surface oil in the northern Gulf of Mexico. Marine Pollution Bulletin 64: 679-687.
Myers, M.S., Johnson, L.L. and T.K. Collier. 2003. Establishing the causal relationship between polycyclic aromatic
hydrocarbon (PAH) exposure and hepatic neoplasms and neoplasia-related liver lesions in English sole
(Pleuronectes vetulus). Human and Ecological Risk Assessment 9: 67-94.
Nahrgang, J., Camus, L., Carls, M.G., Gonzalez, P., Jonsson, M., Taban, I.C., Bechmann, R.K., Christiansen, J.S.
and H. Hop. 2010. Biomarker responses in polar cod (Boreogadus saida) exposed to the water soluble
fraction of crude oil. Aquatic Toxicology 97(3): 234-242.
Neff, J., Ostazeski, S., Gardiner, W. and I. Stejskal. 2000. Effects of weathering on the toxicity of three offshore
Australian crude oils and a diesel fuel to marine animals. Environmental Toxicology and Chemistry 19:
1809-1821.
Neff, J.M., Page, D.S., Landrum, P.F. and P.M. Chapman. 2013. The importance of both potency and mechanism in
dose-response analysis: An example from exposure of Pacific herring (Clupea pallasi) embryos to low
concentrations of weathered crude oil. Marine Pollution Bulletin 67: 7-15.
Nordtug, T., Olsen, A.J., Altin, D., Meier, S., Overrein, I., Hansen, B.H., and A. Johansen. 2011a. Method for
generating parameterized ecotoxicity data of dispersed oil for use in environmental modelling. Marine
Pollution Bulletin 62: 2106-2113.
Nordtug, T., Olsen, A.J., Altin, D., Overrein, I., Storoy, W., Hansen, B.H., and F. De Laender. 2011b. Oil droplets
do not affect assimilation and survival probability of first feeding larvae of North-East Arctic cod. Science
of the Total Environment 412-413: 148-153.
Nordtug, T., Olsen, A.J., Salaberria, I., Øverjordet, I.B., Altin, D., Størdal, I.F. and B.H. Hansen. 2015. Oil droplet
ingestion and oil fouling in the copepod Calanus finmarchicus exposed to mechanically and chemically
dispersed crude oil. Environmental Toxicology and Chemistry 34(8): 1899-1906.
NRC (National Research Council). 2005. Toxicological effects of dispersants and dispersed oil. In: Oil Spill
Dispersants. Efficacy and Effects. National Academies Press, Washington, DC. pp. 193-275.
NRC. 2013. An Ecosystem Services Approach to Assessing the Impacts of the Deepwater Horizon Oil Spill in the
Gulf of Mexico. National Academies Press Washington, DC.
Oakley, K. L. and K.J. Kuletz. 1996. Population, reproduction, and foraging of pigeon guillemots at Naked Island,
Alaska, before and after the Exxon Valdez oil spill. In: Proceedings of the “Exxon Valdez Oil Spill
Symposium” held at Anchorage, Alaska, February 2-5, 1993. American Fisheries Society Symposium.
1996. American Fisheries Society, Bethesda MD.
Olsen, G.H., Klok, C., Hendriks, A.J., Geraudie, P., De Hoop, L., De Laender, F., Farmen, E., Grøsvik, B.E.,
Hansen, B.H., Hjorth, M., Jansen, C.R., Nordtug, T., Ravagnan, E., Viaene, K. and J. Carroll. 2013.
Toxicity data for modeling impacts of oil components in an Arctic ecosystem. Marine Environmental
Research 90: 9-17.
Olsvik, P.l.A., Hansen, B.r.H., Nordtug, T., Moren, M., Holen, E., and K.K. Lie. 2011. Transcriptional evidence for
low contribution of oil droplets to acute toxicity from dispersed oil in first feeding Atlantic cod (Gadus
morhua) larvae. Comparative Biochemistry and Physiology Part C: Toxicology & Pharmacology 154(4):
333-345.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 440
Olsvik, P., Lie, K., Nordtug, T., and B.H. Hansen. 2012. Is chemically dispersed oil more toxic to Atlantic cod
(Gadus morhua) larvae than mechanically dispersed oil? A transcriptional evaluation. BMC Genomics
13(1): 702.
Owens, E.H., Sienkiewicz, A.M. and G.A. Sergy. 1999, Evaluation of Shoreline Cleaning Versus Natural Recovery:
The Metula Spill and the Komi Operations. In: Proceedings of the 1999 International Oil Spill Conference.
American Petroleum Institute, Washington, DC. Pp. 503-509.
Parkerton, T.F., Stone, M.A. and D.J. Letinski. 2000. Assessing the aquatic toxicity of complex hydrocarbon
mixtures using solid phase microextraction. Toxicology Letters 112–113: 273–282.
Pearson, W.H., Elston, R.A., Bienert, R.W., Drum, A.S. and L.D. Antrim. 1999. Why did the Prince William Sound,
Alaska, Pacific herring (Clupea pallasi) fisheries collapse in 1993 and 1994? Review of hypotheses.
Canadian Journal of Fisheries and Aquatic Sciences 56: 711-737.
Perkins, R.A., Rhoton, S., and C. Behr-Andres. 2005. Comparative marine toxicity testing: A cold-water species and
standard warm-water test species exposed to crude oil and dispersant. Cold Regions Science and
Technology 42(3): 226-236.
Place, B., Anderson, B., Mekebri, A., Furlong, E.T., Gray, J.L., Tjeerdema, R., and J. Field. 2010. A role for
analytical chemistry in advancing our understanding of the occurrence, fate, and effects of Corexit oil
dispersants. Environmental Science & Technology 44: 6016-6018.
Piatt, J. F. and P. Anderson. 1996. Response of common murres to the Exxon Valdez oil spill and long-term changes
in the Gulf of Alaska marine ecosystem. In: Proceedings of the “Exxon Valdez Oil Spill Symposium” held
at Anchorage, Alaska, February 2-5, 1993. American Fisheries Society, Bethesda, MD. pp. 720-737
Pilcher, W., Miles, S., Tang, S., Mayer, G., and A. Whitehead. 2014. Genomic and genotoxic responses to
controlled weathered-oil exposures confirm and extend field studies on impacts of the Deepwater Horizon
oil spill on native killifish. PLoS ONE 9: e106351.
Pollino, C.A. and D.A. Holdway. 2002. Toxicity testing of crude oil and related compounds using early life stages of
the crimson-spotted rainbowfish (Melanotaenia fluviatilis). Ecotoxicology and Environmental Safety 52:
180-189.
Raimondo, S., Jackson, C.R., Krzykwa, J., Hemmer, B.L., Awkerman, J.A., and M.G. Barron. 2014. Developmental
toxicity of Louisiana crude oil-spiked sediment to zebrafish. Ecotoxicology and Environmental Safety 108:
265-272.
Ramachandran, S.D., Hodson, P.V., Khan, C.W. and K. Lee. 2004a. Oil dispersant increases PAH uptake by fish
exposed to crude oil. Ecotoxicology and Environmental Safety 59: 300-308.
Ramachandran, S.D., Khan, C.W. and P.V. Hodson. 2004b. Role of droplets in promoting uptake of PAHs by fish
exposed to chemically dispersed crude oil. In: Proceedings of the Twenty-Seventh Arctic and Marine
Oilspill Program (AMOP) Technical Seminar Edmonton, AB. Environment Canada, Ottawa, ON. pp. 765–
772.
Ramachandran, S.D., Sweezey, M.J., Hodson, P.V., Boudreau, M., Courtenay, S., King, T., Dixon, J.A. and K. Lee.
2006. Influence of salinity and fish species on PAH uptake from dispersed MESA crude oil. Marine
Pollution Bulletin 52: 1182-1189.
Räsänen, K.M., Vehniäinen, E.-R., and A.O.J. Oikari. 2011. Different sensitivities of whitefish (Coregonius
lavaretus) and northern pike (Esox lucius) eleutheroembryos to photoinduced toxicity of polycyclic
aromatic hydrocarbons. Polycyclic Aromatic Hydrocarbons 31: 65-83.
Redman, A.D. 2015. Role of entrained droplet oil on the bioavailability of petroleum substances in aqueous
exposures. Marine Pollution Bulletin 97: 342-348.
Redman, A.D., McGrath, J.A., Stubblefield, W.A., Maki, A.W. and D.M. Di Toro. 2012. Quantifying the
concentration of crude oil microdroplets in oil–water preparations. Environmental Toxicology and
Chemistry 31(8): 1814-1822.
Redman, A.D., Parkerton, T.F., Letinski, D.J., Manning, R.G., Adams, J.E. and P.V. Hodson. 2014. Evaluating
toxicity of heavy fuel oil fractions using complimentary modeling and biomimetic extraction methods.
Environmental Toxicology and Chemistry 33: 2094-2104.
Redman, A.D., and T.F. Parkerton. 2015. Guidance for improving comparability and relevance of oil toxicity tests.
Marine Pollution Bulletin 98 (1-2): 156-170.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 441
Ren, L., Huang, X.D., MacConkey, B.J., Dixon, D.G., and B.M. Greenberg. 1994. Photoinduced toxicity of three
polycyclic aromatic hydrocarbons (Fluoranthene, pyrene, and naphthalene) to the duckweed Lemna gibba
L G-3. Ecotoxicology and Environmental Safety 28: 160-171.
Reynaud, S. and P. Deschaux. 2006. The effects of polycyclic aromatic hydrocarbons on the immune system of fish:
A review. Aquatic Toxicology 77: 229-238.
Rhodes, S., Farwell, A., Hewitt, L.M., MacKinnon, M., and D.G. Dixon. 2005. The effects of dimethylated and
alkylated polycyclic aromatic hydrocarbons on the embryonic development of the Japanese medaka.
Ecotoxicology and Environmental Safety 60: 247-258.
Rico-Martinez, R., Snell, T.W., and T.L. Shearer. 2013. Synergistic toxicity of Macondo crude oil and dispersant
Corexit 9500A® to the Brachionus plicatilis species complex (Rotifera). Environmental Pollution 173: 5-
10.
Rood, S. and E. Hillman. 2013. The 2012 Red Deer River oil spill: Analyzing environmental impacts in the
floodplain zone. Western Canadian Spill Services Ltd Newsletter 2(8): 4.
Rowe, C.L., Mitchelmore, C.L., and J.E. Baker. 2009. Lack of biological effects of water accommodated fractions
of chemically- and physically-dispersed oil on molecular, physiological, and behavioral traits of juvenile
snapping turtles following embryonic exposure. Science of the Total Environment 407(20): 5344-5355.
Rozas, L., Minello, T. and M.S. Miles. 2014. Effect of Deepwater Horizon Oil on growth rates of juvenile penaeid
shrimps. Estuaries and Coasts 37: 1403-1414.
RPI (Research Planning, Inc.). 1991. Sea Turtles and Oil–A Synopsis of the Available Literature. RPI/R/91/10/14-9.
Prepared for National Oceanic and Atmospheric Administration, Seattle, WA.
Saba, V.S., and J.R. Spotila. 2003. Survival and behavior of freshwater turtles after rehabilitation from an oil spill.
Environmental Pollution 126: 213-223.
Schein, A., Scott, J.A., Mos, L. and P.V. Hodson. 2009. Oil dispersion increases the apparent bioavailability and
toxicity of diesel to rainbow trout (Oncorhynchus mykiss). Environmental Toxicology and Chemistry 28:
595-602.
Schwacke, L., Smith, C., Townsend, F., Wells, R., Hart, L., Balme, B., Collier, T., De Guise, S., Fry, M., Guillette
Jr., L., Lamb , S., Lane, S., McFee , W., Place , N., Tumlin, M., Ylitalo, G., Zolman, E., and T. Rowles.
2014. Health of common bottlenose dolphins (Tursiops truncatus) in Barataria Bay, Louisiana, following
the Deepwater Horizon Oil Spill. Environmental Science & Technology 48: 93-103.
Scott, J.A. and P.V. Hodson. 2008. Evidence for multiple mechanisms of toxicity in larval rainbow trout
(Oncorhynchus mykiss) co-treated with retene and a-naphthoflavone. Aquatic Toxicology 88: 200-206.
Scott, J.A., Ross, M., Lemire, B.C. and P.V. Hodson. 2009. Embryotoxicity of retene in cotreatment with 2-
aminoanthracene, a cytochrome P4501A inhibitor, in rainbow trout (Oncorhynchus mykiss). Environmental
Toxicology and Chemistry 28: 1304-1310.
Scott, J.A., Incardona, J.P., Pelkki, K., Shepardson, S. and P.V. Hodson. 2011. AhR2-mediated, CYP1A-
independent cardiovascular toxicity in zebrafish (Danio rerio) embryos exposed to retene. Aquatic
Toxicology 101: 165–174.
Sellin Jeffries, M.K., Claytor, C., Stubblefield, W., Pearson, W.H. and J.T. Oris. 2013. Quantitative risk model for
polycyclic aromatic hydrocarbon photoinduced toxicity in Pacific herring following the Exxon Valdez oil
spill. Environmental Science & Technology 47: 5450-5458.
Sergy, G.A. and P.J. Blackhall. 1987. Conclusions of the Baffin Island oil spill project. Arctic 40: 1-Q.
Sergy, G.A. and E.H. Owens. 2011. Differences and similarities in freshwater and marine shoreline oil spill
response. In: Proceedings of the 2011 International Oil Spill Conference. p. abs62.
Shafir, S., VanRijn, J. and B. Rinkevich. 2007. Short and long term toxicity of crude oil and oil dispersants to two
representative coral species. Environmental Science & Technology 41: 5571-5574.
Shen, A., Tang, F. and X. Shen. 2010. Toxicity of crude oil and fuel oil on the embryonic development of the large
yellow croaker (Larmichthys crocea). In: Proceedings of 3rd International Conference on Biomedical
Engineering and informatics, Oct 16-18, 2010. Yantai, China.
Sharp, B. E., Cody, M. and R. Turner. 1996. Effects of the Exxon Valdez oil spill on the black oystercatcher. In:
Proceedings of the “Exxon Valdez Oil Spill Symposium” held at Anchorage, Alaska, February 2-5, 1993.
American Fisheries Society, Bethesda MD.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 442
Short, J.W. 2008. An Evaluation of Oil Fate, Persistence and Toxic Potential to Fish Following the August 2005
Canadian National Derailment Spill into Lake Wabamun, Alberta. August 27, 2008. https://docs.neb-
one.gc.ca/ll-eng/llisapi.dll/Open/2586376.
Short, J.W. 2015. The behaviour and environmental impacts of crude oil released into aqueous environments.
Presentation to the Royal Society of Canada Expert Panel on Oil in Water, Halifax, NS, April 10, 2015.
Shugart, L.R. 1995. Chap 13. Environmental genotoxicology. In: Rand, G.M. (Ed) Fundamentals of Aquatic
Toxicology. Effects, Environmental Fate, and Risk Assessment. 2nd ed.. Taylor and Francis, Washington,
DC. pp 405-419.
Sinclair, J. 2012. Appendix 2. Critique of the exposure and hazard assessment for river otters (Lutra lutra). In:
Hodson, P.V. and J. Martin (eds.) Review for the Haisla Nation Council, Kitimaat Village, BC, of the
Technical Data Report “Ecological and Human Health Risk Assessment for Pipeline Spills” by Enbridge
Northern Gateway Project”. pp 46-52.
Singer, M.M., Smalheer, D.L., Tjeerdema, R.S., and M. Martin. 1991. Effects of spiked exposure to an oil dispersant
on the early life stages of four marine species. Environmental Toxicology and Chemistry 10(10): 1367-
1374.
Singer, M.M., Aurand, D., Bragin, G.E., Clark, J.R., Coelho, G.M., Sowby, M.L. and R.S. Tjeerdema. 2000.
Standardization of the preparation and quantitation of water-accommodated fractions of petroleum for
toxicity testing. Marine Pollution Bulletin 40(11): 1007-1016.
Sørhus, E., Edvardsen, R.B., Karlsen, Ø., Nordtug, T., van der Meeren, T., Thorsen, A., Harman, C., Jentoft, S. and
S. Meier. 2015. Unexpected interaction with dispersed crude oil droplets drives severe toxicity in Atlantic
haddock embryos. PLoS ONE 10(4): e0124376.
Stagg, R.M., Rusin, J., McPhail, M.E., McIntosh, A.D., Moffat, C.F. and J.A. Craft. 2000. Effects of polycyclic
aromatic hydrocarbons on expression of cyp1a in salmon (Salmo salar) following experimental exposure
and after the Braer oil spill. Environmental Toxicology and Chemistry 19: 2797-2805.
Stout, S.A. 2015. Chemical Fingerprinting Assessment of Exposure of Dolphins to Macondo Oil During and After
the Deepwater Horizon Oil Spill. DWH NRDA Chemistry Technical Working Group Report (September 1,
2015)
Szedlmayer, S.T. and P.A. Mudrak. 2014. Influence of age-1 conspecifics, sediment type, dissolved oxygen, and the
Deepwater Horizon oil spill on recruitment of age-0 red snapper in the northeast Gulf of Mexico during
2010 and 2011. North American Journal of Fisheries Management 34: 443-452.
Tamis, J.E., Jongbloed, R.H., Karman, C.C., Koops, W. and A.J. Murk. 2012. Rational application of chemicals in
response to oil spills may reduce environmental damage. Integrated Environmental Assessment and
Management 8: 231-241.
Teal, J.M. and R.W. Howarth. 1984. Oil spill studies: A review of ecological effects. Environmental Management 8:
27-44.
Thormann, M.N. and S.E. Bayley. 2008. Impacts of the CN Rail Oil Spill on Softstem Bulrush-Dominated
Lacustrine Marshes in Wabamun Lake. Report submitted to Alberta Environment, Edmonton, AB.
Thorne, R.E. and G.L. Thomas. 2008. Herring and the “Exxon Valdez” oil spill: An investigation into historical data
conflicts. ICES Journal of Marine Sciences 65: 44-50.
Tollefsen, K.E., Petersen, K. and S.J. Rowland. 2012. Toxicity of synthetic naphthenic acids and mixtures of these
to fish liver cells. Environmental Science & Technology 46: 5143-4150.
Toor, N.S., E.D. Franza,, P.M. Fedorak, M.D. MacKinnon, and L. Karsten. 2013. Degradation and aquatic toxicity
of naphthenic acids in oil sands process-affected waters using simulated wetlands. Chemosphere 90: 449-
458.
Tran, T., Yazdanparast, A. and E.A. Suess. 2014. Effect of oil spill on birds: a graphical assay of the Deepwater
Horizon oil spill's impact on birds. Computational Statistics 29: 133-140.
Truscott, B., Walsh, J.M., Burton, M.P., Payne, J.F. and D.R. Idler. 1983. Effect of acute exposure to crude
petroleum on some reproductive hormones in salmon and flounder. Comparative Biochemistry and
Physiology Part C: Toxicology 75: 121-130.
Turcotte, D., Akhtar, P., Bowerman, M., Kiparissis, Y., Brown, R.S. and P.V. Hodson. 2011. Measuring the toxicity
of alkyl-phenanthrenes to early life stages of medaka (Oryzias latipes) using partition controlled delivery.
Environmental Toxicology and Chemistry 30: 487–495.
USFWS. 2015. Draft Damage Assessment and Restoration Plan/Environmental Assessment for the July 25-26, 2010
Enbridge Line 6B Oil Discharges near Marshall, MI. U.S. Fish and Wildlife Service, Nottawaseppi Huron

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 443
Band of the Potawatomi Tribe, Match-E-Be-Nash-She-Wish Band of the Pottawatomi Indians, National
Oceanic and Atmospheric Administration, Michigan Department of Environmental Quality, and Michigan
Department of Natural Resources.
Valentine, D.L., Burch Fisher, G., Bagby, S.C., Nelson, R.K., Reddy, C.M., Sylva, S.P. and M.A. Woo. 2014.
Fallout plume of submerged oil from Deepwater Horizon. Proceedings of the National Academy of
Sciences 111(45): 15906-15911
van der Ham, J.L. and K. de Mutsert. 2014. Abundance and size of Gulf shrimp in Louisiana's coastal estuaries
following the Deepwater Horizon Oil spill. PLoS ONE 9: e108884.
Van Tamelen, P. G. and M.S. Stekoll. 1996. Population response of the brown alga Fucus gardneri and other algae
in Herring Bay, Prince William Sound, to the Exxon Valdez oil spill. In: Proceedings of the “Exxon Valdez
Oil Spill Symposium” held at Anchorage, Alaska, February 2-5, 1993. American Fisheries Society,
Bethesda MD.
Vehniainen, E.R., Hakkinen, J. and A. Oikari. 2003. Photoinduced lethal and sublethal toxicityof retene, a
polycyclic aromatic hydrocarbon derived from resin acid, to coregonid larvae. Environmental Toxicology
and Chemistry 22: 2995-3000.
Venn-Watson, S., Colegrove, K.M., Litz, J., Kinsel, M., Terio, K., Saliki, J., Fire, S., Carmichael, R., Chevis, C.,
Hatchett, W., Pitchford, J., Tumlin, M., Field, C., Smith, S., Ewing, R., Fauquier, D., Lovewell, G.,
Whitehead, H., Rotstein, D., McFee, W., Fougeres, E. and T. Rowles. 2015. Adrenal gland and lung lesions
in Gulf of Mexico Common Bottlenose Dolphins (Tursiops truncatus) found dead following the Deepwater
Horizon oil spill. PLoS ONE 10(5): e0126538.
Venosa, A.D., Lee, K., Suidan, M.T., Garcia-Blanco, S. Cobanli, S., Moteleb, M., Haines, J.R., Tremblay, G. and M.
Hazelwood. 2002. Bioremediation and biorestoration of crude oil-contaminated freshwater wetland on the
St. Lawrence River. Bioremediation Journal 6: 261-281.
Vikebø, F.B., Rønningen, P., Meier, S., Grøsvik, B.E. and V.S. Lien. 2015. Dispersants have limited effects on
exposure rates of oil spills on fish eggs and larvae in Shelf seas. Environmental Science & Technology 49:
6061–6069
Vrabie, C.M., Candido, A., van Duursen, M.B.M. and M.T.O. Jonker. 2010. Specific in vitro toxicity of crude and
refined petroleum products: II. Estrogen (α and β) and androgen receptor-mediated responses in yeast
assays. Environmental Toxicology and Chemistry 29(7): 1529-1536.
Vrabie, C.M., Candido, A., van den Berg, H., Murk, A.J., van Duursen, M.B.M. and M.T.O. Jonker. 2011. Specific
in vitro toxicity of crude and refined petroleum products: 3. Estrogenic responses in mammalian assays.
Environmental Toxicology and Chemistry 30(4): 973-980.
Vrabie, C.M., Sinnige, T.L., Murk, A.J. and M.T.O. Jonker. 2012. Effect-directed assessment of the
bioaccumulation potential and chemical nature of ah receptor agonists in crude and refined oils.
Environmental Science & Technology 46(3): 1572-1580.
Wang, Y., Zhou, Q., Peng, S., Ma, L. and X. Niu. 2009. Toxic effects of crude-oil-contaminated soil in aquatic
environment on Carassius auratus and their hepatic antioxidant defense system. Journal of Environmental
Sciences 21: 612-617.
Wang, Z., Hollebone, B. P., Fingas, M., Fieldhouse, B., Sigouin, L., Landriault, M., Smith, P., Noonan, J., Thouin,
G. and J.W. Weaver. 2003. Characteristics of Spilled Oils, Fuels, and Petroleum Products: 1. Composition
and Properties of Selected Oils. National Exposure Research Laboratory, Office of Research and
Development, United States Environmental Protection Agency, Research Triangle Park, North Carolina.
Weber, D.D., Maynard, D.J., Gronlund, W.D. and V. Konchin. 1981. Avoidance reactions of migrating adult salmon
to petroleum hydrocarbons. Canadian Journal of Fisheries and Aquatic Sciences 38: 779-781.
Wernick, B., deBruyn, A.H., Patterson, L., and P. Chapman. 2009. Effects of an oil spill on the regrowth of
emergent vegetation in a Northern Alberta lake. Archives of Environmental Contamination and Toxicology
57: 697-706.
Whitehead, A., Dubansky, B., Bodinier, C., Garcia, T.I., Miles, S., Pilley, C., Raghunathan, V., Roach, J.L., Walker,
N., Walter, R.B., Rice, C.D. and F. Galvez. 2012. Genomic and physiological footprint of the Deepwater
Horizon oil spill on resident marsh fishes. Proceedings of the National Academy of Sciences 109(50):
20298-20302.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 444
Wiese, F.K. and P.C. Ryan. 2003. The extent of chronic marine oil pollution in southeastern Newfoundland waters
assessed through beached bird surveys 1984-1999. Marine Pollution Bulletin 46(9): 1090-1101.
Wikelski, M., Wong, V., Chevalier, B., Rattenborg, N. and H.L. Snell. 2002. Galapagos Islands: Marine iguanas die
from trace oil pollution. Nature 417(6889): 607-608.
Williams, R., Gero, S., Bejder, L., Calambokidis, J., Kraus, S.D., Lusseau, D., Read, A.J., and J. Robbins. 2011.
Underestimating the damage: interpreting cetacean carcass recoveries in the context of the Deepwater
Horizon/BP incident. Conservation Letters 4: 228-233.
Wise, J.P., Jr., Wise, J.T.F., Wise, C.F., Wise, S.S., Gianios, C., Jr., Xie, H., Thompson, W.D., Perkins, C., Falank,
C., and J.P. Wise, Sr. 2014. Concentrations of the genotoxic metals, chromium and nickel, in whales, tar
balls, oil slicks, and released oil from the Gulf of Mexico in the immediate aftermath of the Deepwater
Horizon oil crisis: Is genotoxic metal exposure part of the Deepwater Horizon legacy? Environmental
Science & Technology 48: 2997-3006.
Wu, D., Wang, Z., Hollebone, B., McIntosh, S., King, T., and P.V. Hodson. 2012. Comparative toxicity of four
chemically-dispersed and undispersed crude oils to rainbow trout embryos. Environmental Toxicology and
Chemistry 31: 754-765.
Yang, C., Wang, Z., Yang, Z., Hollebone, B., Brown, C.E., Landriault, M. and B. Fieldhouse. 2011. Chemical
fingerprints of Alberta Oil sands and related petroleum products. Environmental Forensics 12:173-188.
Zhang, J.F., Wang, X.R., Guo, H.Y., Wu, J.C., and Y.Q. Xue. 2004. Effects of water-soluble fractions of diesel oil
on the antioxidant defenses of the goldfish, Carassius auratus. Ecotoxicology and Environmental Safety
58(1): 110-116.

Chapter 5

Aeppli, C., Carmichael, C.A., Nelson, R.K., Lemkau, K.L., Graham, W.M., Redmond, M.C., Valentine, D.L., and
C.M. Reddy. 2012. Oil weathering after the Deepwater Horizon disaster led to the formation of oxygenated
residues. Environmental Science & Technology 46(16): 8799-8807.
Ajijolaiya, L. O., Hill, P.S., Khelifia, A., Islam, R.M., and K. Lee. 2006. Laboratory investigation of the effects of
mineral size and concentration on the formation of oil–mineral aggregates. Marine pollution bulletin 52(8):
920-927.
Allen, A.A. 1983. Oil Spill Response in the Arctic. Part 2: Field Demonstrations in Broken Ice. Prepared for Shell
Oil Company, Sohio Alaska Petroleum Company, Exxon Company, U.S.A., Amoco Production,
Archorage, AK.
ASCE Task Committee on Modeling of Oil Spills of the Water Resources Engineering Division. 1996. State-of-the-
art review of modeling transport and fate of oil spills. Journal of Hydraulic Engineering 122(11): 594-609.
Atlas, R. M. and R. Bartha. 1972. Degradation and mineralization of petroleum in sea water: limitation by nitrogen
and phosphorus. Biotechnology and Bioengineering 14: 309-318.
Atlas, R. and T. Hazen. 2011. Oil biodegradation and bioremediation: A tale of the two worst spills in U.S. history.
Environmental Science & Technology 45: 6709–6715.
Barker, C. 2011. A statistical outlook for the Deepwater Horizon oil spill. Geophysical Monograph Series 195: 237-
244.
Blokker, P. 1964. Spreading and evaporation of petroleum products on water. In: Proceedings of the Fourth
International Harbour Conference, Antwerp, Belgium. pp. 911-919.
Bobra, M. 1992. A Study of the Evaporation Of Petroleum Oils. Environment Canada, Environmental Protection
Directorate, River Road Environmental Technology Centre.
Bobra, M and Fingas, M. 1986. The behaviour and fate of arctic oil spills. Water Science & Technology 18:13–23.
Borden, R. C. and P. B. Bedient. 1986. Transport of dissolved hydrocarbons influenced by oxygen-limited
biodegradation. 1. Theoretical development. Water Resources Research 22(13): 1973-1982.
Boufadel, M. C. and A. M. Bobo. 2011. Feasibility of high pressure injection of chemicals into the subsurface for
the bioremediation of the Exxon Valdez oil. Ground Water Monitoring and Remediation 31(1): 59-67.
Boufadel, M. C., Suidan, M.T. and A.D. Venosa. 1999. A numerical model for density-and-viscosity-dependent
flows in two- dimensional variably-saturated porous media. Journal of Contaminant Hydrology 37: 1-20.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 445
Boufadel, M. C., Bechtel, R.D., and J. Weaver. 2006. The movement of oil under non-breaking waves. Marine
pollution Bulletin 52(9): 1056-1065.
Boufadel, M. C., Du, K., Kaku, V. and J. Weaver. 2007. Lagrangian simulation of oil droplets transport due to
regular waves. Environmental Modelling & Software 22(7): 978-986.
Boufadel, M.C., Sharifi, Y., Van Aken, B., Wrenn, B.A. and K. Lee. 2010. Nutrient and oxygen concentrations
within the sediments of an Alaskan beach polluted with the Exxon Valdez oil spill. Environmental Science
& Technology 44: 7418-7424.
Boufadel, M. C., Bobo, A.M. and Y. Xia. 2011. Feasibility of Deep Nutrients Delivery into a Prince William Sound
Beach for the Bioremediation of the Exxon Valdez Oil Spill. Ground Water Monitoring and Remediation
31(2): 80-91.
Boufadel, M.C., Abdollahi-Nasab, A., Geng, X., Galt, J. and J. Torlapati. 2014. Simulation of the landfall of the
Deepwater Horizon oil on the shorelines of the Gulf of Mexico. Environmental Science & Technology 48:
9496-9505.
Bragg, J.R., Prince, R.C., Harner, E.J. and R.M. Atlas. 1994. Effectiveness of bioremediation for the Exxon Valdez
oil spill. Nature 368: 413-418.
Brandvik, P. J, Johansen, O., Leirvik, F., Farooq, U. and P. S. Daling. 2013. Droplet breakup in sub-surface oil
releases—Part 1: Experimental study of droplet breakup and effectiveness of dispersant injection. Marine
Pollution Bulletin 73: 319-326.
Brost, E. J. and G.E. DuVall. 2000. Non-aqueous phase liquid (NAPL) mobility limits in soil. API Soil &
Groundwater Research Bulletin No.9.
Brustaert, W. 1982. Evaporation into the Atmosphere: Theory, History and Application. D. Reidel Publising
Company, Boston.
Buffington, J. M. and D. Tonina. 2009. Hyporheic exchange in mountain rivers II: Effects of channel morphology
on mechanics, scales, and rates of exchange. Geography Compass 3(3): 1038-1062.
Camilli, R., Di Iorio, D., Bowen, A., Reddy, C.M., Techet, A.H., Yoerger, D.R., Whitcomb, L.L., Seewald, J.S.,
Sylva, S.P. and J. Fenwick. 2012. Acoustic measurement of the Deepwater Horizon Macondo well flow
rate. Proceedings of the National Academy of Sciences USA 109(50): 20235-20239.
Camus, L., Olsen, G.H. and M. Carroll. 2008. Arctic Ecotoxicological Studies Review. Akvaplanniva AS Report
3975. Tromsø, Norway.
Carls, M.G., Thomas, R.E., Lilly, M.R. and S.D. Rice. 2003. Mechanism for transport of oil-contaminated
groundwater into pink salmon redds. Marine Ecology Progress Series 248: 245–255.
Chapelle, F. 2001. Ground-Water Microbiology and Geochemistry, John Wiley & Sons.
Clark, M. M. 1996. Transport Modeling for Environmental Engineers and Scientists. John Wiley and Sons Inc., New
York.
Clement, T. 1997. A Modular Computer Code for Simulating Reactive Multispecies Transport in 3–Dimensional
Groundwater Aquifers. Pacific Northwest National Laboratory, Richland, WA. http://bioprocess. pnl.
gov/rt3d. htm.
Cozzarelli, I. M., Bekins, B.A., Eganhouse, R.P., Warren, E. and H.I. Essaid. 2010. In situ measurements of volatile
aromatic hydrocarbon biodegradation rates in groundwater. Journal of Contaminant Hydrology 111(1-4):
48-64.
D’Auria, M., Emanuele, L., Racioppi, R. and V. Velluzzi V. 2009. Photochemical degradation of crude oil:
Comparison between direct irradiation, photocatalysis, and photocatalysis on zeolite. Journal of hazardous
materials 164(1): 32-38.
Dean, R. G. and R. A. Dalrymple. 1984. Water Wave Mechanics for Engineers and Scientists. Prentice-Hall,Inc.,
New Jersey.
Delvigne, G. A. L. 1994. Natural and chemical dispersion of oil. Journal of Advanced Marine Technology 11: 23-
63.
Delvigne, G. A. L. and C. E. Sweeney. 1988. Natural dispersion of oil. Oil & Chemical Pollution 4: 281-310.
Dickins, D., Brandvik, P.J., Bradford, J., Faksness, L-G., Liberty, L. and R. Daniloff. 2008. Svalbard 2006
experimental oil spill under ice: Remote sensing, oil weathering under Arctic conditions and assessment of
oil removal by in-situ burning. In: Proceedings of the 2008 International Oil Spill Conference. American
Petroleum Institute. pp. 681-688.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 446
Dickins, D., Hearon, G., Morris, K., Ambrosius, K. and W. Horowitz. 2011. Mapping sea ice overflood using
remote sensing: Alaskan Beaufort Sea. Cold Regions Science and Technology 65(3): 275-285.
El-Kadi, A. I. 2001. Modeling hydrocarbon biodegradation in tidal aquifers with water-saturation and heat inhibition
effects. Journal of Contaminant Hydrology 51(1-2): 97-125.
Elliot, A. J., Hurford, N. and C.J. Penn. 1986. Shear diffusion and the spreading of oil slicks. Marine pollution
Bulletin 17(7): 308-313.
Elliott, A. H. and N. H. Brooks. 1997. Transfer of nonsorbing solutes to a streambed with bed forms: Laboratory
experiments. Water Resources Research 33(1): 137-151.
Essaid, H. I., Bekins, B.A., Godsy, E.M., Warren, E., Baedecker, M.J. and Cozzarelli, I.M. 1995. Simulation of
aerobic and anaerobic biodegradation processes at a crude oil spill site. Water Resources Research 31(12):
3309-3327.
Essaid, H. I., Cozzarelli, I. M., Eganhouse, R.P., Herkelrath, W.N., Bekins, B.A., and Delin, G.N. 2003. Inverse
modeling of BTEX dissolution and biodegradation at the Bemidji, MN crude-oil spill site. Journal of
Contaminant Hydrology 67(1): 269-299.
Etkin, D. S., Michel, J., French McCay, D., Boufadel, M. and H. Li. 2008. Integrating state-of-the-art shoreline
interaction knowledge into spill modeling. In: Proceedings of the 2008 International Oil Spill Conference.
American Petroleum Institute.
Farmer, D. M. and M. Li. 1994. Oil dispersion by turbulence and coherent circulations. Ocean Engineering 21(6):
575-586.
Fay, J. A. 1971. The Spread of Oil Slicks on a Calm Sea. Dissertation/Thesis. Cambridge, MA, Massachusetts
Institute of Technology.
Fingas, M. 2011. Evaporation modeling. Chapter 9. In: Fingas, M.F. (ed.). Oil Spill Science and Technology:
Prevention, Response, and Cleanup. Elsevier/Gulf Professional Publication. pp. 201-242.
Fingas, M. 2013. Modeling oil and petroleum evaporation. Journal of Petroleum Science Research 2(3): 104-115.
Fingas, M. and B. Fieldhouse. 2004. Formation of water-in-oil emulsions and application to oil spill modelling.
Journal of Hazardous Materials 107(1): 37-50.
Fingas, M. and B. Hollebone. 2003. Review of behaviour of oil in freezing environments. Marine Pollution Bulletin
47(9): 333-340.
Fingas, M., Fieldhouse, B. and J. Mullin. 1996. Studies of water-in-oil emulsions: The role of asphaltenes and
resins. In: Proceedings of the Nineteenth Arctic and Marine Oilspill Program (AMOP) Technical Seminar,
Calgary, Alberta, Canada. Environment Canada. p. 73-88.
Fingas, M., Fieldhouse, B., Lane, J., and J. Mullin. 2000. Studies of water-in-oil emulsions: Long-term stability, oil
properties, and emulsions formed at sea. In: Proceedings of the Twenty-Third Arctic and Marine Oil Spill
Program (AMOP) Technical Seminar, Vancouver, British Columbia. Environment Canada. p. 145-160.
Fingas, M. F., Fieldhouse,B., Lambert, P., Wang, Z., Noonan, J. Lane, J. and J. Mullin. 2002. Water-in-oil
emulsions formed at sea, in test tanks, and in the laboratory. Environment Canada, Ottawa, ON.
Environment Canada Manuscript Report EE-169.
Fieldhouse, B., Lambert, P., Wang, Z., Noonan, J., Lane, J. and J.V. Mullin. 2002. Water-in-oil Emulsions Formed
at Sea, in Test Tanks, and in the Laboratory. Environment Canada Manuscript Report EE-169.Environment
Canada, Ottawa, Ontario.
Fischer, H. B., List, J.E., Koh, C.R., Imberger, J. and N.H. Brooks. 1979. Mixing in Inland and Coastal Waters.
Academic Press, New York.
Fitzpatrick, F.A., Boufadel, M.C., Johnson, R., Lee, K., Graan, T.P., Bejarano, A.C., Zhu, Z., Waterman, D.,
Capone, D.M., Hayter, E., Hamilton, S.K., Dekker, T., Garcia, M.H. and J.S. Hassan. 2015. Oil-Particle
Interactions and Submergence from Crude Oil Spills in Marine and Freshwater Environments – Review of
the Science and Future Science Needs. U.S. Geological Survey Open File Report 2015-1076.
http://dx.doi.org/10.3133/ofr20151076.
Frelichowska, J., Bolzinger, M.A., and Y. Chevalier. 2010. Effects of solid particle content on properties of o/w
Pickering emulsions. Journal of Colloid And Interface Science 351(2): 348-356.
French‐McCay, D. P. 2004. Oil spill impact modeling: Development and validation. Environmental Toxicology and
Chemistry 23(10): 2441-2456.
Garrett, R. M., Pickering, I.J., Haith, C.E. and R. C. Prince. 1998. Photooxidation of crude oils. Environmental
Science & Technology 32: 3719-3723.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 447
Geng, X., Boufadel, M.C. and B. Wrenn. 2013. Mathematical modeling of the biodegradation of residual
hydrocarbon in a variably-saturated sand column. Biodegradation 24(2): 153-163.
Geng, X., Boufadel, M.C., Personna, Y.R., Lee, K., Tsao, D. and E.D. Demicco. 2014. BioB: a mathematical model
for the biodegradation of low solubility hydrocarbons. Marine Pollution Bulletin 83(1): 138-147.
Geng, X., Boufadel, M.C., Lee, K., Abrams, S. and M. Suidan. 2015. Biodegradation of subsurface oil in a tidally
influenced sand beach: Impact of hydraulics and interaction with pore water chemistry. Water Resources
Research 51: 3193–3218.
Gong, Y., Zhao, X., Cai, Z., O'Reilly, S.E., Hao, X., and D. Zhao. 2014. A review of oil, dispersed oil and sediment
interactions in the aquatic environment: Influence on the fate, transport and remediation of oil spills.
Marine Pollution Bulletin 79: 16-33.
Guo, Q., Li, H., Boufadel,M.C. and Y. Sharifi. 2010. Hydrodynamics in a gravel beach and its impact on the Exxon
Valdez oil. Journal of Geophysical Research 115: C12077.
Hamoda, M., Hamam, S.E.M. and S.B. Shaban. 1989. Volatilization of crude oil from saline water. Oil and
Chemical Pollution 5(5): 321-331.
Harvey, J. W. and C. C. Fuller. 1998. Effect of enhanced manganese oxidation in the hyporheic zone on basin‐scale
geochemical mass balance. Water Resources Research 34(4): 623-636.
Hodson, P. V., Collier, T. K.and J. D. Martin. 2011. Toxicity to Fish -Potential Effects on an Oil Spill into the
Kitimat River from a Northern Gateway Pipeline Rupture. https://docs.neb-one.gc.ca/ll-
eng/llisapi.dll/fetch/2000/90464/90552/384192/620327/624910/693017/776343/D80-27-05_-
_Haisla_Nation_-_Hodson_Collier_Martin-Toxicity_of_Oil_to_Fish_-
_A2K3D7.pdf?nodeid=776350&vernum=-2
Hoult, D. P. 1972. Oil spreading on the sea. Annual Review of Fluid Mechanics 4: 341-368.
Johansen, Ø. 2000. Deepblow-A lagrangian plume model for deep water blowouts. Spill Science and Technology
Bulletin 6(2): 103-111.
Johansen, Ø., Rye, H., Melbye, A.G., Jensen, H.V., Serigstad, B. and T. Knutsen. 2001. Deep Spill JIP -
Experimental Discharges of Gas and Oil at Helland Hansen - June 2000. Technical Report. Prepared for
DeepSpill JIP. SINTEF, Trondheim, Norway.
Johansen, Ø., Brandvik, P.J. and U.Forooq. 2013. Droplet breakup in subsea oil releases – Part 2: Predictions of
droplet size distributions with and without injection of chemical dispersants. Marine Pollution Bulletin 73:
327-335.
Jokuty, P., Fingas, M., Whiticar, S. and B. Fieldhouse. 1995. A Study of Viscosity and Interfacial Tension of Oils
and Emulsions. Environmental Protection Service, Environment Canada, Ottawa, ON.
Jones, J. B. and P. J. Mulholland. 1999. Streams and Ground Waters. Academic Press.
Kazunga, C. and M. D. Aitken. 2000. Products from the incomplete metabolism of pyrene by polycyclic aromatic
hydrocarbon-degrading bacteria. Applied and Environmental Microbiology 66(5): 1917-1922.
Kindred, J. S. and M. A. Celia. 1989. Contaminant transport and biodegradation: 2. Conceptual model and test
simulations. Water Resources Research 25(6): 1149-1159.
King, T. L., Clyburne, J.A., Lee, K. and B.J. Robinson. 2013. Interfacial film formation: Influence on oil spreading
rates in lab basin tests and dispersant effectiveness testing in a wave tank. Marine Pollution Bulletin 71: 83-
91.
Kujawinski, E.B., Kido Soule, M.C., Valentine, D.L., Boysen, A.K., Longnecker, K. and M.C. Redmond. 2011. Fate
of dispersants associated with the Deepwater Horizon spill. Environmental Science and Technology 45:
1298-1306.
Langmuir, I. 1938. Surface motion of water induced by wind. Science 87(2250): 119-123.
Le Floch, S., Guyomarch, J., Merlin, F-X., Stoffyn-Egli, P. Dixon, J. and K. Lee.2002. The influence of salinity on
oil–mineral aggregate formation. Spill Science & Technology Bulletin 8(1): 65-71.
Lee, K. 2002. Oil–particle interactions in aquatic environments: Influence on the transport, fate, effect and
remediation of oil spills. Spill Science & Technology Bulletin 8(1): 3-8.
Lee, K. and P. Stoffyn-Egli. 2001. Characterization of oil mineral aggregates. In: Proceedings of the 2001
International Oil Spill Conference, Tampa, Florida, March 26-29, 2001. pp. 991-996.
Lee, K., Boudreau, M., Bugden, J., Burridge, L., Cobanli, S.E., Courtenay, S., Grenon, S., Hollebone, B., Kepkay,
P., Li, Z., Lyons, M., Niu, H., King, T.L., MacDonald, S., McIntyre, E.C., Robinson, B., Ryan, S.S. and G.
Wohlgeschaffen. 2011. State of Knowledge Review of Fate and Effect of Oil in the Arctic Marine
Environment. A report prepared for the National Energy Board of Canada Arctic Roundtable, Fisheries and
Oceans Canada. 01/2011; National Energy Board of Canada. https://www.neb-one.gc.ca/ll-

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 448
eng/livelink.exe/fetch/2000/90463/621169/700096/704342/A2A8R2_-
_NEB_State_of_Knowledge_Review_of_Fate_and_Effect_of_Oil_DFO_COOGER.pdf?nodeid=704343&
vernum=0.
Lehr, W., Fraga, R. J., Belen, M. S. and H.M. Cekirge. 1984. A new technique to estimate initial spill size using a
modified Fay-type spreading formula. Marine Pollution Bulletin 15(9): 326-329.
Leibovich, S. and D. Ulrich. 1972. A note on the growth of small‐scale Langmuir circulations. Journal of
Geophysical Research 77(9): 1683-1688.
Li, H. and M. C. Boufadel. 2010. Long-term persistence of oil from the Exxon Valdez spill in two-layer beaches.
Nature Geoscience 3(2): 96-99.
Li, H., Zhao, Q., Boufadel, M.C. and A.D. Venosa.2007. A universal nutrient application strategy for the
bioremediation of oil-polluted beaches. Marine Pollution Bulletin 54(8): 1146-1161.
Mackay, D. and P. L. Leinonen. 1977. Mathematical Model of the Behavior of Oil Spills on Water with
NaturalaAnd Chemical Dispersion. Environment Canada, Ottawa.
Mackay, D. and W. Zagorski. 1982. Studies of Water-in-Oil Emulsions. Report prepared for Environment Canada,
Environmental Protection Service, Environmental Impact Control Directorate, Environmental Emergency
Branch, Research And Development Division.
Mackay, D., Buist, I., Mascarenhas, R. and S. Paterson. 1980. Oil spill Processes and Models. Environment Canada
Report EE8. Ottawa, ON.
Maki, H., Sasaki, T. and S. Harayama. 2001. Photo-oxidation of biodegraded crude oil and toxicity of the photo-
oxidized products. Chemosphere 44(5): 1145-1151.
Molz, F. J., Widdowson, M.A. and L. D. Benefield. 1986. Simulation of microbial-growth dynamics coupled to
nutrient and oxygen-transport in porous-media. Water Resources Research 22(8): 1207-1216.
Mooney, M. 1951. The viscosity of a concentrated suspension of spherical particles. Journal of Colloid Science 6(2):
162-170.
Ng, C. G.-H., Bekins, B.A., Cozzarelli, I.M., Baedecker, M.J., Bennett, P.C. and R.T. Amos. 2014. A mass balance
approach to investigating geochemical controls on secondary water quality impacts at a crude oil spill site
near Bemidji, MN. Journal of Contaminant Hydrology 164: 1-15.
Nicol, J. P., Wise, W.R., Molz, F.J. and L.D. Benefield. 1994. Modeling biodegradation of residual petroleum in a
saturated porous column. Water Resources Research 30(12): 3313-3325.
Omotoso, O.E., Munoz, V.A. and R.J. Mikula. 2002. Mechanisms of crude oil–mineral interactions. Spill Science &
Technology Bulletin 8(1): 45-54
OSAT (Operational Science Advisory Team). 2011. Summary Report for Fate and Effects of Remnant Oil in the
Beach Environment, Prepared for U.S. Coast Guard Federal On-Scene Coordinator Deepwater Horizon
MC252.
Owens, E. H. and K. Lee. 2003. Interaction of oil and mineral fines on shorelines: review and assessment. Marine
Pollution Bulletin 47(9-12): 397-405.
Prince, R. C. 1993. Petroleum spill bioremediation in marine environments. Critical Reviews in Microbiology 19:
217-242.
Prince, R. C., Clark, J.R., Lindstrom, J.E., Butler, E.L., Brown, E.J., Winter, G., Grossman, M.J., Parrish, P.R.,
Bare, R.E., Braddock, J.F., Steinhauer, W.G., Douglas, G.S., Kennedy, J.M. and P.J. Barter. 1994.
Bioremediation of the Exxon Valdez oil spill: monitoring safety and efficacy. In: Hinchee, R. E., Miller,
R.N. and R. E. Hoeppel (eds.) Hydrocarbon Bioremediation. Lewis Publishers, Boca Raton, FL. pp. 107-
124.
Prince, R.C., Garrett, R. M., Bare, R.E., Grossman, M.J., Townsend, T., Suflita, J.M., Lee, K., Owens, E.H., Sergy,
G.A., Braddock, J.F., Lindstrom, J.E. and R. R. Lessard. 2003. The roles of photooxidation and
biodegradation in long-term weathering of crude and heavy fuel oils. Spill Science and Technology Bulletin
8: 145-156.
Pritchard, P.H., Mueller, J.G., Rogers, J.C., Kremer, F.V. and J.A. Glaser. 1992. Oil spill bioremediation:
experiences, lessons and results from the Exxon Valdex oil spill in Alaska. Biodegradation 3: 315-335.
Radović, J. R., Aeppli, C., Nelson, R. K., Jimenez, N., Reddy, C. M., Bayona, J. M., and J. Albaigés. 2014.
Assessment of photochemical processes in marine oil spill fingerprinting. Marine Pollution Bulletin 79(1):
268-277.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 449
Rittmann, B. E. and P. L. McCarty. 2001. Environmental Biotechnology: Principles and Applications. McGraw-Hill,
Boston.
Rønn, R., McCaig, A. E., Griffiths, B. S.and J. I. Prosser. 2002. Impact of protozoan grazing on bacterial community
structure in soil microcosms. Applied and Environmental Microbiology 68(12): 6094-6105.
Ross, S. and I. Buist.1995. Preliminary laboratory study to determine the effect of emulsification on oil spill
evaporation. In: Proceedings of the Eighteenth Arctic and Marine Oilspill Program (AMOP) Technical
Seminar, Edmonton, Alberta. Environment Canada.
Ross, S. and D. Dickins. 1987. Field Research Spills to Investigate the Physical and Chemical Fate of Oil in Pack
Ice. Environmental Studies Revolving Fund Report No. 062. Natural Resources Canada, Ottawa, ON.
Rutherford, J. C. 1994. River Mixing. John Wiley and Sons.
Ryan, R. J. and M. C. Boufadel. 2006. "Influence of streambed hydraulic conductivity on solute exchange with the
hyporheic zone." Environmental Geology 51(2): 203-210.
Ryan, R. J. and M. C. Boufadel. 2007a. Evaluation of streambed hydraulic conductivity heterogeneity in an urban
watershed. Stochastic Environmental Research and Risk Assessment 21(4): 309-316.
Ryan, R. J. and M. C. Boufadel. 2007b. Lateral and longitudinal variation of hyporheic exchange in a piedmont
stream pool. Environmental Science & Technology 41(12): 4221-4226.
Ryerson, T. B., Camilli, R., Kessler, J.D., Kujawinski, E.B., Reddy, C.M., Valentine, D.L., Atlas, E., Blake, D.R.,
de Gouw, J. and S. Meinardi. 2012. Chemical data quantify Deepwater Horizon hydrocarbon flow rate and
environmental distribution. Proceedings of the National Academy of Sciences 109(50): 20246-20253.
Schirmer, M., Molson, J. W., Frind, E.O. and J. F. Barker. 2000. Biodegradation modelling of a dissolved gasoline
plume applying independent laboratory and field parameters. Journal of Contaminant Hydrology 46(3-4):
339-374.
Short, J. 2015. Fate and Effect of Oil Spills from the Trans Mountain Expansion Project in Burrard Inlet and the
Fraser River Estuary A report Prepared for Tsleil-Waututh Nation, City of Vancouver, and Living Oceans
Society.
Stiver, W. and D. Mackay. 1984. Evaporation rate of spills of hydrocarbons and petroleum mixtures. Environmental
Science & Technology 18(11): 834-840.
Stoffyn-Egli, P. and K. Lee. 2002. Formation and characterization of oil–mineral aggregates. Spill Science &
Technology Bulletin 8(1): 31-44.
Sun, J., Khelifa, A., Zheng, X., Wang, Z., So, L.L., Wong, S., Yang, C. and B. Fieldhouse. 2010. A laboratory study
on the kinetics of the formation of oil-suspended particulate matter aggregates using the NIST-1941b
sediment. Marine Pollution Bulletin 60(10): 1701-1707.
Sutton, O. G. 1934. Wind structure and evaporation in a turbulent atmosphere. In: Proceedings of the Royal Society
of London. Series A, Containing Papers of a Mathematical and Physical Character. pp. 701-722.
Tchobanoglous, G., Burton, F. and H. D. Stensel. 2002. Wastewater Engineering: Treatment and Reuse. Metcalf and
Eddy, McGraw-Hill.
Tkalin, A. 1986. Evaporation of petroleum hydrocarbons from films on a smooth sea surface. Oceanology ONLGAE
26: 473-474.
Tonina, D. and J.M. Buffington. 2007. Hyporheic exchange in gravel bed rivers with pool-riffle morphology:
Laboratory experiments and three-dimensional modeling. Water Resources Research 43(W01421).
Tonina, D. and J.M. Buffington. 2009. Hyporheic exchange in mountain rivers I: Mechanics and environmental
effects. Geography Compass 3(3): 1063-1086.
Torlapati, J. and M. C. Boufadel. 2014. Evaluation of the biodegradation of Alaska North Slope oil in microcosms
using the biodegradation model BIOB. Frontiers in Aquatic Microbiology 5(212): 1-15.
Tsouris, C. and L. Tavlarides. 1994. Breakage and coalescence models for drops in turbulent dispersions. AIChE
Journal 40(3): 395-406.
Venkatesh, S., El-Tahan, H., Comfort, G. and R. Abdelnour. 1990. Modeling the behavior of oil spills in Ice-
infested waters. Atmosphere-Ocean 28: 303-329.
Venosa, A.D and E.L. Holder. 2007. Biodegradability of dispersed crude oil and two different temperatures. Marine
Pollution Bulletin 54(5): 545-553.
Venosa, A.D., Suidan, M.T., Wrenn, B.A., Strohmeier, K.L., Haines, J.R., Eberhart, B.L., King, D. and E. Holder.
1996. Bioremediation of an experimental oil spill on the shoreline of Delaware Bay. Environmental Science
and Technology 30(5): 1764-1775.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 450
Venosa, A. D., Campo, P. and M. T. Suidan. 2010. Biodegradability of lingering crude oil 19 years after the Exxon
Valdez oil spill. Environmental Science & Technology 44(19): 7613-7621.

Vilcáez, J., Li, L. and S. S. Hubbard. 2013. A new model for the biodegradation kinetics of oil droplets: application
to the Deepwater Horizon oil spill in the Gulf of Mexico. Geochemical Transactions 14(4).
Viveros, D., Jones, R. and M. Boufadel. 2015. The biodegradation model of the NOAA oil spill software ADIOS3.
In: Proceedings of the Thirty-Eighth Arctic and Marine Oilspill Program (AMOP), Vancouver, Canada,
June 2-5.
Wang, W., Zheng, Y. and K. Lee. 2013. Chemical dispersion of oil with mineral fines in a low temperature
environment. Marine Pollution Bulletin 72: 205-212
Weise, W. and G. Rheinheimer. 1977. Scanning electron microscopy and epifluorescence investigation of bacterial
colonization of marine sand sediments. Microbial Ecology 4(3): 175-188.
Widdowson, M., Waddill, D., Brauner, J., Chapelle, F. and P. Bradley. 2002. SEAM3D: A Numerical Model for
Three-Dimensional Solute Transport Coupled to Sequential Electron Acceptor-Based Biological Reactions
in Groundwater. Technical Report 24061. The Via Department of Civil and Environmental Engineering,
Virginia Polytechnic Institute and State University, Blacksburg, Virginia.
Wrenn, B.A., Suidan, M.T., Strohmeier, K.L., Eberhart, B.L., Wilson, G.J. and A.D. Venosa. 1997. Nutrient
transport during bioremediation of contaminated beaches: Evaluation with lithium as a conservative tracer.
Water Research 31: 515-524.
Xia, Y., Li, H., Boufadel, M.C. and Y. Sharifi. 2010. Hydrodynamic factors affecting the persistence of the Exxon
Valdez oil in a shallow bedrock beach. Water Resources Research 46: W10528-W10528.
Xie, H., Yapa, P. D. and K. Nakata. 2007. Modeling emulsification after an oil spill in the sea. Journal of Marine
Systems 68(3): 489-506.
Yang, W. C. and H. Wang. 1977. Modeling of oil evaporation in aqueous environment. Water Research 11(10):
879-887.
Yang, Z., Hollebone, B. P., Wang, Z., Yang, C. Brown, C., Zhang, G., Landriault, M. and X. Ruan. 2015. A
preliminary study for the photolysis behavior of biodiesel and its blends with petroleum oil in simulated
freshwater. Fuel 139: 248-256.
Yassine, M. H., Suidan, M. T. and A. D. Venosa. 2013. Microbial kinetic model for the degradation of poorly
soluble organic materials. Water Research 47(4): 1585-1595.
Zeidan, E., Zahariev, K., Li, M. and C. Garrett. 1997. The breakup of oil spills in the marine environment. In:
Proceedings of 20th Arctic Marine Oilspill Program (AMOP) Technical Seminar, Vancouver, Canada.
Environment Canada.
Zhao, L., Boufadel, M.C., Socolofsky, S.A., Adams, E., King, T. and K. Lee. 2014a. Evolution of droplets in subsea
oil and gas blowouts: Development and validation of the numerical model VDROP. Journal of Marine
Pollution Bulletin 83: 58-69.
Zhao, L., Torlapati, J., Boufadel, M.C., King, T., Robinson, B. and K. Lee. 2014b. VDROP: A comprehensive
model for droplet formation of oils and gases in liquids-Incorporation of the interfacial tension and droplet
viscosity. Chemical Engineering Journal 253: 93-106.
Zhao, L., Boufadel, M. C., Adams, E.E., Socolofsky, S., King, T. and K. Lee. 2015. Simulation of scenarios of oil
droplet formation from the Deepwater Horizon blowout. Marine Pollution Bulletin. Accepted.
Zheng, L., Yapa, P. D. and F. Chen. 2003. A model for simulating deepwater oil and gas blowouts-Part I: Theory
and model formulation. Journal of Hydraulic Research 41(4): 339-351.
Zhu, X., Venosa, A.D., Suidan M.T. and K. Lee. 2001. Guidelines for the Bioremediation of Marine Shorelines and
Freshwater Wetlands. U.S. Environmental Protection Agency.

Chapter 6

API. 2015. Field operations guide for in-situ burning of on-water oil spills, API Technical Report 1252, 1st Edition.
American Petroleum Institute, Washington, DC.
Andrade, F., Lutz, G., Lyra, A. P., Ranieri, A. and B. Lemos. 2013. Oil Spill Response Equipment and Techniques
for Waxy Oils in Brazil. In: Proceedings of the SPE Latin-American and Caribbean Heath, Safety,
Environment and Social Responsibility Conference. Society of Petroleum Engineers.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 451
Arctic Oil Spill Response Technology Joint Industry Programme. 2013. Dispersant Testing Under Realistic
Conditions. Final Report 2.1. Report from Joint Industry Programme (JIP) to identify and summarise the
state-of-the-art on research conducted to date on the effectiveness of dispersant and mineral fines in ice.
http://www.arcticresponsetechnology.org/wp-content/uploads/2013/10/Report%202.1%20-
%20DISPERSANT%20TESTING%20UNDER%20REALISTIC%20CONDITIONS.pdf
Atlas, R. and T. Hazen. 2011. Oil biodegradation and bioremediation: A tale of the two worst spills in U.S. history.
Environmental Science & Technology 45: 6709–6715.
Aurand, D. and G. Coelho. 2006a. Ecological Risk Assessment Workshop: Environmental Tradeoffs Associated
with Oil Spill Response Technologies. Cape Flattery, Washington. A report to Regional Response Team X.
Ecosystem Management & Associates, Inc., Lusby, MD. Technical Report 05-01.
Aurand, D. and G. Coelho. 2006b. Ecological Risk Assessment: Consensus Workshop. Environmental Tradeoffs
Associated with Oil Spill Response Technologies. A Report to the US Coast Guard, Sector Delaware Bay.
Ecosystem Management & Associates, Inc., Lusby, MD. Technical Report 06-01.
Bento, F.M., de Oliveira-Camargo, F.A., Okeke, B.C. and W.T. Frankenberger, Jr. 2005. Diversity of biosurfactant
producing microorganisms isolated from soils contaminated with diesel oil. Microbiological Research 160
(3): 249-255.
Bragg, J.R., Prince, R.C., Harner, E.J. and R.M. Atlas. 1994. Effectiveness of bioremediation for the Exxon Valdez
oil spill. Nature 368: 413-418.
Brinkman, J.S. 1998. Absorbents vs. adsorbents. Hazmat Management Sept. 1, 1998.
http://www.hazmatmag.com/features/absorbents-vs-adsorbents/
Buist I., Potter, S., Nedwed, T. and J. Mullin. 2008. Herding agents thicken oil spills in drift ice to facilitate in situ
burning: A new trick for an old dog. In: Proceedings of the 2008 International Oil Spill Conference,
Savannah. American Petroleum Institute, Washington, DC. Pp. 673–680.
Buist I., Canevari, G. and T. Nedwed. 2010a. New herding agents for thickening oil slicks in drift ice for in situ
burning. In: Proceedings of the Thirty-third AMOP Technical Seminar on Environmental Contamination
and Response. Environment Canada, Ottawa. pp 725–742.
Buist I. , Potter, S. and S.E. Sørstrøm. 2010b. Barents Sea field test of herder to thicken oil for in situ burning. In:
Proceedings of the Thirty-third AMOP Technical Seminar on Environmental Contamination and Response.
Environment Canada, Ottawa. pp 1085–1108.
Buist, I., Potter, S. and T. Nedwed. 2011. Herding agents to thicken oil spills in drift ice for in situ burning: New
developments. In: Proceedings of the 2011 International Oil Spill Conference. American Petroleum
Institute, Washington, DC. pp. abs230.
Burns, K.A., Codi, S. and N.C. Duke. 2000. Gladstone, Australia field studies: weathering and degradation of
hydrocarbons in oiled mangrove and salt marsh sediments with and without the application of an
experimental bioremediation protocol. Marine Pollution Bulletin 41: 392-402.
Cai, Q., Zhang, B., Chen, B., Zhu, Z., Lin, W. and T. Cao. 2014a. Screening of biosurfactant producers from
petroleum hydrocarbon contaminated sources in cold marine environments. Marine Pollution Bulletin
86(1): 402-410.
Cai, Q.H., Zhang, B.Y., Chen, B. and T. Cao. 2014b. Biodispersant produced by a Rhodococcus Erythropolis mutant
as an oil spill response agent. In: Proceedings of the International Conference on Marine and Freshwater
Environments (iMFE 2014), August 6-8, St. John's, Canada. MOS1137:1-12.
Canadian Coast Guard. 2005. Oil Spill Response Field Guide (4th Printing).
Chroma, L., Mackova, M., Kucerova, P., der Wiesche, C., Burkhard, J. and T. Macek. 2002. Enzymes in plant
metabolism of PCBs and PAHs. Acta Biotechnologica 22 (1-2): 35–41.
Chung, S., Suidan, M.T. and A.D. Venosa, 2012. Effectiveness of plant-derived sorbents for the remediation of low-
energy intertidal wetlands contaminated by oil spills. Water, Air, and Soil Pollution 223 (6): 3137-3144.
Clayton, J.R., Payne, J.R. and J.S. Farlow. 1993. Oil Spill Dispersants: Mechanisms of Action and Laboratory Tests.
CRC Press, Inc., Boca Raton, FL.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 452
Colten, C. E., Hay, J. and A. Giancarlo. 2012. Community resilience and oil spills in coastal Louisiana. Ecology and
Society 17(3): 5.
Criquet, S., Joner, E., Leglize, P. and C. Leyval. 2000. Anthracene and mycorrhiza affect the activity of
oxidoreductases in the roots and the rhizosphere of lucerne (Medicago sativa L.). Biotechnology Letters 22
(21): 1733–1737.
Crosby, S., Fay, R., Groark, C., Kani, A., Smith, J.R. and T. Sullivan. 2013. Transporting Alberta’s Oil Sands
Products: Defining the Issues and Assessing the Risks.
http://www.rrt10nwac.com/Files/FactSheets/131217071129.pdf
Cunningham, S.D., Anderson, T.A., Schwab, A.P. and F.C. Hsu. 1996. Phytoremediation of soils contaminated with
organic pollutants. Advances in Agronomy 56: 55-114.
Dahl, W.A., Lessard, R.R. and E.A. Cardello. 1997. Recent research on the application and practical effects of
solidifiers. In: Proceedings of the 1997 International Oil Spill Conference. American Petroleum Institute,
Washington, DC. Pp. 391-395.
Daling, P.S., Holumsnes, A., Rasmussen, C., Brandvik, P.J. and F. Leirvik. 2010. Development and Testing of a
Containerized Dispersant Spray System for Use in Cold and Ice-Covered Areas. Oil-in-Ice JIP Report No.
13/ARCTECH P4. SINTEF Materials and Chemistry, Trondheim, Norway.
http://www.sintef.no/project/JIP_Oil_In_Ice/Dokumenter/publications/JIP-rep-no-13-Development-of-
spray-arm-final.pdf. (accessed June 12, 2015).
Dejhosseini, M., Aida, T., Watanabe, M., Takami, S., Hojo, D., Aoki, N., Arita, T., Kishita, A. and T. Adschiri.
2013. Catalytic cracking reaction of heavy oil in the presence of cerium oxide nanoparticles in supercritical
water. Energy & Fuels 27(8): 4624-4631.
Delaune, R.D., Lindau, C.W. and A. Jugsujinda. 1999. Effectiveness of “Nochar” solidifier polymer in removing oil
from open water in coastal wetlands. Spill Science and Technology Bulletin 5: 357–359.
Demirbas, A., Alidrisi, H. and M.A. Balubaid. 2015. API gravity, sulfur content, and desulfurization of crude oil.
Petroleum Science and Technology 33(1): 93-101.
Deng, D., Prendergast, D. P., MacFarlane, J., Bagatin, R., Stellacci, F. and P.M. Gschwend. 2013. Hydrophobic
meshes for oil spill recovery devices. ACS Applied Materials & Interfaces 5(3): 774-781.
Dickins, D. 2011. Behavior of oil spills in ice and implications for arctic spill response. In: Arctic Technology
Conference, OTC 22126, February 7-9, Texas. Offshore Technology Conference.
Duke, N.C., Burns, K.A., Swannell, R. P. J., Dalhaus, O. and R.D. Rupp. 2000. Dispersant use and a bioremediation
strategy as alternate means of reducing impacts of large oil spills on mangroves: The Gladstone field trials.
Marine Pollution Bulletin 41: 403-412.
Doerffer, J.W. 1992. Oil Spill Response in the Marine Environment. Pergamon Press, Oxford, U.K.
Doerffer, J. W. 2013. Oil spill Response in the Marine Environment. Elsevier.
EPPR (Emergency Prevention, Preparedness and Response). 2015. Guide to Oil Spill Response in Snow and Ice
Conditions. Narayana Press, Denmark.
Etkin, D.S. and P. Tebeau. 2003. Assessing progress and benefits of oil spill response technology development since
Exxon Valdez. In: Proceedings of the 2003 International Oil Spill Conference. American Petroleum
Institute, Washington, D.C.
Favas, P.J., Pratas, C.J., Varun, M., D’Souza, R. and M.S. Paul. 2014. Chapter 17. Phytoremediation of Soils
Contaminated with Metals and Metalloids at Mining Areas: Potential of Native Flora. In: Soriano, M.C.
(ed.) Environmental Risk Assessment of Soil Contamination. InTech: DOI: 10.5772/57086.
Federal Interagency Solutions Group. 2010. Oil Budget Calculator: Deepwater Horizon. A Report to the National
Incident Command, NOAA, USGS, and NIST.
http://www.noaanews.noaa.gov/stories2010/PDFs/OilBudgetCalc_Full_HQ-Print_111110.pdf
Ferek, R., Allen, A. and J.H. Kucklick. 1997. Air Quality Considerations Involving In-Situ Burning. Marine
Preservation Association, Scottsdale, AZ.
Ferro, A. M., Kennedy, J., Doucette, W., Nelson, S., Jauregui, G., McFarland, B. and B. Bugbee. 1997. Fate of
benzene in soils planted with alfalfa: uptake, volatilization, and degradation. In: Kruger, E.L., Anderson,

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 453
T.A. and J.R. Coates (eds) Phytoremediation of Soil and Water Contaminants. American Chemical Society,
Washington, DC. pp. 223-237.
Fieldhouse, B. and M. Fingas. 2009. Solidifier effectiveness: variation due to oil composition, oil thickness and
temperature. In: Proceedings of the 32nd AMOP Technical Seminar on Environmental Contamination and
Response. Environment Canada, Ottawa. 337-363.
Fingas, M. 2008. A Review of Literature Related to Oil Spill Solidifiers 1990–2008. Prince Williams Sound
Regional Citizens’ Advisory Council (PWSRCAC), Anchorage, AK.
Fingas, M.F. 2011. Oil Spill Science and Technology: Prevention, Response, and Cleanup. Elsevier/Gulf
Professional Publication.
Fingas, M.F. 2012. Buoys and devices for oil spill tracking. In: Proceedings of the 35th AMOP Technical Seminar
on Environmental Contamination and Response, Vancouver, Canada. Environment Canada, Ottawa. Pp.
213-228.
Fingas, M. 2013a. Review of Solidifiers: An Update 2013. Prince Williams Sound Regional Citizens’ Advisory
Council (PWSRCAC). http://www.pwsrcac.org/resources/reports-documents/
Fingas, M. 2013b. Review of Oil Spill Herders. Prince Williams Sound Regional Citizens’ Advisory Council
(PWSRCAC). http://www.pwsrcac.org/resources/reports-documents/
Fingas, M. 2013c. Macondo well blowout mass balance: a chemical view. In: Proceedings of the 36th AMOP
Technical Seminar on Environmental Contamination and Response, Environment Canada, Ottawa.
Fingas, M.F., Kolokowski, B. and E.J. Tennyson. 1990a. Study of oil spill dispersants effectiveness and physical
studies. In: Proceedings of the 13th Arctic and Marine Oilspill Program Technical Seminar, Edmonton,
Alberta. Environment Canada. Pp. 265-287.
Fingas, M.F., Stoodley, R. and N. Laroche. 1990b. Effectiveness testing of spill-treating agents. Oil Chemistry and
Pollution 7: 337–348.
Fingas, M.F., Halley, G., Ackerman, F., Nelson, R., Bissonnette, M.C., Laroche, N., Wang, Z., Lambert, P., Li, K.,
Jokuty, P., Sergy, G., Halley, W., Latour, J., Galarneau, R., Ryan, B., Campagna, P.R., Turpin,R.D.,
Tennyson, E.J., Mullin, J., Hannon, L., Aurand, D. and R. Hiltabrand, 1995a. The Newfoundland Offshore
Burn Experiment. In: Proceedings of the 1995 International Oil Spill Conference. American Petroleum
Institute, Washington, D.C. pp. 123-132.
Fingas, M.F., Kyle, D.A., Larouche, N., Fieldhouse, B., Sergy, G. and G. Stoodley. 1995b. Effectiveness Testing of
Oil Spill-treating Agents. ASTM Special Technical Publication 1252. pp. 286–298.
Forsyth, J.V., Tsao, Y.M. and R.D. Blem. 1995. Bioremediation: when is augmentation needed? In: Hinchee, R.E.,
Alleman, B.C. and J. Frederickson (eds.) Bioaugmentation for Site Remediation. Battelle Press, Columbus,
OH. pp1-14.
Frick, C.M., Germida, J.J. and R.E. Farrell. 1999. Assessment of phytoremediation as an in-situ technique for
cleaning oil-contaminated sites. In: Proceedings of the Phytoremediation Technical Seminar. Environment
Canada, Ottawa. Pp. 105-124.
Frielo, D.A., Mylroie, J.R. and A.M. Chakrabarty. 2001. Use of genetically engineered multi-plasmid
microorganisms for rapid degradation of fuel hydrocarbons. International Biodeterioration &
Biodegradation 48: 233–242.
Garcia-Blanco, S., Venosa, A.D., Suidan, M.T., Lee, K., Cobanli, S. and J.R. Haines. 2007. Biostimulation for the
treatment of an oil-contaminated coastal salt marsh. Biodegradation 18(1): 1-15.
Garrett W.D. and W.R. Barger. 1972. Control and Confinement of Oil Pollution on Water with Monomolecular
Surface Films. Final Report to U.S. Coast Guard, Nov 1971, Project No. 724110.1/4.1 (also reprinted as
U.S. Navy Naval Research Laboratory Memorandum Report 2451, June 1972, AD 744–943).
Genwest Systems, Inc. 2012. EDRC Project Final Report. GSA Contract GS-00F-0002W, BSEE Order #E12-PD-
00012. Prepared for Bureau of Safety and Environmental Enforcement.
http://www.genwest.com/Genwest_EDRC-Project_Final_Report.pdf.
Ghalambor, A. 1997. The Effectiveness of Solidifiers for Combating Oil Spills. Louisiana Applied and Educational
Oil Spill Research and Development Program. Baton Rouge, LA. Technical Report Series, 96-006.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 454
Gong, K., Tkalich, P. and H. Xu. 2014. The numerical investigation on oil slick behavior behind the oil boom.
Journal of Environmental Protection 5(8): 739.
Goodbody-Gringley, G., Wetzel, D. L., Gillon, D., Pulster, E., Miller, A. and K.B. Ritchie. 2013. Toxicity of
Deepwater Horizon source oil and the chemical dispersant, Corexit® 9500, to coral larvae. PLoS One 8(1):
e45574.
Gramss, G., Voigt, K-D. and B. Kirshe, 1999. Oxidoreductase enzymes liberated by plant roots and their effects on
soil humic material. Chemosphere 38(7): 1481-1494.
Gray, J.L., Kanagy, L.K., Furlong, E.T., McCoy, J.W. and C.J. Kanagy. 2011. Determination of the Anionic
Surfactant Di(Ethylhexyl) Sodium Sulfosuccinate in Water Samples Collected from Gulf of Mexico
Coastal Waters Before and After Landfall of Oil From the Deepwater Horizon Oil Spill, May to October,
2010: U.S. Geological Survey Open-File Report 2010–1318. http://pubs.usgs.gov/of/2010/1318/.
Greer, C.W., Fortin, N., Sanschagrin, S., Yergeau, E., Whyte, L.G., King, T.L. and K. Lee. 2015. Natural
degradation potential for crude oil at sub-zero temperatures in the Canadian Arctic marine environment.
Abstract, 38th Arctic and Marine Oilspill Program (AMOP) Technical Seminar. Environment Canada,
Ottawa.
Gülüm, M. and A. Bilgin. 2015. Density, flash point and heating value variations of corn oil biodiesel–diesel fuel
blends. Fuel Processing Technology 134: 456-464.
Hassanshahian, M., Emtiazi, G. and S. Cappello. 2012. Isolation and characterization of crude-oil-degrading bacteria
from the Persian Gulf and the Caspian Sea. Marine Pollution Bulletin 64(1): 7-12.
Hoffman, A. J. and P.D. Jennings. 2011. The BP oil spill as a cultural anomaly? Institutional context, conflict, and
change. Journal of Management Inquiry: DOI: 10.1177/1056492610394940.
Hu, G., Li, J. and G. Zeng. 2013. Recent development in the treatment of oily sludge from petroleum industry: A
review. Journal of Hazardous Wastes 261: 470-490.
IMO. 2005. Manual on Oil Pollution: Section IV – Combating Oil Spills (2nd edition). International Maritime
Organization.
ITOPF (International Tanker Owners Pollution Federation Ltd.). 2014. Use of Dispersants to Treat Oil Spills.
Technical Information Paper #4. London, UK. http://www.itopf.com/knowledge-resources/documents-
guides/document/tip-4-use-of-dispersants-to-treat-oil-spills/.
ITOPF. 2014. Use of Sorbent Materials in Oil Spill Response. Technical Information Paper #8. London.
http://www.itopf.com/knowledge-resources/documents-guides/document/tip-8-use-of-sorbent-materials-in-
oil-spill-response/.
Ivanov, B., Bezgreshnov, A., Kubyshkin, N. and A. Kursheva. 2005. Spreading of oil products in sea ice and their
influence on the radiation properties of the snow-ice cover. In: Proceedings of the 18th International Port
and Oceans Engineering Under Arctic Conditions. Vol. 2. pp. 853-862.
Jacobs, T., Jacobi, M., Rogers, M., Adams, J., Coffey, J., Walker, J. and B. Johnston. 2015. Testing and evaluating
low altitude unmanned aircraft system technology for maritime domain awareness and oil spill response in
the arctic. Marine Technology Society Journal 49(2): 145-150.
Jacques, R.J.S., Okeke, B.C., Bento, F.M., Teixeira, A.S., Peralba, M.C.R. and F.A.O. Camargo. 2008. Microbial
consortium bioaugmentation of polycyclic aromatic hydrocarbons in contaminated soil. Bioresource
Technology 99: 2637-2643.
JRP (Joint Review Panel). 2013. Connections: Report of the Joint Review Panel for the Enbridge Northern Gateway
Project, Vol. 1. National Energy Board, Calgary, AB.. http://gatewaypanel.review-examen.gc.ca/clf-
nsi/dcmnt/rcmndtnsrprt/rcmndtnsrprtvlm1-eng.pdf.
Kleindienst, S., Paul, J. H. and S.B. Joye. 2015. Using dispersants after oil spills: impacts on the composition and
activity of microbial communities. Nature Reviews Microbiology 13(6): 388-396.
Kok, M. V. 2011. Characterization of medium and heavy crude oils using thermal analysis techniques. Fuel
Processing Technology 92(5): 1026-1031.
Kraus, J.J., Munir, I.Z., McEldoon, J.P., Clark, D.S. and J.S. Dordick. 1999. Oxidation of polycyclic aromatic
hydrocarbons catalyzed by soybean peroxidise. Applied Biochemical Biotechnology 80 (3): 221–230.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 455
Lampela, K. and K. Jolma. 2011. Mechanical oil spill recovery in ice: Finnish approach. In: Proceedings of the 2011
International Oil Spill Conference Proceedings (IOSC). American Petroleum Institute. p. abs189.
Leahy, J.G. and R.R. Colwell. 1990. Microbial Degradation of hydrocarbons in the environment. Microbial Reviews
53(3): 305-315.
Lee, K. 2000. In situ bioremediation of oiled shoreline environments. Opportunities for Advancement of
Environmental Applications of Marine Biotechnology. In: Proceedings of the October 5-6, 1999,
Workshop. The National Research Council of the National Academy of Sciences and the National
Academy of Engineering. National Academy Press, Washington, DC. pp. 44-60.
http://www.nap.edu/read/9988/chapter/9.
Lee, K. and E.M. Levy. 1991. Bioremediation: Waxy crude oils stranded on low-energy shorelines. In: Proceedings
of the 1991 Oil Spill Conference, American Petroleum Institute, Washington, D. C. pp. 541-547.
Lee, K., Lunel, T., Wood, P., Swannell, R. and P. Stoffyn-Egli. 1997a. Shoreline clean up by acceleration of clay-oil
flocculation processes. In: Proceedings of the 1997 International Oil Spill Conference (Prevention,
Behaviour, Control and Cleanup. Fort Lauderdale, Florida, April 7-10, 1997. pp. 235-240.
Lee, K., Stoffyn-Egli, P., Tremblay, G.H., Owens, E.H., Sergy, G.A., Guénette, C.C. and R.C. Prince. 2003b. Oil-
mineral aggregate formation on oiled beaches: Natural attenuation and sediment relocation. Spill Science
and Technology Bulletin 8: 285-296.
Lee, K., Nedwed, T., Prince, R.C. and D. Palandro, 2013. Lab tests on the biodegradation of chemically dispersed
oil should consider the rapid dilution that occurs at sea. Marine Pollution Bulletin 73: 314–318.
Li, P., Chen, B., Li, Z.L., Zheng, X., Wu, H.J., Jing, L. and K. Lee. 2014a. A Monte Carlo simulation based two-
stage adaptive resonance theory mapping approach for offshore oil spill vulnerability index classification.
Marine Pollution Bulletin 86(1-2): 434-442.
Li, P., Chen, B., Zhang, B.Y., Jing, L. and J.S. Zheng. 2014b. Monte Carlo simulation-based dynamic mixed integer
nonlinear programming for supporting oil recovery and devices allocation during offshore oil spill
responses. Ocean & Coastal Management 89: 58-70.
Li, Z., Lee, K., King, T., Boufadel, M.C. and A.D. Venosa. 2008. Assessment of chemical dispersant effectiveness
in a wave tank under regular nonbreaking and breaking wave conditions. Marine Pollution Bulletin 56(5):
903–912.
Li, Z., Lee, K., King, T., Boufadel, M.C. and A.D. Venosa. 2009. Evaluating chemical dispersant efficacy in an
experimental wave tank: 2, significant factors determining in-situ oil droplet size distribution.
Environmental Engineering Science 26(9): 1407-1418.
Lunel, T., Lee, K., Swannell, R.,Wood, P., Rusin, J., Bailey, N., Halliwell, C., Davies, L., Sommerville, M., Dobie,
A., Mitchell, D. and M. McDonagh. 1996. Shoreline clean up during the Sea Empress Incident: The role of
surf washing (clay-oil flocculation), dispersants and bioremediation. In: Proceedings of the 19th Arctic and
Marine Oilspill Program (AMOP) Technical Seminar, June 12-14, 1996, Calgary, Alberta, Canada.
Environment Canada. pp. 1521-1540.
Macek, T., Mackova, M. and J. Kas. 2000. Exploitation of plants for the removal of organics in environmental
remediation. Biotechnology Advances 18: 23–34.
Macnaughton, S.J., Stephen, J.R., Venosa, A.D., Davis, G.A. Chang, Y-J. and D.C. White. 1999. Microbial
population changes during bioremediation of an experimental oil spill. Applied Environmental
Microbiology 65: 3566-3574.
Maki, H., Sasaki, T. and S. Harayama. 2001. Photo-oxidation of biodegraded crude oil and toxicity of the photo-
oxidized products. Chemosphere 44(5): 1145-1151.
McGenity, T.J., Folwell, B.D., McKew, B.A. and G.O. Sanni. 2012. Marine crude-oil biodegradation: A central role
for interspecies interactions. Aquatic Biosystems 8:10.
McNutt, M.K., Camilli, R., Crone, T.J., Guthrie, G.D., Hsieh, P.A., Ryerson, T.B., Savas, O. and F. Shaffer. 2012.
Review of flow rate estimates of the Deepwater Horizon oil spill. Proceedings of the National Academy of
Sciences of the United States of America 109 (50): 20260-20267.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 456
Merlin, F.X. and P. Le Guerroué. 2009. Use of Sorbents for Spill Response: Operational Guide. Cedre.
http://www.commissionoceanindien.org/fileadmin/resources/Autoroute%20maritime%20CEDRE/Cedre%2
0Use%20of%20Sorbents%20for%20Spill%20Response%20Cedre.pdf
Michaels, D. and J. Howard. 2012. Review of the OSHA-NIOSH response to the Deepwater Horizon oil spill:
Protecting the health and safety of cleanup workers. PLoS Currents 2012 Jul 18. Edition 1. doi:
10.1371/4fa83b7576b6e.
Michel, J. and B. Benggio. 1999. Guidelines for selecting appropriate cleanup endpoints at oil spills. In:
Proceedings of the 1999 International Oil Spill Conference. American Petroleum Institute. pp. 591-595.
Michel, J., Keane, P. and B. Benggio. 2008. Pre-authorization for the use of solidifiers: results and lessons learned.
In: Proceedings of the 2008 International Oil Spill Conference. American Petroleum Institute, Washington,
D.C. pp. 345-348.
Mills, M.A., Bonner, J.S., Simon, M.A., McDonald, T.J. and R.L. Autenrieth. 1997. Bioremediation of a controlled
oil release in a wetland. In: Proceedings of the 20st Arctic and Marine Oilspill Program (AMOP) Technical
Seminar. Environment Canada, Ottawa. Pp. 606-616.
Mortazavi, B., Horel, A., Beazley, M. J. and P.A. Sobecky. 2013. Intrinsic rates of petroleum hydrocarbon
biodegradation in Gulf of Mexico intertidal sandy sediments and its enhancement by organic substrates.
Journal of Hazardous Materials 244: 537-544.
Muratova, A.Y., Kapitonova, V.V., Chernyshova, M.P. and O.V. Turkovskaya. 2009. Enzymatic Activity of Alfalfa
in a Phenanthrene-contaminated Environment. World Academy of Science, Engineering and Technology
58: 569-574.
Muschenheim, D.K. and K. Lee, 2003. Removal of oil from the sea surface through particulate interactions: Review
and prospectus. Spill Science and Technology Bulletin 8: 9-18.
Myers, D. 2006. Surfactant Science and Technology. Third Edition. John Wiley and Sons, Inc., Hoboken, NJ.
NOAA (National Oceanic & Atmospheric Administration). 1992. Shoreline Countermeasure Manual. Seattle,
Washington.
NOAA. 2007. In-Situ Burning. Office of Response and Restoration. http://response.restoration.noaa.gov/oil-and-
chemical-spills/oil-spills/resources/in-situ-burning.html.
NOAA. 2015. Spill Containment Methods. Office of Response and Restoration.
http://response.restoration.noaa.gov/oil-and-chemical-spills/oil-spills/spill-containment-methods.html
NOAA and API (American Petroleum Institute). 1994. Options for Minimizing Environmental Impacts of
Freshwater Spill Response.
NRC (National Research Council). 1989. Using Oil Spill Dispersants on the Sea. National Academy Press,
Washington, DC.
NRC. 2014. Responding to Oil Spills in the U.S. Arctic Marine Environment. National Academies Press,
Washington, D.C. http://www.nap.edu/catalog/18625/responding-to-oil-spills-in-the-us-arctic-marine-
environment. (Accessed June, 2015)
Nuka Research and Planning Group. 2010. Oil Spill Prevention and Response in the U.S. Arctic: Unexamined Risks,
Unacceptable Consequences. U.S. Arctic Program, Pew Environment Group.
Oskins, C.J. and D. Bradley. 2005. “Extreme" cold weather oil spill response techniques & strategies - ice & snow
environments. In: Proceedings of the 2005 International Oil Spill Conference. American Petroleum
Institute. pp. 521-525.
Pelletier, É. and R. Siron. 1999. Silicone-based polymers as oil spill treatment agents. Environmental Toxicology
and Chemistry 18: 813-818.
Pezeshki, S.R., Hester, M.W., Lin, Q. and J.A. Nyman. 2000. The effects of oil spill and clean-up on dominant US
Gulf coast macrophytes: a review. Environmental Pollution 108: 129-139.
Pilipenko, N., Lerondeau, B. and E. Vial. 2013. Oil spill response in the arctic. In: Proceedings of the 2013 SPE
Arctic and Extreme Environments Technical Conference and Exhibition. Society of Petroleum Engineers.
Porter, M.R. 1991. Handbook of Surfactants. Blackie and Sons, Ltd., Glasgow, Scotland.
Powter, C.B. (Compiler). 2002. Glossary of Reclamation and Remediation Terms Used in Alberta – 7th Edition.
Alberta Environment, Science and Standards Branch, Edmonton, AB. Pub. No. T/655; Report No.
SSB/LM/02-1. http://environment.gov.ab.ca/info/library/6843.pdf

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 457
Prince, R.C., Gramain, A. and T.J. McGenity. 2010. Prokaryotic hydrocarbon degraders. In: Timmis, K.N.,
McGenity, T.J., Meer, J.R. and V. Lorenzo (eds.) Handbook of Hydrocarbon and Lipid Microbiology.
Springer, Berlin Heidelberg. pp. 1671-1692.
Ramsay, M.A., Swannell, R.P.J., Shipton, W.A., Duke, N.C. and R.T. Hill. 2000. Effect of bioremediation on the
microbial community in oiled mangrove sediments. Marine Pollution Bulletin 41: 413-419.
Region 10 Regional Response Team and the Northwest Area Committee. 2015. 9301. Oil spill best management
practices. In: NW Area Contingency Plan. RRT/NWAC, Seattle, WA. http://www.rrt10nwac.com/nwacp/
Rosenberg, E. and E.Z. Ron. 1996. Bioremediation of petroleum contamination. In: Crawford, R.L. and D.L.
Crawford (eds.) Bioremediation: Principles and Applications. Cambridge University Press, UK. Pp. 100-
124.
Salt, D., Hughes, E. and S. Gair. 2012. Monitoring and quantifying oil pollution using aerial surveillance~ problems
and solutions. In: Proceedings of the 2012 SPE International Conference on Health, Safety and
Environment in Oil and Gas Exploration and Production. Society of Petroleum Engineers.
Schmidt, B., Meyer, P. and S. Potter. 2014. Testing of oil recovery skimmers in ice at Ohmsett The National Oil
Spill Response Research & Renewable Energy Test Facility. In Proceedings of the 2014 International Oil
Spill Conference Proceedings. American Petroleum Institute. pp. 618-633.
Schulze, R. and H.L. Hoffman. 1993. World Catalog of Oil Spill Response Products. Fourth Edition. Elkride, MD.
Sergy, G. A., and E.H. Owens. 2007. Guidelines for Selecting Shoreline Treatment Endpoints for Oil Spill
Response. Emergency Science and Technology Division, Environment Canada, Ottawa, ON.
Sergy, G. A. and E.H. Owens. 2008. Selection and use of shoreline treatment endpoints for oil spill response. In:
Proceedings of the 2008 International Oil Spill Conference Proceedings. American Petroleum Institute. pp.
435-441.
Serova, I. 1992. Behaviour of oil in ice and water at low temperatures. In: Combating Marine Oil Spills in Ice and
Cold Climates: HELLCOM Seminar, Helsinki, Finland.
Shin, W.S., Tate, P.T., Jackson, W.A. and J.H. Pardue. 1999. Bioremediation of an experimental oil spill in a salt
marsh. In: Means, J.L. and R.E. Hinchee (eds.) Wetlands and Remediation: An International Conference.
Battelle Press, Columbus, OH. pp. 33-40.
Shin, W.S., Pardue, J.H. and W.A. Jackson. 2000. Oxygen demand and sulfate reduction in petroleum hydrocarbon
contaminated salt marsh soils. Water Research 34(4): 1345-1353.
Sundaravadivelu, D., Suidan, M.T. and A.D. Venosa. 2015. Parametric study to determine the effect of temperature
on oil solidifier performance and the development of a new empirical correlation for predicting
effectiveness. Marine Pollution Bulletin 95(1): 297-304.
Transport Canada. 2013. A Review of Canada’s Ship-Source Oil Spill Preparedness and Response Regime: Setting
the Course for the Future. Tanker Safety Panel Secretariat, Ottawa.
www.tc.gc.ca/eng/tankersafetyexpertpanel/menu.htm. (accessed June 16 2015).
Tyagi, M., da Fonseca, M. M. R. and C.C. de Carvalho. 2011. Bioaugmentation and biostimulation strategies to
improve the effectiveness of bioremediation processes. Biodegradation 22(2): 231-241.
U.S. EPA. 1999. Understanding Oil Spills and Oil Spill Response. EPA 540-K-99-007. Office of Emergency and
Remedial Response, U.S. Environmental Protection Agency.
US Coast Guard. 2001. Characteristics of Response Strategies: A Guide for Spill Response Planning in Marine
Environments. A joint publication of the American Petroleum Institute, National Oceanic and Atmospheric
Administration, U.S. Coast Guard, and U.S. Environmental Protection Agency.
Valentine, D.L., Fisher, G.B., Bagby, S.C., Nelson, R.K., Reddy, C.M., Sylva, S.P. and M.A. Woo. 2014. Fallout
plume of submerged oil from Deepwater Horizon. Proceedings of the National Academy of Sciences (USA)
111(45): 15906-15911.
Venosa, A.D. and X. Zhu. 2005. Guidance for the bioremediation of oil-contaminated wetlands, marshes, and
marine shorelines. In: Fingerman, S. and R. Nagabhushanam (eds.) Bioremediation of Aquatic and
Terrestrial Ecosystems. Science Publishers, Inc. Enfield, NH. pp. 141-172.
Venosa, A.D., Suidan, M.T., Wrenn, B.A., Strohmeier, K.L., Haines, J.R., Eberhart, B.L., King, D. and E. Holder.
1996. Bioremediation of an experimental oil spill on the shoreline of Delaware Bay. Environmental Science
and Technology 30(5): 1764-1775.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 458
Venosa, A.D., Lee, K., Suidan, M.T., Garcia-Blanco, S., Cobanli, S., Moteleb, M., Haines, J.R., Tremblay, G. and
M. Hazelwood. 2002. Bioremediation and biorestoration of crude oil-contaminated freshwater wetland on
the St. Lawrence River. Bioremediation Journal 6(3): 261-281.
Venosa, A.D., Anastas, P.T., Barron, M.G., Conmy, R.N., Greenberg, M.S. and G.J. Wilson. 2012. Science-based
decision making on the use of dispersants in the Deepwater Horizon oil spill: EPA's perspectives. In:
Somasundaran, P., Farinato, R., Patra, P. and K. Papadopoulos (eds.) Oil Spill Remediation: Colloid
Chemistry-Based Principles and Solutions. John Wiley and Sons, Inc.
Verma, M., Gendreau, M. and G. Laporte. 2013. Optimal location and capability of oil-spill response facilities for
the south coast of Newfoundland. Omega 41(5): 856-867.
Wadsworth, T. 1995. Containment & Recovery of Oil Spills at Sea. Methods and Limitations. International Tanker
Owners Pollution Federation Limited. http://www.itopf.com/knowledge-resources/documents-
guides/document/containment-recovery-of-oil-spills-at-sea-methods-and-limitations/).
Walker, A.H., Kucklick, J.H. and J. Michel. 1999. Effectiveness and environmental considerations for non-
dispersant countermeasures. Pure Applied Chemistry 71: 67-81.
Wang, Z. and M. Fingas. 2003. Development of oil hydrocarbon fingerprinting and identification techniques.
Marine Pollution Bulletin 47: 423–452.
Wang, Y. Z., Yan, Q. B., Huang, J. J., Li, B., and J.L. Gao. 2012. Recovery of lubricating oil for cold strip rolling.
Applied Mechanics and Materials 189: 44-47.
Webler, T. and F. Lord. 2010. Planning for the human dimensions of oil spills and spill response. Environmental
Management 45(4): 723–738.
Widada, J., Nojiri, H., Kasuga, K., Yoshida, T.,, Habe, H. and T. Omori. 2002. Molecular detection and diversity of
polycyclic aromatic hydrocarbon-degrading bacteria isolated from geographically diverse sites. Applied
Microbiology and Biotechnology 58: 202–209.
Wilkman, G., Ritari, E., Uuskallio, A. and M. Niini. 2014. Technological Development in Oil Recovery in Ice
Conditions. In: OTC Arctic Technology Conference. Offshore Technology Conference.
Wrenn, B.A., Virkus, A., Mukherjee, B. and A.D. Venosa. 2009. Dispersibility of crude oil in freshwater.
Environmental Pollution 157: 1807-1814.
Yasin, G., Bhanger, M. I., Ansari, T. M., Naqvi, S. M. S. R., Ashraf, M., Ahmad, K. and F.N. Talpur. 2013. Quality
and chemistry of crude oils. Journal of Petroleum Technology and Alternative Fuels 4(3): 53-63.
Zhong, Z., and F. You. 2011. Oil spill response planning with consideration of physicochemical evolution of the oil
slick: A multiobjective optimization approach. Computers & Chemical Engineering 35(8): 1614-1630.
Zhu, X., Venosa, A.D., Suidan M.T. and K. Lee. 2001. Guidelines for the Bioremediation of Marine Shorelines and
Freshwater Wetlands. U.S. Environmental Protection Agency
Zhu, X., Venosa, A.D., Suidan, M.T. and K. Lee. 2004. Guidelines for the Bioremediation of Oil-Contaminated Salt
Marshes. EPA/600/R-04-074. U.S. Environmental Protection Agency.

Chapter 7

Akinyokun, O. C. and U.G. Inyang. 2013. Experimental study of neuro fuzzy-genetic framework for oil spillage risk
management. Artificial Intelligence Research 2(4): 13-35.
ACAP. 2008. The Atlantic Coastal Action Program (ACAP) – Year in Review 2006/07. Atlantic Coastal Action
Program, Environment Canada, Dartmouth, NS.
ADEC. 1993. The Exxon Valdez Oil Spill-Final Report. Department of Environmental Conservation (ADEC) Final
Report. State of Alaska Response.
ADEC. 2015. Oil and Other Hazardous Substances Pollution Control. 18 AAC 75, Department of Environmental
Conservation. Alaska.
Adger, W. N. 2006. Vulnerability. Global Environmental Change 16(3): 268-281.
Adler, E., and M. Inbar. 2007. Shoreline sensitivity to oil spills, the Mediterranean coast of Israel: assessment and
analysis. Ocean & Coastal Management 50(1): 24-34.
AER (Alberta Energy Regulator). 2015. Compliance Dashboard.
http://www1.aer.ca/compliancedashboard/incidents.html.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 459
Aguilera, F., Méndez, J., Pásaro, E. and B. Laffon. 2010. Review on the effects of exposure to spilled oils on human
health. Journal of Applied Toxicology 30(4): 291-301.
Allan, S. E., Smith, B. W. and K.A. Anderson. 2012. Impact of the Deepwater Horizon oil spill on bioavailable
polycyclic aromatic hydrocarbons in Gulf of Mexico coastal waters. Environmental Science & Technology
46(4): 2033-2039.
Allen, B., Stokey, R., Forrester, N., Gouldsborough, R., Purcell, M. and C. von Alt. 1997. Remus: a small, low cost
AUV; system description, field trials and performance results. In: Proceedings of IEEE/MTS Oceans
Conference, October 1997, volume 2. Pp. 994–1000.
Anderson, E. J. and D.J. Schwab. 2012. Contaminant transport and spill reference tables for the St. Clair River.
Marine Technology Society Journal 46(5): 34-47.
Andrade, F., Lutz, G., Lyra, A. P., Ranieri, A., and B. Lemos. 2013. Oil Spill Response Equipment and Techniques
for Waxy Oils in Brazil. In: Proceedings of the SPE Latin-American and Caribbean Heath, Safety,
Environment and Social Responsibility Conference. Society of Petroleum Engineers.
AOSRT JIP, 2014. Environmental Impacts of Arctic Oil Spills and Arctic Spill Response Technologies - Literature
Review and Recommendations. December 2014. Arctic Oil Spill Response Technology Joint Industry
Programme.
API (American Petroleum Institute). 2009. Analysis of US Spillage. API Publication No. 356. American Petroleum
Institute, Washington, D.C.
Baars, B. J. 2002. The wreckage of the oil tanker ‘Erika’—human health risk assessment of beach cleaning,
sunbathing and swimming. Toxicology Letters 128(1): 55-68.
Barenboim, G. and A.J. Saveka. 2012. New methodological aspects of assessment of biological hazard of oil spills.
In: Proceedings of the 35th AMOP Technical Seminar on Environmental Contamination and Response,
Vancouver, Canada. Environment Canada, Ottawa. Pp. 13-30.
Barenboim, G., Saveka, A.J. and O. Avandeeva. 2013. Development of a system for the early detection and
monitoring of oil spills on water bodies with a glance to its use in the Arctic Zone. In: Proceedings of the
36th AMOP Technical Seminar on Environmental Contamination and Response, Halifax, Canada.
Environment Canada, Ottawa.
Barron, M. G. 2012. Ecological impacts of the deepwater horizon oil spill: Implications for immunotoxicity.
Toxicologic Pathology 40(2): 315-320.
Bejarano, A. C. and A.J. Mearns. 2015. Improving environmental assessments by integrating Species Sensitivity
Distributions into environmental modeling: Examples with two hypothetical oil spills. Marine Pollution
Bulletin 93(1): 172-182.
Benoit, L.E. 2007. Rail Transport and the Environment in Canada: A Study Prepared in Support of the 2007
Railway Safety Act Review. Benoit & Associates. Report prepared for Railway Safety Act Review
Advisory Panel.
Board of Ocean Studies. 2014. Reducing Coastal Risk on the East and Gulf Coasts. National Academies Press,
Washington, DC.
Board of Ocean Studies and Marine Board. 2014. Responding to Oil Spills in the US Arctic Marine Environment.
The National Academies Press, Washington, DC.
Boehm, P. D., Page, D. S., Brown, J. S., Neff, J. M., and E. Gundlach. 2014. Long-term fate and persistence of oil
from the Exxon Valdez Oil Spill: Lessons learned or history repeated? In: Proceedings of the 2014
International Oil Spill Conference. American Petroleum Institute. pp. 63-79.
Bostrom, A., Walker, A. H., Scott, T., Pavia, R., Leschine, T. M., and K. Starbird. 2015. Oil spill response risk
judgments, decisions, and mental models: findings from surveying US stakeholders and coastal residents.
Human and Ecological Risk Assessment: An International Journal 21(3): 581-604.
British Columbia Ministry of Environment. 2014. Land Based Spill Preparedness and Response in British
Columbia. Ministry of Environment - Policy Intentions Paper for Consultation. British Columbia.
British Columbia Ministry of Environment. 2015. Queen of the North BC Ferry Sinking.
http://www.env.gov.bc.ca/eemp/incidents/2006/queen_north_06.htm. (accessed August 29, 2015)
Brown, C. and M. Fingas. 2003. Review of the development of laser fluorosensors for oil spill application. Marine
Pollution Bulletin 47: 477–484.
Buist, I., Potter, S., and T. Nedwed. 2011. Herding agents to thicken oil spills in drift ice for in situ burning: new
developments. In: Proceedings of the 2011 International Oil Spill Conference, May 23-26, 2011, Portland,
Oregon, USA. American Petroleum Institute. Pp. abs230.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 460
Bush, B. J. 2011. Addressing the regulatory collapse behind the deepwater horizon oil spill: Implementing a best
available technology regulatory regime for deepwater oil exploration safety and cleanup technology.
Journal of Environmental Law and Litigation 26: 535-570.
Cai, Q.H., Zhang, B.Y., Chen, B., and T. Cao. 2014b. Biodispersant produced by a Rhodococcus
Cai, W.X., Zheng, Y.W., Shi, Y.Q., and H.L. Zhong. 2009. Threat Level Forecast for Ship's Oil Spill - Based on BP
Neural Network Model. In: Proceedings of the International Conference on Computational Intelligence
and Software Engineering, 2009 (CiSE 2009). IEEE.
pp. 1-4.
Canadian National Railway Company (CN). 2015. Gogama, Ontario: Get the Facts.
http://www.cngogamainfo.ca/en/index.asp.
Cao, X and S. Fan. 2013. The synthetic assessment modeling of ships’ oil spill risk based on fuzzy neural network.
In: Proceedings of International Conference on Intelligent Systems and Applications (ISA). Wuhan, China.
2013 May 22-23. Pp. 368-371.
CAPP. 2015. Infrastructure and Transportation: Pipelines. http://www.capp.ca/canadian-oil-and-natural-
gas/infrastructure-and-transportation/pipelines. (accessed August 24, 2015).
CBC. 2005. Province unprepared for Wabamun spill: Klein. CBC News, Aug 11, 2005.
http://www.cbc.ca/news/canada/province-unprepared-for-wabamun-spill-klein-1.542144
Chang, S. E., Stone, J., Demes, K., and M. Piscitelli. 2014. Consequences of oil spills: A review and framework for
informing planning. Ecology and Society 19(2): 26.
Chapin III, F. S., Peterson, G., Berkes, F., Callaghan, T. V., Angelstam, P., Apps, M. and C. Beier. 2004. Resilience
and vulnerability of northern regions to social and environmental change. AMBIO: A Journal of the Human
Environment 33(6): 344-349.
Chapman, H., Purnell, K. and R.J. Law. 2007. The use of chemical dispersants to combat oil spills at sea: A review
of practice and research needs in Europe. Marine Pollution Bulletin 54(7): 827-838.
Chatterjee, S. 2015. Oil spill cleanup: Role of environmental biotechnology. In: Applied Environmental
Biotechnology: Present Scenario and Future Trends. The Springer Press.
Chen, B., Zhang, B.Y., Li, P., Cai, Q.H., Lin, W.Y., and B. Liu. 2012. From Challenges to Opportunities: Towards
Future Strategies and a Decision Support Framework for Oil Spill Preparedness and Response in Offshore
Newfoundland and Labrador. Technical Report for the Harris Centre, St. John’s, Canada.
Christoph, M., Vis, M. A., Rackliff, L. and H. Stipdonk. 2013. A road safety performance indicator for vehicle fleet
compatibility. Accident Analysis & Prevention 60: 396-401.
Christopherson, S. and K. Dave. 2014. A New Era of Crude Oil Transport: Risks and Impacts in the Great Lakes
Basin. Community and Regional Development Institute, Department of Development Sociology, Cornell
University.
https://cardi.cals.cornell.edu/sites/cardi.cals.cornell.edu/files/shared/documents/CardiReports/A-New-Era-
of-Crude-Oil-Transport.pdf.
Chung, P. P., Ballard, J. W. O. and R.V. Hyne. 2012. It's All in the Genes: How Genotype Can Impact Upon
Response to Contaminant Exposure and the Implications for Biomonitoring in Aquatic Systems. INTECH
Open Access Publisher.
Cleveland, C. 2010. The Challenges of Oil Spill Response in the Arctic. National Commission on the BP Deepwater
Horizon Oil Spill and Offshore Drilling, US.
Cluzel, F., Yannou, B., Leroy, Y. and D. Millet. 2012. Proposition of an adapted management process to evolve
from an unsupervised LCA of complex industrial systems towards an eco-designing organisation.
Concurrent Engineering: Research and Applications 20(2): 111-126.
Coelho, G., Clark, J. and D. Aurand. 2013. Toxicity testing of dispersed oil requires adherence to standardized
protocols to assess potential real world effects. Environmental Pollution 177: 185-188.
Coolbaugh, T., DeMicco, E. and E. Kennedy. 2014. Dispersant-related oil spill response communication tools:
Toward an enhanced approach to conveying complex topics in an approachable manner. In: Proceedings of
the 2014 International Oil Spill Conference. American Petroleum Institute. pp. 1163-1171.
Cox, R. 2014. Risk assessment and planning for offshore oil spill response preparedness. In: SPE International
Conference on Health, Safety, and Environment. Society of Petroleum Engineers.
Crosby, S., Fay, R., Groark, C., Kani, A., Smith, J.R., Sullivan, T., and R. Pavia. 2013. Transporting Alberta Oil
Sands Products: Defining the Issues and Assessing the Risks. NOAA Technical Memorandum NOS OR&R
44. NOAA Office of Response and Restoration, Seattle.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 461
Dalbey, B. 2015. Oil Company Agrees to Mackinac Straits Heavy Crude Oil Ban. September 3, 2015, Patch
Network. http://patch.com/michigan/wyandotte/oil-company-agrees-mackinac-straits-heavy-crude-oil-ban-
0. [accessed September 18, 2015]
Daling, P. S., Leirvik, F., Almås, I. K., Brandvik, P. J., Hansen, B. H., Lewis, A. and M. Reed. 2014. Surface
weathering and dispersibility of MC252 crude oil. Marine Pollution Bulletin 87(1): 300-310.
Davies, A. J., and M.J. Hope. 2015. Bayesian inference-based environmental decision support systems for oil spill
response strategy selection. Marine Pollution Bulletin: doi:10.1016/j.marpolbul.2015.05.041.
DeCola, E., Robertson, T., Fletcher, S., and S. Harvey. 2006. Offshore Oil Spill Response in Dynamic Ice
Conditions. A Report to WWF on Considerations for the Sakhalin II Project, Alaska. Nuka Research.
DeMicco, E., Schuler, P. A., Omer, T. and B. Baca. 2011. Net environmental benefit analysis (NEBA) of dispersed
oil on nearshore tropical ecosystems: Tropics–the 25th year research visit. In: Proceedings of the 2011
International Oil Spill Conference. American Petroleum Institute. p. abs282.
Department of Attorney General. 2015. Michigan Petroleum Pipeline Task Force Report. Department of Attorney
General and Department of Environmental Quality, State of Michigan. July 2015. Lansing, USA.
http://www.michigan.gov/documents/deq/M_Petroleum_Pipeline_Report_2015-
10_reducedsize_494297_7.pdf
Dix, M., Kolodziejski, A. and J. Owens. 2011. 19 down, 34 nations to go: AFRICOM and spill response capacity
building. In: Proceedings of the 2011 International Oil Spill Conference. American Petroleum Institute. p.
abs134.
Doerffer, J. W. 2013. Oil Spill Response in the Marine Environment. Elsevier.
Dollhopf, R. and M. Durno. 2011. Kalamazoo River\Enbridge Pipeline Spill 2010. In: Proceedings of the 2011
International Oil Spill Conference. American Petroleum Institute. p. abs422.
Dollhopf, R. H., Fitzpatrick, F. A., Kimble, J. W., Capone, D. M., Graan, T. P., Zelt, R. B., and R. Johnson. 2014.
Response to heavy, non-floating oil spilled in a Great Lakes river environment: a multiple-lines-of-
evidence approach for submerged oil assessment and recovery. In: Proceedings of the 2014 International
Oil Spill Conference Proceedings. American Petroleum Institute. pp. 434-448.
Echavarria-Gregory, M. A., and J.D. Englehardt. 2015. A predictive Bayesian data-derived multi-modal Gaussian
model of sunken oil mass. Environmental Modelling & Software 69: 1-13.
Ecosocialist Network. 2013. Lac-Mégantic: A Social and Ecological Tragedy. The Bullet No. 856.
El-Zahaby A.M., Kabeel A.E., Bakry A.I., and A.M. Khaira. 2011. Effect of disk offset angle on the hydrodynamic
performance of rotating disk skimmers during oil spill recovery. Environmental Engineering Science 28(2):
113-119.
Environment Canada. 1992. Oilspill SCAT Manual for the Coastlines of British Columbia. Environment Canada,
Technology Development Branch, Edmonton, AB, Regional Programme Report 92-03, 245 pp.
Environment Canada. 1999. National Environmental Emergencies Contingency Plan. Environmental Emergencies
Program. Environment Canada. No. de cat. En40-584/1999E.
Environment Canada. 2015a. Canadian Aquatic Biomonitoring Network (CABIN). http://www.ec.gc.ca/rcba-cabin.
(accessed June 21, 2015).
Environment Canada. 2015b. Environmental Effects Monitoring. http://ec.gc.ca/esee-
eem/default.asp?lang=En&n=4B14FBC1-1. (accessed June 21, 2015).
Environment Canada. 2015c. Proceedings of the 38th Arctic and Marine Oilspill Program (AMOP) Technical
Seminar on Environmental Contamination and Response, June 2-4, Vancouver, Canada. Environment
Canada, Ottawa.
ERCB (Energy Resources Conservation Board). 2013. Plains Midstream Canada ULC NPS 20 Pipeline Failure
Licence No. 5592, Line No. 1 April 28, 2011. ERCB Investigation Report.
https://www.aer.ca/documents/reports/IR_20130226-PlainsMidstream.pdf.
Ermakov, S.A. 2013. Oil slicks on the sea surface and their remote sensing. In: Proceedings of the 36th AMOP
Technical Seminar on Environmental Contamination and Response, Halifax, NS. Environment Canada,
Ottawa. Pg, 539-547.
Etkin, D.S. 2012a. Cook Inlet Maritime Risk Assessment: Spill Baseline & Accident Casualty Study. Spill Scenarios
and Impacts. A report prepared for the Glosten Associates.
Etkin, D.S. 2012b. Development of a Quantitative Risk Model for Inland Facility and Pipeline Spills. In:
Proceedings of the 35th AMOP Technical Seminar on Environmental Contamination and Response,
Vancouver, Canada.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 462
Etkin, D.S., Culpepper, D.B. and L. Martinez. 2013. Application of some remote sensing techniques for oil spill
monitoring, response, and planning. In: Proceedings of the 36th AMOP Technical Seminar on
Environmental Contamination and Response, Halifax, Canada. Environment Canada, Ottawa. Pp. 522-538.
Etkin, D.S., Reilly, T. and D. French-McCay. 2011. State-of-the-art risk-based approach to spill contingency
planning and risk management. In: Proceedings of the 34th AMOP Technical Seminar on Environmental
Contamination and Response, Banff, Canada. Environment Canada, Ottawa. Pp. 906-927.
ExxonMobil Corporation. 2013. Offshore Arctic Oil Spill Prevention, Preparedness and Response.
http://www.exxonmobil.com/Corporate/Files/news_pub_2013-arctic-spill-prevent.pdf. (accessed June 15
2015).
Fernandes, R., Neves, R. and C. Viegas. 2013. Integration of an oil and inert spill model in a framework for risk
management of spills at sea: A case study for the Atlantic area. In: Proceedings of the 36th AMOP
Technical Seminar on Environmental Contamination and Response, Halifax, Canada. Environment
Canada, Ottawa. Pp. 326-353.
Ferraro, G., Meyer-Roux, S., Muellenhoff, O., Pavlih, M., Svetak, J., Tarchi, D. and K. Topouzelis. 2009. Long term
monitoring of oil spills in European seas. International Journal of Remote Sensing 30(3): 627-645.
Fingas, M.F. (ed.). 2011. Oil Spill Science and Technology: Prevention, Response, and Cleanup. Elsevier/Gulf
Professional Publication.
Fingas, M.F. 2012. Buoys and Devices for Oil Spill Tracking. In: Proceedings of the 35th AMOP Technical Seminar
on Environmental Contamination and Response, Vancouver, Canada
Fitzpatrick, F., Bejarano, A., Michel, J. and L. Williams. 2012. Net Environmental Benefit Analysis (NEBA)
Relative Risk Ranking Conceptual Design Kalamazoo River System Enbridge Line 6B MP 608 Marshall,
MI Pipeline Release. USGS, Research Planning, Inc., and US Fish and Wildlife Service.
http://www.epa.gov/enbridgespill/ar/enbridge-AR-0963.pdf.
French-McCay, D. P. 2002. Development and application of an oil toxicity and exposure model, OilToxEx.
Environmental Toxicology and Chemistry 21(10): 2080-2094.
Fritt-Rasmussen, J., Wegeberg, S., Boertmann, D., Clausen, D.S., Johansen, K.L., and A. Mosbech. 2013. Net
Environmental Benefit Analysis as a Decision Tool for Greenlandic Oil Spill Response. In: Proceedings of
the 36th AMOP Technical Seminar on Environmental Contamination and Response, Halifax, Canada.
Environment Canada, Ottawa. Pp. 765-776.
Gala, W., Lipton, J., Cernera, P., Ginn, T., Haddad, R., Henning, M., Jahn, K., Landis, W., Mancini, E., Nicoll, J.,
Peters, V., and J. Peterson. 2009. Ecological risk assessment and natural resource damage assessment:
Synthesis of assessment procedures. Integrated Environmental Assessment and Management 5(4): 515-522.
Gill, D.A., Ritchie, L.A., Picou, J.S., Langhinrichsen-Rohling, J., Long, M.A. and J.W. Shenesey. 2014. The Exxon
and BP oil spills: a comparison of psychosocial impacts. Natural Hazards 74(3): 1911-1932.
Gill, S.D., Bonke, C.A., and J. Carter. 1985. Management of the UNIACKE G-72 incident. In: Proceedings of the
1985 International Oil Spill Conference. American Petroleum Institute. pp. 311-314.
Goodman, R. 1994. Overview and future trends in oil spill remote sensing. Spill Science & Technology Bulletin
1(1): 11–21.
Government of Canada. 2010. Northern Canada Vessel Traffic Services Zone Regulations (SOR/2010-127). Annex
of Canada Shipping Act 2001, Ottawa.
Graham, B., Reilly, W. K., Beinecke, F., Boesch, D. F., Garcia, T. D., Murray, C. A. and F. Ulmer. 2011. Deep
Water: The Gulf Oil Disaster and the Future of Offshore Drilling. Report to the President, National
Commission on the BP Deepwater Horizon Oil Spill and Offshore Drilling, Washington, DC.
Grattan, L.M., Roberts, S., Mahan Jr. W.T., McLaughlin, P.K., Otwell, W.S. and J.G. Morris Jr. 2011. The early
psychological impacts of the Deepwater Horizon oil spill on Florida and Alabama communities. Environ
Health Perspectives 119:838-843.
GRCA. 2015. Grand River Conservation Authority. http://www.grandriver.ca. (accessed June 21, 2015).
Griffiths, G., Stevenson, P., Webb, A. T., Millard, N. W., McPhail, S. D., Pebody, M. and J. R. Perrett. 1999. Open
ocean operational experience with the autosub-1 autonomous underwater vehicle. In: Proceedings of the
11th International Symposium on Unmanned Untethered Submersible Technology, Durham, NH, USA.
Hausberger, O., Hoegn, L. A. and K. Soliman. 2012. Management decision matrix for shale gas projects in Europe.
In: SPE Hydrocarbon Economics and Evaluation Symposium. Society of Petroleum Engineers. Pp. 545-556
Hodges, B. R., Orfila, A., Sayol, J. M., and X. Hou. 2015. Operational oil spill modelling: from science to
engineering applications in the presence of uncertainty. In: Ehrhardt, M. (ed.) Mathematical Modelling
and Numerical Simulation of Oil Pollution Problems. Springer International Publishing. pp. 99-126.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 463
Hollnagel, E. and D.D. Woods. 2006. Epilogue: Resilience engineering precepts. In: Hollnagel, E., Woods, D.D. and
N. Leveson (eds.) Resilience Engineering–Concepts and Precepts. Ashgate Publishing Company,
Aldershot. UK. Pp. 347-358.
Holt, T. J., D. R Jones, S. J. Hawkins, and R. G. Hartnoll. 1995. The sensitivity of marine communities to man-
induced change: a scoping report. Contract science report 65. Maes-y-Ffynnon, Penrhosgarnedd, Bangor,
Wales.
Hopkins, A. 2000. Lessons from Longford. The Esso Gas Explosion. CCH Australia limited, Sydney.
Hou, X., and B.R. Hodges. 2014. Visualizing hydrodynamic uncertainty in operational oil spill modeling. In:
Proceedings of the 2014 International Oil Spill Conference. American Petroleum Institute. p. 299013.
Hout, T., Parkin, T., Smith, S., and R. Ward. 2014. Combining best available technology: A systems approach for
effective mechanical oil spill response. In: Proceedings of the 2014 International Oil Spill Conference.
American Petroleum Institute. p. 300184
Howden, S. D., Barrick, D., and H. Aguilar. 2011. Applications of high frequency radar for emergency response in
the coastal ocean: utilization of the Central Gulf of Mexico Ocean Observing System during the Deepwater
Horizon oil spill and vessel tracking. In: Hou, W.W. and R. Armone (eds.) Proceedings of SPIE 8030
Ocean Sensing and Monitoring III, 80300O. International Society for Optics and Photonics.
Huang, G.H., Sae-Lim, N., Liu, L., and Z. Chen. 2001. An interval-parameter fuzzy-stochastic programming
approach for municipal solid waste management and planning. Environmental Modeling and Assessment 6:
271-283.
Ihaksi, T., Kokkonen, T., Helle, I., Jolma, A., Lecklin, T., and S. Kuikka. 2011. Combining conservation value,
vulnerability, and effectiveness of mitigation actions in spatial conservation decisions: An application to
coastal oil spill combating. Environmental Management 47(5): 802-813.
IMO. 1990. International Convention on Oil Pollution Preparedness, Response and Co-operation, 1990, and Final
Act of the Conference. International Maritime Organization. International Legal Materials. Vol. 30, No. 3
(MAY 1991), pp. 733-761. Published by American Society of International Law. 29pp.
Inyang, U. G., and O. C. Akinyokun. 2014. A hybrid knowledge discovery system for oil spillage risks pattern
classification. Artificial Intelligence Research 3(4): 77.
IPIECA (International Petroleum Industry Environmental Conservation Association). 2000. Choosing Spill
Response Options to Minimize Damage Net Environmental Benefit Analysis. London, UK.
IPIECA. 2015. Oil Spill Preparedness and Response: An Introduction. International Association of Oil & Gas
Producers Report 520. International Petroleum Industry Environmental Conservation Association, London.
ITOPF. 2013. Country Profiles: Canada. A Summary of Oil Spill Response Arrangements. International Tanker
Owners Pollution Federation.
http://www.itopf.com/fileadmin/data/Documents/Country_Profiles/canada.pdf.
ITOPF. 2015. Oil Tanker Spill Statistics 2014. http://www.itopf.com/knowledge-resources/data-statistics/statistics/.
(accessed August 24, 2015).
Jennings, K. and F. Schulberg. 2009. Guidance on developing safety performance indicators. Process Safety
Progress 28(4): 362-366.
Jing, L., Chen, B., Zhang, B.Y. and P. Li. 2012. A stochastic simulation-based hybrid interval fuzzy programming
approach for optimizing the treatment of recovered oily water. Journal of Ocean Technology 7(4): 59-72.
Jing, L., Chen, B., Zhang, B.Y. and P. Li. 2015. Process simulation and dynamic control for marine oily wastewater
treatment using UV irradiation. Water Research 81: 101-112.
Johnson, E. I. 2014. Louisiana’s Coastal Stewardship Program: Beach-nesting Bird Protection, Monitoring, and
Community Outreach. Final Report. National Audubon Society, Baton Rouge, LA.
JRP (Joint Review Panel). 2013a. Connections: Report of the Joint Review Panel for the Enbridge Northern
Gateway Project, Vol. 1. National Energy Board, Calgary, AB.. http://gatewaypanel.review-
examen.gc.ca/clf-nsi/dcmnt/rcmndtnsrprt/rcmndtnsrprtvlm1-eng.pdf.
Judson, R. S., Martin, M. T., Reif, D. M., Houck, K. A., Knudsen, T. B., Rotroff, D. M., Xia, M.H., Sakamuru, S.,
Huang, R.L., Shinn, P., Austin, C.P., Kavlock, R.J., and D.J. Dix. 2010. Analysis of eight oil spill
dispersants using rapid, in vitro tests for endocrine and other biological activity. Environmental Science &
Technology 44(15): 5979-5985.
Kang, L., Shi, S., Deng, Z., Jin, J., Zhang, F., and H. Huang. 2013. Simulation, visualization and GIS analysis based
on globe model for PL19-3 oil spill of Bohai Sea. In: Ulrich, M., Civco, D.L., Schulz, K., Ehlers, M. and
K.G. Nikolakopoulos (eds.) Proceedings of SPIE 8893 Earth Resources and Environmental Remote
Sensing/GIS Applications IV, 88931A . International Society for Optics and Photonics.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 464
Keller, A.A. and K. Clark. 2008. Oil recovery with novel skimmer surfaces under cold climate conditions. In:
Proceedings of the 2008 International Oil Spill Conference, May 4-8, Savannah, Georgia. American
Petroleum Institute. Pp. 667-671.
Khan, F.I., Sadiq, R., and T. Husain. 2002. Risk-based process safety assessment and control measures design for
offshore process facilities. Journal of Hazardous Materials 94: 1-36.
Kirby, M. F., and R.J. Law. 2010. Accidental spills at sea–Risk, impact, mitigation and the need for co-ordinated
post-incident monitoring. Marine Pollution Bulletin 60(6): 797-803.
Kulapina, G.M., Markova, O.V., and E.V. Kulapina. 2015. Adapted management techniques for innovative
networks. Review of European Studies 7(6): DOI: 10.5539/res.v7n6p.1
Laborde, D., Wenke, A., Aarnes, Ø., Rudberg, A., and Ø. Endresen. 2015. Platform for oil spill response gap
analysis case Barents and Norwegian Seas. In: Offshore Mediterranean Conference and Exhibition , 25-27
March, Ravenna, Italy. Offshore Mediterranean Conference.
Lacoursière, J. P., Dastous, P. A., and S. Lacoursière. 2015. Lac-Mégantic accident: What we learned. Process
Safety Progress 34(1): 2-15.
Lamarche, A., MacDonald, S., and S. Grenon. 2013. Development of a Google Earth based contingency planning
support system. In: Proceedings of the 36th AMOP Technical Seminar on Environmental Contamination
and Response, Halifax, Canada. Environment Canada, Ottawa. Pp. 809-817.
Lambert, P. 2003. A literature review of portable fluorescence-based oil-in-water monitors. Journal of Hazardous
Materials 102(1): 39-55.
Le Floch, S., Dussauze, M., Merlin, F. X., Claireaux, G., Theron, M., Le Maire, P., and A. Nicolas-Kopec. 2014.
DISCOBIOL: Assessment of the impact of dispersant use for oil spill response in coastal or estuarine areas.
In: Proceedings of the 2014 International Oil Spill Conference. American Petroleum Institute. pp. 491-503.
Leifer, I., Clark, R., Jones, C., Holt, B., Svejkovsky, J., and G. Swayze. 2011. Satellite and airborne oil spill remote
sensing: State of the art and application to the BP Deepwater Horizon oil spill. In: Proceedings of the 34th
AMOP Technical Seminar on Environmental Contamination and Response, Banff, Canada. Environment
Canada, Ottawa. Pp. 270-295.
Leifer, I., Murray, J., Streett, D., Stough, T., Ramirez, E., and S. Gallegos. 2015. The Federal Oil Spill Team for
Emergency Response Remote Sensing, FOSTERRS: Enabling remote sensing technology for marine
disaster response. In: Lippitt, C., Stow, D. and L. Coulter (eds.) Time-Sensitive Remote Sensing. Springer,
New York. pp. 91-111.
Leong, W. F. 2012. Applicability of dynamic risk assessment in the oil and gas industry. In: SPE Middle East
Health, Safety, Security, and Environment Conference and Exhibition. Society of Petroleum Engineers.pp.
70-76.
Leschine, T. M., Pavia, R., Walker, A. H., Bostrom, A., and K. Starbird. 2015. What-if scenario modeling to support
oil spill preparedness and response decision-making. Human and Ecological Risk Assessment: An
International Journal 21(3): 646-666.
Li, P. 2014. Development of an Integrated Decision Support System for Supporting Offshore Oil Spill Response in
Harsh Environments. Doctoral Dissertation, Memorial University of Newfoundland.
Li, P., Chen, B., Li, Z.L., Zheng, X., Wu, H.J., Jing, L., and K. Lee. 2014a. A Monte Carlo simulation based two-
stage adaptive resonance theory mapping approach for offshore oil spill vulnerability index classification.
Marine Pollution Bulletin 86(1-2): 434-442.
Li, P., Chen, B., Zhang, B.Y., Jing, L., and J.S. Zheng. 2012a. A multiple-stage simulation-based mixed integer
nonlinear programming approach for supporting offshore oil spill recovery with weathering processes.
Journal of Ocean Technology 7(4): 87-105.
Li, P., Chen, B., Zhang, B.Y., Jing, L., and J.S. Zheng. 2014b. Monte Carlo simulation-based dynamic mixed integer
nonlinear programming for supporting oil recovery and devices allocation during offshore oil spill
responses. Ocean & Coastal Management 89: 58-70.
Li, P., Chen, B., Li, Z.L. and Jing, L. 2014c. An agent-based simulation-optimization coupling approach for device
allocation and operation control in response to offshore oil spills under uncertainty. Proceedings of the
International Conference on Marine and Freshwater Environments (iMFE 2014), August 6-8, 2014, St.
John's, Canada.
Li, Y., Lan, G. X., Liu, B. X., and D. Chen. 2012b. Dynamic analysis on oil spill in Mexico Bay based on remote
sensing and GIS. Marine Environmental Science 31(1): 79-82.
Liao, Z., Hannam, P.M., Xia, X. and T. Zhao. 2012b. Integration of multi-technology on oil spill emergency
preparedness. Marine Pollution Bulletin 64: 2117–2128.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 465
Lieske, D.J., Mahoney, M., Wilhelm, S.I., Weir, L., and P. O’Hara. 2011. Eyes in the sky: A comparative spatial
analysis of aerial and satellite surveillance of each coast Canadian oil pollution. In: Proceedings of the 34th
AMOP Technical Seminar on Environmental Contamination and Response, Banff, Canada. Environment
Canada, Ottawa. Pp. 244-258.
Lin, G.H., Chen, X.J. and M. Fukushima. 2009. Solving stochastic mathematical programs with equilibrium
constraints via approximation and smoothing implicit programming with penalization. Mathematical
Programming 116: 343-368.
Linkov, I., Satterstrom, F. K., Kiker, G., Batchelor, C., Bridges, T., and E. Ferguson. 2006. From comparative risk
assessment to multi-criteria decision analysis and adaptive management: Recent developments and
applications. Environment International 32(8): 1072-1093.
Liu, A. X., Lang, Y. H., Xue, L. D., Liao, S. L., and H. Zhou. 2009. Probabilistic ecological risk assessment and
source apportionment of polycyclic aromatic hydrocarbons in surface sediments from Yellow Sea. Bulletin
of Environmental Contamination and Toxicology 83(5): 681-687.
MacPherson, J. 2014. Hiland Crude Pipeline Spills Oil Near Alexander, ND. Huffington. 23 March 2014.
http://www.huffingtonpost.com/2014/03/21/hiland-crude-pipeline-spill_n_5008405.html (accessed
September 10, 2015)
Margesin, R. and F. Schinner. 2001. Bioremediation (natural attenuation and biostimulation) of diesel-oil-
contaminated soil in an alpine glacier skiing area. Applied and Environmental Microbiology 67(7): 3127-
3133.
Maritime New Zealand. 2006. New Zealand Marine Oil Spill Response Strategy. Maritime New Zealand, New
Zealand.
Marthiniussen, R. Vestgard, K., Klepaker, R. A. and N. Storkersen. 2004. HUGIN-AUV concept and operational
experiences to date. In: Ocean ’04 - MTS/IEEE Techno-Ocean ’04: Bridges across the Oceans -
Conference Proceedings, volume 2, , Kobe, Japan, Nov. 2004. Pg. 846–850.
Martin, J. D., Adams, J., Hollebone, B., King, T., Brown, R. S., and P.V. Hodson. 2014. Chronic toxicity of heavy
fuel oils to fish embryos using multiple exposure scenarios. Environmental Toxicology and Chemistry
33(3): 677-687.
Martini, N., and R. Patruno. 2005. Oil pollution risk assessment and preparedness in the east Mediterranean. In:
Proceedings of the 2005 International Oil Spill Conference. American Petroleum Institute. pp. 259-264.
Marty, J., and S. Potter. 2014. Risk Assessment for marine spills in Canadian waters, Phase 1: Oil spills south of the
60th Parallel. In: Proceedings of the 37th AMOP Technical Seminar on Environmental Contamination and
Response, Canmore, Canada. Environment Canada, Ottawa. Pp. 537-552.
Matheny, K. 2014. Coast Guard: We can’t adequately respond to Great Lakes heavy oil spill, Detroit Free Press
(September 11, 2014). http://www.freep.com/story/news/local/michigan/2014/09/11/coast-guard-we-cant-
adequately-respond-to-great-lakes-heavy-oil-spill/15422415/ (accessed September 18, 2015
Maykuth, A. 2014. 2,550 barrels of crude recovered from oil spill. Philly.com. http://articles.philly.com/2014-10-
22/business/55284991_1_oil-spill-mid-valley-pipeline-underground-pipeline. (accessed September 10,
2015)
MCA. 2007. The UK SCAT Manual: A Field Guide to the Documentation of Oiled Shorelines in the UK. Maritime
& Coastguard Agency, Southampton, UK.
McGowan, E., Song, L., Marshall, M., and K. River. 2012. The dilbit disaster: Inside the biggest oil spill you’ve
never heard of, Part 1. Inside Climate News, June 26, 2012.
McGrath, C. G. G. 2014. Preventing a nationally-significant oil spill on the Grand Banks of Newfoundland. In:
Proceedings of the 2014 International Oil Spill Conference. American Petroleum Institute. pp. 1226-1238.
Menzie, C.A., Deardorff, T., Booth, P., and T. Wickwire. 2012. Refocusing on nature: Holistic assessment of
ecosystem services. Integrated Environmental Assessment and Management 8(3): 401-411.
Merchant, R. M., Elmer, S., and N. Lurie. 2011. Integrating social media into emergency-preparedness efforts. New
England Journal of Medicine 365(4): 289-291.
Metzger, M. J., Rounsevell, M. D. A., Acosta-Michlik L., Leemans, R. and D. Schröter. 2006. The vulnerability of
ecosystem services to land use change. Agriculture, Ecosystems & Environment 114(1): 69-85.
Michigan Department of Community Health. 2014. Public Health Assessment: Kalamazoo River/Enbridge Spill:
Evaluation of Kalamazoo River Surface Water and Fish after a Crude Oil Release. Michigan Department of
Community Health, Lansing.
Minister of Justice. 2001. Canada Shipping Act, 2001 (2001, c. 26). Published by the Minister of Justice. Ottawa,
Canada

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 466
Morović, M., and A. Ivanov. 2011. Oil Spill Monitoring in the Croatian Adriatic Waters: needs and possibilities.
Acta Adriat 52(1): 45-56.
Morris, B. F. and T.R. Loughlin. 1994. Overview of the Exxon Valdez oil spill, 1989-1992. In: Loughlin, T.R. (ed.)
Marine Mammals and the Exxon Valdez. Academic Press.
NBC. 2014. Israel Oil Spill: 600,000-Gallons Lost in Eilat Pipeline Breach. NBC News Dec 4 2014.
http://www.nbcnews.com/news/world/israel-oil-spill-600-000-gallons-lost-eilat-pipeline-breach-n261386.
(accessed September 7, 2015)
NEB (National Energy Board). 2014a. Order XO-E101-003-2014. Orders to Enbridge Pipelines Inc. – Application
for the Line 9B Revesrsal and Line 9 Capacity Expansion Project (A59174). National Energy Board.
NEB. 2014b. Pipeline Performance Measures and Integrity Inspection Information Reporting Form. National Energy
Board. http://www.neb-one.gc.ca/sftnvrnmnt/sft/pplnprfrmncmsr/index-eng.html.
NEB. 2015a. Safety. National Energy Board. http://www.neb-one.gc.ca/sftnvrnmnt/sft/index-eng.html. (accessed
August 20, 2015).
NEB. 2015b. Enbridge Pipelines Inc. - Line 9B Reversal and Line 9 Capacity Expansion Project (OH-002-2013).
National Energy Board. http://www.neb-one.gc.ca/pplctnflng/mjrpp/ln9brvrsl/index-eng.html. (accessed
August 20, 2015
NEB. 2015c. Safety and Environmental Performance Dashboard. National Energy Board Updated July 31, 2015.
http://www.neb-one.gc.ca/sftnvrnmnt/sft/dshbrd/index-eng.html. (accessed August 2015)
Newbrey, J. L., Paszkowski, C. A., and B.A. Gingras. 2012. Trophic relationships of two species of grebe on a
prairie lake based on stable isotope analysis. Hydrobiologia 697(1): 73-84.
Nicholls, K., Picou, J. S., and J. Lowman. 2014. Enhancing the utility of community health workers in disaster
preparedness, resiliency, and recovery. In: Proceedings of the 2014 International Oil Spill Conference.
American Petroleum Institute. pp. 170-183.
NRC. 2013. An Ecosystem Services Approach to Assessing the Impacts of the Deepwater Horizon Oil Spill in the
Gulf of Mexico. Report by the Committee on the Effects of the Deepwater Horizon Mississippi Canyon-
252 Oil Spill on Ecosystem Services in the Gulf of Mexico. Ocean Studies Board, Division on Earth and
Life Studies, National Research Council. The National Academies Press, Washington, D.C.
TSB. 2015. Statistical Summary Railway Occurrences 2014. Transportation Safety Board of Canada. Catalogue No.
TU1-2E-PDF. June 2015.
Nuka Research and Planning Group. 2010. Oil Spill Prevention and Response in the US Arctic: Unexamined Risks,
Unacceptable Consequences. US Arctic Program, Pew Environment Group.
NYSDEC. 2015. Biomonitoring. Stream Biomonitoring Unit, Division of Water. New York State Department of
Environmental Conservation (NYSDEC), Troy, NY. http://www.dec.ny.gov/chemical/23847.html.
(accessed June 20, 2015)
OGP (International Association of Oil & Gas Producers). 2011. Process Safety-Recommended Practice on Key
Performance Indicators.
OGP. 2011. Process safety - recommended practice on key performance indicators. International Association of Oil
& Gas Producers. Report No. 456.
Øien, K. 2010. Remote operation in environmentally sensitive areas: development of early warning indicators. In:
Proceedings of the 2nd iNTeg-Risk Conference: New Technologies and Emerging Risks - Dealing with
multiple and interconnected emerging risks, Stuttgart, Germany. Steinbeis Editon.
Øien, K., Massaiu, S., Tinmannsvik, R. K., and F. Størseth. 2010. Development of early warning indicators based on
Resilience Engineering. In: PSAM10, International Probabilistic Safety Assessment and Management
Conference. pp. 7-11.
Olita, A., Cucco, A., Simeone, S., Ribotti, A., Fazioli, L., Sorgente, B., and R. Sorgente. 2012. Oil spill hazard and
risk assessment for the shorelines of a Mediterranean coastal archipelago. Ocean & Coastal Management
57: 44-52.
Ornitz, B. and M. Champ. 2003. Oil Spills First Principles: Prevention and Best Response. Elsevier, Netherlands.
Oskins, C.J., and D. Bradley. 2005. "Extreme" cold weather oil spill response techniques & strategies - ice & snow
environments. In: Proceedings of the 2005 International Oil Spill Conference. American Petroleum
Institute. pp. 521-525.
Owens, E. H. and P.D. Reimer. 2013. Surveying oil on the shoreline. In: Wiens, J.A. (ed.) Oil in the Environment:
Legacies and Lessons of the Exxon Valdez Oil Spill. Cambridge University Press. Pp. 78-97.
Owens, E.H. and G.A. Sergy. 1994. Field Guide to the Documentation and Description of Oiled Shorelines.
Environment Canada, Edmonton, AB, 66 pp.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 467
Owens, E.H. and G.A. Sergy. 2000. The SCAT Manual: A Field Guide to the Documentation and Description of
Oiled Shorelines. Second Edition. Environment Canada, Edmonton, AB.
Owens, E.H. and G.A. Sergy. 2004. The Arctic SCAT Manual: A Field Guide to the Documentation of Oiled
Shorelines in Arctic Environments. Environment Canada, Edmonton, AB.
Pacific Streamkeepers Federation. 2003. The Streamkeepers Program.
http://www.pskf.ca/program/program.html.(accessed June 21, 2015)
Paltrinieri, N., Cozzani, V., Øien, K., and T.O. Grøtan. 2012. Prevention of atypical accident scenarios through the
use of resilience based early warning indicators. In: Berenger, C., Grall, A. and C. Guedes Soares (eds.)
Advances in Safety, Reliability and Risk Management. Taylor & Francis, London.
Pan, G., Tang, D., and Y. Zhang. 2012. Satellite monitoring of phytoplankton in the East Mediterranean Sea after
the 2006 Lebanon oil spill. International Journal of Remote Sensing 33(23): 7482-7490.
Panzar, J., Reyes, E.A. and J. Mozingo. 2015. Santa Barbara County oil cleanup continues; pipeline may be dug up
soon. Los Angeles Times May 23, 201.5
Paul, J. H., Hollander, D., Coble, P., Daly, K. L., Murasko, S., English, D., Basso, J., Delaney, J., McDaniel, L., and
C.W. Kovach. 2013. Toxicity and mutagenicity of Gulf of Mexico waters during and after the Deepwater
Horizon oil spill. Environmental Science & Technology 47(17): 9651-9659.
Payne, K. C., Jackson, C. D., Aizpurua, C. E., Rojas, O. J., and M.A. Hubbe. 2012. Oil spills abatement: factors
affecting oil uptake by cellulosic fibers. Environmental Science & Technology 46(14): 7725-7730.
Peltier, H., Jepson, P. D., Dabin, W., Deaville, R., Daniel, P., Van Canneyt, O., and V. Ridoux. 2014. The
contribution of stranding data to monitoring and conservation strategies for cetaceans: Developing spatially
explicit mortality indicators for common dolphins (Delphinus delphis) in the eastern North-Atlantic.
Ecological Indicators 39: 203-214.
Pilipenko, N., Lerondeau, B., and E. Vial. 2013. Oil spill response in the arctic. In: Proceedings of the 2013 SPE
Arctic and Extreme Environments Technical Conference and Exhibition. Society of Petroleum Engineers.
Pisano, A. 2011. Development of Oil Spill Detection Techniques for Satellite Optical Sensors and their Application
to Monitor Oil Spill Discharge in the Mediterranean Sea. PhD Thesis. University of Bologna.
Potter, S., Buist, I., Trudel, K, Dickins, D. and E. Owens. 2012. Spill Response in the Arctic Offshore. The
American Petroleum Institute and the Joint Industry Programme on Oil Spill Recovery in Ice.
Prendergast, D. P., and P.M. Gschwend. 2014. Assessing the performance and cost of oil spill remediation
technologies. Journal of Cleaner Production 78: 233-242.
Prince, R. C. 2015. Oil spill dispersants: boon or bane? Environmental Science & Technology 49(11): 6376-6384.
Pula, R., Khan, F.I., Veitch, B., and P.R. Amyotte. 2005. Revised fire consequence models for offshore quantitative
risk assessment. Journal of Loss Prevention in the Process Industries 18: 443-454.
Radović, J. R., Aeppli, C., Nelson, R. K., Jimenez, N., Reddy, C. M., Bayona, J. M., and J. Albaigés. 2014.
Assessment of photochemical processes in marine oil spill fingerprinting. Marine Pollution Bulletin 79(1):
268-277.
Ramseur, J., Lattanzio, R., Luther, L., Parfomak, P., and N. Carter. 2014. Oil Sands and the Keystone XL Pipeline:
Background and Selected Environmental Issues. Congressional Research Service, Washington, DC.
Reich, D., Balouskus, R., McCay, D.F., Etkin, D.S., Michel, J., and J. Lehto. 2014. An environmental vulnerability
model for oil spill risk analyses: Examples from an assessment for the state of Alaska. In: Proceedings of
the 37th AMOP Technical Seminar on Environmental Contamination and Response, Canmore, Canada.
Pp. 65-88.
Reilly, T.I. and R.K. York. 2001. Guidance on Sensory Testing and Monitoring of Seafood for Presence of
Petroleum Taint Following an Oil Spill. NOAA Technical Memorandum NOS OR&R 9, Seattle, WA.
Reynolds, K. 2014. Applying international standards in response to oil spills in remote areas. In: Proceedings of the
2014 International Oil Spill Conference Proceedings. American Petroleum Institute. pp. 1859-1868.
Rise, I. H. 2014. The Agreement on Cooperation on Marine Oil Pollution Preparedness and Response in the Arctic.
Master Thesis. The Arctic University of Norway.
Rost, G., Londong, J., Dietze, S., and G. Osor. 2014. Integrated urban water management–development of an
adapted management approach Darkhan, Kharaa catchment, Mongolia. Environmental Earth Science 73:
709–718.
Salt, D., Hughes, E., and S. Gair. 2012. Monitoring and quantifying oil pollution using aerial surveillance-problems
and solutions. In: Proceedings of the 2012 SPE International Conference on Health, Safety and
Environment in Oil and Gas Exploration and Production. Society of Petroleum Engineers.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 468
Sanyal, J., Zhang, S., Dyer, J., Mercer, A., Amburn, P., and R.J. Moorhead. 2010. Noodles: A tool for visualization
of numerical weather model ensemble uncertainty. IEEE Transactions on Visualization and Computer
Graphics 16(6): 1421-1430.
Shupe, D., Ott, G., and K. Preble. 2014. Evaluating place of refuge risk in the Chesapeake Bay. In: Proceedings of
the 2014 International Oil Spill Conference Proceedings. American Petroleum Institute. pp. 814-824.
Silliman., B. D. 2014. Guidelines to Prepare for Oil Sands Product Spills in Varied Aquatic Environments. In:
Proceedings of the 2014 International Oil Spill Conference. American Petroleum Institute. pp. 426-433.
Skogdalen, J. E., Utne, I. B., and J.E. Vinnem. 2011. Developing safety indicators for preventing offshore oil and
gas deepwater drilling blowouts. Safety Science 49(8): 1187-1199.
Song, K. C., Hu, S. P., Yan, Y. H., and J. Li. 2014. Surface defect detection method using saliency linear scanning
morphology for silicon steel strip under oil pollution interference. ISIJ International 54(11): 2598-2607.
Soomere, T. 2012. A preventative method for minimizing environmental risks based on the optimization of the
location of potentially dangerous activities. In: Proceedings of the 35th AMOP Technical Seminar on
Environmental Contamination and Response, Vancouver, Canada. Pp. 836-881.
Stephenson, M., Mazzocco, P., and A. St-Amand. 2015. Ecological risk assessment of hypothetical spills of diluted
bitumen in rivers. In: Proceedings of the 38th AMOP Technical Seminar on Environmental Contamination
and Response, Vancouver, Canada.
Struzik, E. 2015. Oil drilling in Arctic Ocean: A push into uncharted waters. Yale Environment 360, June 8, 2015.
Sweeten, D.W., Wittman, J., Le Roux, P., and T. Bleier. 2012. Improving oil spill response decision making using a
rapidly deployed aerial imaging platform. In: Proceedings of the 35th AMOP Technical Seminar on
Environmental Contamination and Response, Vancouver, Canada. Pp. 658-666.
Tan, S. H. 2011. In-situ fluorometry and SMART Protocol – The Montara Wellhead Experience. Proceedings of the
2011 International Oil Spill Conference. American Petroleum Institute. P. Abs141.
Taylor, E., Moyano, M. and A. Steen. 2014. Upgraded RETOS™: An International Tool to Assess Oil Spill
Response Planning and Readiness. In: Proceedings of the 2014 International Oil Spill Conference.
American Petroleum Institute. pp. 1353-1363.
The Canadian Press. 2015a. Alberta pipeline spill: Too soon to know what caused Nexen leak, says regulator.
HuffPost Alberta July 17, 2015. http://www.huffingtonpost.ca/2015/07/17/nexen-pipeline-spills-
13_n_7819438.html.
The Canadian Press. 2015b. Regulator orders Nexen Energy to suspend operation of 95 pipelines in Alberta. CTV
News August 28, 2015. http://www.ctvnews.ca/business/regulator-orders-nexen-energy-to-suspend-
operation-of-95-pipelines-in-alberta-1.2538815.
Transport Canada. 2013. A Review of Canada’s Ship-Source Oil Spill Preparedness and Response Regime: Setting
the Course for the Future. Tanker Safety Panel Secretariat, Ottawa.
www.tc.gc.ca/eng/tankersafetyexpertpanel/menu.htm. (accessed June 16 2015).
Transport Canada. 2014. Marine Oil Spill Preparedness and Response Regime. Report to Parliament 2006 – 2011,
TP 14539E, Transport Canada, Ottawa, ON.
Transport Canada 2015a. FAQs - Rail Safety and the Transportation of Dangerous Goods by rail.
http://www.tc.gc.ca/eng/mediaroom/faq-rail-safety-tsb-7565.html
Transport Canada. 2015b. Joint Industry-Government Guidelines for the Control of Oil Tankers and Bulk Chemical
Carriers in Ice Control Zones of Eastern Canada. TP 15163, Transport Canada, Ottawa, ON.
Transport Canada. 2015c. Motor Carriers, Commercial Vehicles and Drivers.
https://www.tc.gc.ca/eng/motorvehiclesafety/roadsafety-motor-carriers-commercial-vehicles-drivers-
1312.htm
Transport Canada. 2015d. Risk Assessment for Marine Spills in Canadian Waters.
http://www.tc.gc.ca/eng/marinesafety/menu-4100.htm#abs. (accessed June 24, 2015)
Transport Canada. 2015e. Tanker Safety and Spill Prevention. http://www.tc.gc.ca/eng/marinesafety/menu-
4100.htm. (accessed August 24, 2015)
Turner, M. 2010. Review of Offshore Oil-Spill Prevention and Remediation Requirements and Practices in
Newfoundland and Labrador. Department of Natural Resources, Government of Newfoundland and
Labrador, Canada.
Tyler-Walters, H., and A. Jackson. 1999. Assessing seabed species and ecosystems sensitivities; rationale and user
guide. Report to English Nature, Scottish Natural Heritage and the Department of the Environment
Transport, and the Regions from the Marine Life Information Network (MarLIN report 4). Marine
Biological Association of the United Kingdom, Plymouth.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 469
US Chemical Safety and Hazard Investigation Board. 2007. Investigation Report, Refinery Explosion and Fire (15
Killed, 180 Injured), BP Texas City, Texas, March 23, 2005. Report No. 2005-04-I-TX, March 2007.
US Coast Guard. 2001. Characteristics of Response Strategies: A Guide for Spill Response Planning in Marine
Environments. A joint publication of the American Petroleum Institute, National Oceanic and Atmospheric
Administration, US Coast Guard, and US Environmental Protection Agency.
US EPA. 1999. Screening Level Ecological Risk Assessment Protocol (SLERAP) for Hazardous Waste Combustion
Facilities. EPA 530-D-99-001A, US Environmental Protection Agency.
US EPA. 2004. Generic Ecological Assessment Endpoints (GEAE) for Ecological Risk Assessment. EPA/630/P-
02/004F, US Environmental Protection Agency, Risk Assessment Forum, Washington, DC.
US GAO (US Government Accountability Office). 2012. Oil and Gas: Interior Has Strengthened its Oversight of
Subsea Well Containment, but Should Improve its Documentation. Report to Congressional Requesters.
GAO-12-244. http://www.gao.gov/assets/590/ 588961.pdf. (accessed June 15 2015).
US NOAA (National Oceanic and Atmospheric Association). 1990. Excavation and rock washing treatment
technology: Net environmental benefit analysis. Hazardous Materials Response Branch, National Oceanic
and Atmospheric Administration, Seattle, Washington.
US NOAA. 1994. Alaska Shoreline Countermeasures Manual. Hazardous Materials Response & Assessment
Division, National Oceanic and Atmospheric Administration. 99pp
US NOAA. 2013. Shoreline Assessment Manual, 4th Edition. US Dept. of Commerce, Seattle, WA: Emergency
Response Division, Office of Response and Restoration, National Oceanic and Atmospheric
Administration. 73 pp + appendices.
von Stackelberg, K.E. 2013. Decision analytic strategies for integrating ecosystem services and risk assessment.
Integrated Environmental Assessment and Management 9(2): 260-268.
Wang, C., Chen, B., Zhang, B., Guo, P. and M. Zhao. 2014. Study of weathering effects on the distribution of
aromatic steroid hydrocarbons in crude oils and oil residues. Environmental Science: Processes Impacts
(Formerly Journal of Environmental Monitoring) 16(10): 2408-2414.
Wang, C.Y., Chen, B., Zhang, B., He, S. and M. Zhao. 2013. Fingerprint and weathering characteristics of crude oils
after Dalian oil spill, China. Marine Pollution Bulletin 71(1): 64-68
Wang , Y. Z., Yan, Q. B., Huang, J. J., Li, B., and J.L. Gao. 2012. Recovery of lubricating oil for cold strip rolling.
Applied Mechanics and Materials 189: 44-47.
Wang , Z., Li, K., Fingas, M., Sigouin, L. and L. Ménard. 2002. Characterization and source identification of
hydrocarbons in water samples using multiple analytical techniques. Journal of Chromatography A 971:
173-184.
Wang, Z., Hollebone, B. P., Fingas, M., Fieldhouse, B., Sigouin, L., Landriault, M., Smith, P., Noonan, J., Thouin,
G. and J.W. Weaver. 2003. Characteristics of Spilled Oils, Fuels, and Petroleum Products: 1. Composition
and Properties of Selected Oils. National Exposure Research Laboratory, Office of Research and
Development, United States Environmental Protection Agency, Research Triangle Park, North Carolina.
Wang, Z., Yang, C., Yang, Z., Sun, J., Hollebone, B., Brown, C., and M. Landriault. 2011. Forensic fingerprinting
and source identification of the 2009 Sarnia (Ontario) oil spill. Journal of Environmental Monitoring
13(11): 3004-3017.
Wang, Z.D. and S. Stout. 2006. Oil Spill Environmental Forensics: Fingerprinting and Source Identification, 1st
Edition. Academic Press.
Wei, A., Jianwei, L., Yupeng, Z., Haibo, C., Yonggang, W., and S. Tianyun. 2015. R&D of underwear oil spill
numerical simulation and 3D visualization system in deepwater area. Aquatic Procedia 3: 165-172.
White, H. K., Hsing, P. Y., Cho, W., Shank, T. M., Cordes, E. E., Quattrini, A. M., Nelson, R.K., Camilli, R.,
Demopoulos, A.W.J., German, C.R., Brooks, J.M., Roberts, H.H., Shedd, W., Reddy, C.M., and C.R.
Fisher. 2012. Impact of the Deepwater Horizon oil spill on a deep-water coral community in the Gulf of
Mexico. Proceedings of the National Academy of Sciences 109(50): 20303-20308.
WHOI. 2010. WHOI Scientists Map and Confirm Origin of Large, Underwater Hydrocarbon Plume in Gulf. Media
Relations Office, August 19, 2010. Woods Hole Oceanographic Institute.
http://www.whoi.edu/page.do?pid=7545&tid=282&cid=79926&. (accessed June 20, 2015)
WSP. 2014a. Risk Assessment for Marine Spills in Canadian Waters. Phase 1: Oil Spills South of 60th Parallel. Final
Study Report for Transport Canada. 131-17593-00.
WSP. 2014b. Risk Assessment for Marine Spills in Canadian Waters. Phase 2, Part B: Spills of Oil and Select
Hazardous and Noxious Substances (HNS) Transported in Bulk North of the 60th Parallel North. Final
Study Report for Transport Canada. 131-17593-00.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 470
Yan H. 2015. After oil spilled in Yellowstone River, residents told not to drink water. CNN. January 21, 2015.
http://edition.cnn.com/2015/01/20/us/yellowstone-river-spill. (accessed September 7, 2015)
Yang, Z., Hollebone, B. P., Wang, Z., Yang, C., Brown, C., and M. Landriault. 2013a. Forensic identification of
spilled biodiesel and its blends with petroleum oil based on fingerprinting information. Journal of
Separation Science 36(11): 1788-1796.
Yang, C., Wang, Z., Liu, Y., Yang, Z., Li, Y., Shah, K., and S. Tian. 2013b. Aromatic steroids in crude oils and
petroleum products and their applications in forensic oil spill identification. Environmental Forensics
14(4): 278-293.
Yee, D., Dion, F-O., Smith, C., Videla, P., and T. Rodolakis. 2015. Human health risk assessment: acute and chronic
biological effects from hypothetical releases of diluted bitumen into sensitive watercourses. In:
Proceedings of the 38h AMOP Technical Seminar on Environmental Contamination and Response,
Vancouver, Canada.
Yim, U. H., Ha, S. Y., An, J. G., Won, J. H., Han, G. M., Hong, S. H., Kim, M., Jung, J. H. and W.J. Shim. 2011.
Fingerprint and weathering characteristics of stranded oils after the Hebei Spirit oil spill. Journal of
Hazardous Materials 197: 60-69.
Yim, U. H., Kim, M., Ha, S. Y., Kim, S., and W.J. Shim. 2012. Oil spill environmental forensics: the Hebei Spirit
oil spill case. Environmental Science & Technology 46(12): 6431-6437.
Ying, L., Li, J., and Z. Qingling. 2011. Microwave remote sensing sea surfaces covered in oil. In: 2011 International
Conference on Electric Information and Control Engineering (ICEICE). IEEE. pp. 2319-2322.
You, F., and S. Leyffer. 2011. Mixed‐integer dynamic optimization for oil-spill response planning with integration
of a dynamic oil weathering model. AIChE Journal 57(12): 3555-3564.
Yu, F., and Y. Yin. 2011. Simulation and 3D visualization of oil spill on the sea. In: 2011 IEEE International
Symposium on VR Innovation (ISVRI). IEEE. pp. 213-216.
Zacharias, M. A. and E.J. Gregr. 2005. Sensitivity and vulnerability in marine environments: an approach to
identifying vulnerable marine areas. Conservation Biology 19(1): 86-97.
Zhong, Z., and F. You. 2011. Oil spill response planning with consideration of physicochemical evolution of the oil
slick: A multiobjective optimization approach. Computers & Chemical Engineering 35(8): 1614-1630.
Zhu, L., Wu, Y. Q., and J. Yin. 2013. Segmentation of marine spill oil SAR image based on Gabor, Krawtchouk
Moments and KFCM. Advanced Materials Research 760: 1462-1466.

Chapter 8

AER (Alberta Energy Regulator). 2015a. News Release. AER responding to a pipeline failure near Fort McMurray.
Alberta Energy Regulator, Calgary, AB. https://www.aer.ca/documents/news-releases/AERNR2015-08.pdf
AER. 2015b. News Release. AER issues Suspension Order to Nexen. Alberta Energy Regulator, Calgary, AB.
https://www.aer.ca/documents/news-releases/AERNR2015-13.pdf
Alpine Environmental and EBA Engineering. 2001. Interim Report on the Effects of the Pembina Oil Spill on Fish
Inhabiting the Pine River System, Northeastern British Columbia. Submitted to Pembina Pipeline
Corporation, Calgary, AB. 260 pp.
Amec. 2001a. Environmental Assessment of the Pembina Pipeline Oil Spill to the Pine River near Chetwynd, BC.
Volume 2 – Water Quality. Submitted to District of Chetwynd. 183 pp.
Amec. 2001b. Environmental Assessment of the Pembina Pipeline Oil Spill to the Pine River near Chetwynd, BC.
Volume 1 – Sediment Quality. Submitted to District of Chetwynd. 104 pp.
Amec. 2001c. Environmental Assessment of the Pembina Pipeline Oil Spill to the Pine River near Chetwynd, BC.
Volume 3 – Fish Surveys. Submitted to District of Chetwynd. 142 pp.
Anderson, A-M. 2005. Wabamun Lake Oil Spill August 2005: Data Report for Water and Sediment Quality in the
Pelagic Area of the Lake (August 4-5 to September 15, 2005). Alberta Environment, Edmonton, AB. 99 pp.
Anderson, M.E. 2014. Early review of potential impacts of the Deepwater Horizon oil spill on Gulf of Mexico
wetlands and their associated fisheries. In: Alford, J. B., Peterson, M. S. and C. C. Green (eds.) Impacts of
Oil Spill Disasters on Marine Habitats and Fisheries in North America. CRC Press, Boca Raton, FL. Pp
97-112.
AOSRT 2014. Environmental Impacts of Arctic Oil Spills and Arctic Spill Response Technologies - Literature
Review and Recommendations. December 2014. Arctic Oil Spill Response Technology Joint Industry
Programme.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 471
http://neba.iogp.org/assets/files/Environmental%20Impacts%20of%20Arctic%20Oil%20Spills%20-
%20report.pdf
Atlas, R. and T. Hazen. 2011. Oil biodegradation and bioremediation: A tale of the two worst spills in U.S. history.
Environmental Science & Technology 45: 6709–6715.
Baccante, N. 2000. Preliminary report on fish mortalities in the Pine River, near Chetwynd, B.C., following an oil
spill from Pembina’s pipeline on August 1, 2000. BC Ministry of Environment, Fort St. John, BC. 20 pp.
Ballachey, B.E., Bodkin, J.L., Esler, D. and S.D. Ricer. 2014. Lessons from the 1989 Exxon Valdez oil spill: A
biological perspective. In: Alford, J. B., Peterson, M. S. and C. C. Green (eds.) Impacts of Oil Spill
Disasters on Marine Habitats and Fisheries in North America. CRC Press, Boca Raton, FL. Pp. 181-197.
Bejarano, A.C., Michel, J., and Williams, L. 2012. Net Environmental Benefit Analysis (NEBA) Relative Risk
Ranking Conceptual Design, Kalamazoo River System, Enbridge Line 6B Release. August 8, 2012;
document and appendixes; AR-0963. Cited in Fitzpatrick, F.A., Boufadel, M.C., Johnson, R., Lee, K.,
Graan, T.P., Bejarano, A.C., Zhu, Z., Waterman, D., Capone, D.M., Hayter, E., Hamilton, S.K., Dekker, T.,
Garcia, M.H. and J.S. Hassan. 2015. Oil-Particle Interactions and Submergence from Crude Oil Spills in
Marine and Freshwater Environments – Review of the Science and Future Science Needs. U.S. Geological
Survey. Open File Report 2015-1076.
Biber, P. D., Wu, W., Peterson, M.S., Liu, Z. and L. Pham. 2014. Oil contamination in Mississippi saltmarsh
habitats and the impacts to Spartina alterniflora photosynthesis. In: Alford, J. B., Peterson, M. S. and C. C.
Green (eds.) Impacts of Oil Spill Disasters on Marine Habitats and Fisheries in North America. CRC
Press, Boca Raton, FL. Pp. 133-171.
Birtwell, I.K. 2003. Comments on the Effects of Crude Oil on Fish and Their Habitat in the Pine River, British
Columbia. Fisheries and Oceans Canada, West Vancouver, BC. 25 pp.
Birtwell, I.K. 2008b. Comments on the Effects of Oil Spillage on Fish and Their Habitat – Lake Wabamun, Alberta.
Fisheries and Oceans Canada, West Vancouver, BC. 198 pp.
Boehm, P.D., M.S. Steinhauer, D.R. Green, B. Fowler, B. Humphrey, D.I Feist and W.J. Cretney. 1987.
Comparative fate of chemically dispersed and beached crude oil in subtidal sediments of the Arctic
nearshore. Arctic 40 (Supp, 1): 133-148.
Brandvik, P. J, Johansen, O., Leirvik, F., Farooq, U. and P. S. Daling. 2013. Droplet breakup in sub-surface oil
releases—Part 1: Experimental study of droplet breakup and effectiveness of dispersant injection. Marine
Pollution Bulletin 73: 319-326.
Brannon, E.L., Collins, K., Cronin, M.A., Moulton, L.L., Maki, A.L. and K.R. Parker. 2012. Review of the Exxon
Valdez spill effects on pink salmon in Prince William Sound, Alaska. Reviews in Fisheries Science 20: 20-
60.
Boufadel, M.C., Abdollahi-Nasab, A., Geng, X., Galt, J. and J. Torlapati. 2014. Simulation of the landfall of the
Deepwater Horizon oil on the shorelines of the Gulf of Mexico. Environmental Science & Technology 48:
9496-9505.
Britannica. 2015. Deepwater Horizon Oil Spill of 2010. http://www.britannica.com/event/Deepwater-Horizon-oil-
spill-of-2010
Brodersen, C., J. Short, L. Holland, M. Carls, J. Pella, M. Larsen, and S. Rice. 1997. Evaluation of oil removal from
beaches 8 years after the Exxon Valdez oil spill. pp 325-336. In: Proceedings of the 22nd Arctic & Marine
Oilspill Program (AMOP) Technical Seminar, Environment Canada, Ottawa, ON, Canada.
Brown-Peterson, N.J., Brewton, R.A., Griffitt, R.J. and R.S. Fulford. 2014. Impacts of the Deepwater Horizon oil
spill on the reproductive biology of Spotted Seatrout (Cynoscion nebulosus). In: Alford, J. B., Peterson,
M. S. and C. C. Green (eds.) Impacts of Oil Spill Disasters on Marine Habitats and Fisheries in North
America. CRC Press, Boca Raton, FL.pp. 237-251.
Bunch, J.N. 1987. Effects of petroleum releases on bacterial numbers and microheterotrophic activity in the water
and sediment of an arctic marine ecosystem. Arctic 40 (Supp. 1): 172-183.
Camilli, R., Reddy, C. M., Yoerger, D.R., Van Mooy, B.A.S., Jakuba, M.V., Kinsey, J.C., McIntyre, C.P., Sylva,
S.P. and J. V. Maloney. 2010. Tracking hydrocarbon plume transport and biodegradation at Deepwater
Horizon. Science 330: 201-204.
Campo, P., Venosa, A.D. and M. T. Suidan. 2013. Biodegradability of Corexit 9500 and Dispersed South Louisiana
Crude Oil at 5 and 25 °C. Environmental Science & Technology 47(4): 1960-1967.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 472
CBC. 2006. Alberta establishes new disaster response agency. CBC News October 17, 2006.
http://www.cbc.ca/news/canada/edmonton/alberta-establishes-new-disaster-response-agency-1.582879.
CBC. 2008. Alberta band settle Wabamun oil-spill lawsuit. CBC News September 12, 2008.
http://www.cbc.ca/news/canada/calgary/alberta-band-settles-wabamun-oil-spill-lawsuit-1.767873
CCG (Canadian Coast Guard). 2015a. Independent Review of the M/V Marathassa Fuel Oil Spill Environmental
Response Operation. Report presented to Commissioner Jody Thomas, Canadian Coast Guard. July 19,
2015. http://www.ccg-gcc.gc.ca/independent-review-Marathassa-oil-spill-ER-operation
CCG. 2015b. Marathassa Spill Operational Update – April 23, 2015. Canadian Coast Guard. http://www.dfo-
mpo.gc.ca/media/infocus-alaune/2015/Marathassa/index-eng.htm
Chanton, J., Zhao, T., Rosenheim, B.E., Joye, S., Bosman, S., Brunner, C., Yeager, K.M., Diercks, A.R. and D.
Hollander. 2015. Using natural abundance radiocarbon to trace the flux of petrocarbon to the seafloor
following the Deepwater Horizon oil spill. Environmental Science & Technology 49: 847-854
Cretney, W.J., D.R. Green, B.R. Fowler, B. Humphrey, D.L. Feist and P.D. Boehm. 1987a. Hydrocarbon
biogeochemical setting of the Baffin Island oil spill experimental sites. I: Sediments. Arctic 40 (Supp. 1):
51-65.
Cretney, W.J. 1987b. Hydrocarbon biogeochemical setting of the Baffin Island oil spill experimental sites. II: Water
Arctic 40 (Supp. 1): 66-70.
Cretney, W.J., D.R. Green, B.R. Fowler, B. Humphrey, F.R. Engelhardt, R.J. Nordstrom, M. Simon, D.L. Feist and
P.D. Boehm. 1987a. Hyddrocarbon biogeochemical setting of the Baffin Island oil spill experimental sites.
III: Biota. Arctic 40 (Supp. 1): 71-79.
Cross, W.E., and C.M. Martin. 1983. In situ studies of effects of oil and chemically treated oil on primary
productivity of ice algae and on under-ice meiofanual and macrofaunal communities. Special studies –
1983 study results. Baffin Island Oil Spill Project Working Report 82-7. Environmental Protection Service,
Environment Canada. 103 pp.
Cross, W.E. and D.H. Thompson. 1987. Effects of experimental releases of oil and dispersed oil on Arctic nearshore
macrobenthos. I. Infauna. Arctic 40 (Supp. 1): 184-200.
Cross, W.E., R.T. Wilce and M.F. Fabijan. 1987b. Effects of experimental releases of oil and dispersed oil on Arctic
nearshore macrobenthos. III; Macroalgae. Arctic 40 (Supp. 1): 211-219.
Cross, W.E., C.M. Martin and D.H. Thomson. 1987b. Effects of experimental releases of oil and dispersed oil on
Arctic nearshore macrobenthos. II; Epibenthos. Arctic 40 (Supp. 1): 201-210.
DeBruyn, A.M.H., Wernick, B.G., Stefura, C., McDonald, B.G., Rudolph, B-L., Patterson, L. and P.M. Chapman.
2007. In situ experimental assessment of Lake whitefish development following a freshwater oil spill.
Environmental Science and Technology 41: 6983-6989.
Dickins, D.F. and J.A. Hellebust. 1981. Return to Balaena Bay: Long-term effect of a large scale crude oil spill
under Arctic sea ice. Gylf Canada Resources Inc., Calgary, AB.
Dickins, D. 2011. Behavior of oil spills in ice and implications for Arctic spill response. Offshore Technology
Conference. OTC 22126. Houston, Texas, February 709, 2011.
Diercks, A-R., Highsmith, R.C., Asper, V.L., Joung, D., Zhou, Z., Guo, L., Shiller, A.M., Joye, S.B., Teske, A.P.,
Guinasso Jr., N., Wade, T.L. and S.E. Lohrenz. 2010. Characterization of subsurface polycyclic aromatic
hydrocarbons at the Deepwater Horizon Site, Geophysical Research Letters 37.
Di Toro, D.M., McGrath, J.A and D.J. Hansen. 2000. Technical basis for narcotic chemicals and polycyclic aromatic
hydrocarbon criteria. I. Water and tissue. Environmental Toxicology and Chemistry 19: 1951-1970.
Di Toro, D.M. and J.A. McGrath. 2000. Technical basis for narcotic chemicals and polycyclic aromatic hydrocarbon
criteria. II. Mixtures and sediments. Environmental Toxicology and Chemistry 19:1971-1982.
DNV (Det Norske Veritas). 2011. Final Report Assessment of the Risk of Pollution from Marine Oil Spills in
Australian Ports and Waters. Report to Australian Maritime Safety Authority. Report No. PP002916, Rev.
5. Cited in WSP and SL Ross. 2014a. Risk Assessment for Marine Spill in Canadian Waters. Phase 1: Oil
Spills South of 60th Parallel. Final Study Report prepared for Transport Canada.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 473
http://wcel.org/sites/default/files/file-downloads/131-17593-00_ERA_Oil-Spill-South_150116_pp1-
124.pdf.
Donahue, W.F., Allen, E.W., and D.W. Schindler. 2006. Impacts of coal-fired power plants on trace metals and
polycyclic aromatic hydrocarbons (PAHs) in lake sediments in central Alberta, Canada. Journal of
Paleolimnology 35: 111-128.
Eimhjellen,K. and K. Josefson. 1984. Microcbilogy 2: Biodegradation of stranded oil – 1982 study results. Baffin
Island Oil Spill Project Worksing Report 83-6. Environmental Proection Service, Environment Canada,
Ottawa. 58 pp.
Energy East. 2014a. Accidents and Malfunctions. Volume 6 of the Environmental Impact Assessment. Section 3:
Crude oil characteristics, environmental fate, transport and effects. Energy East Pipeline Project. 48 pp.
Available online from the National Energy Board at https://docs.neb-one.gc.ca/ll-
eng/llisapi.dll/fetch/2000/90464/90552/2432218/2540913/2543426/2543068/A63989-
4_ESA_V6_S3_OilFateTranspEffects_-_A4E1F4.pdf?nodeid=2543560&vernum=-2
Energy East. 2014b. Accidents and Malfunctions. Volume 6 of the Environmental Impact Assessment. Section 4:
Sites of interest. 112 pp. Available on-line from the National Energy Board at https://docs.neb-one.gc.ca/ll-
eng/llisapi.dll/fetch/2000/90464/90552/2432218/2540913/2543426/2543068/A63989-
5_ESA_V6_S4_SitesOfInterest_-_A4E1F5.pdf?nodeid=2543268&vernum=-2
EPA (US Environmental Protection Agency) 1998. Guidelines for Ecological Risk Assessment. EPA/630/R-
95/002F.
EPA. 2012. Response Technologies for Oil Sands Products: Enbridge Oil Spill Case Study. Kalamazoo River,
Michigan. Presentation. http://www.southportland.org/files/8813/9509/2248/Water_Quaility_Impacts_-
_Response_Technologies_for_Oil_Sands_Products___Enbridge_Oil_Spill_Case_Study_Kalamazoo.pdf.
Etkin, D.S. 2015. New risks from crude-by-rail transportation. In: Proceedings of the Thirty-Eight AMOP Technical
Seminar, Environment Canada, Ottawa, ON. pp 900-923.
Evans, C.L. 2008. An Evaluation of the Harmful Alteration, Disruption and Destruction (HADD) of Fish Habitat
Following the Bunker “C” And Pole Treating Oil Release on Wabamun Lake, Alberta. Report submitted to
Public Prosecution Service of Canada, Alberta Regional Office, Fisheries and Oceans Canada, Prairies
Area Operations, Central and Arctic Region.
EVOSTC. 2015. Status of Injured Resources and Services. Exxon Valdez Oil Spill Trustee Council.
http://www.evostc.state.ak.us/index.cfm?FA=status.injured
Federal Interagency Solutions Group. 2010. Oil Budget Calculator: Deepwater Horizon.
http://www.restorethegulf.gov/sites/default/files/documents/pdf/OilBudgetCalc_Full_HQ-Print_111110.pdf
Fitzpatrick, F.A. 2014. Effects of Diluted Bitumen in the Environment: Introductory Remarks. Presentation.
NAS/NRC 1st Committee Meeting. December 3, 2014. http://nas-
sites.org/dilbit/files/2014/12/5_Fitzpatrick_USGS.pdf.
Fitzpatrick, F.A., Boufadel, M.C., Johnson, R., Lee, K., Graan, T.P., Bejarano, A.C., Zhu, Z., Waterman, D.,
Capone, D.M., Hayter, E., Hamilton, S.K., Dekker, T., Garcia, M.H. and J.S. Hassan. 2015. Oil-Particle
Interactions and Submergence from Crude Oil Spills in Marine and Freshwater Environments – Review of
the Science and Future Science Needs. U.S. Geological Survey. Open File Report 2015-1076.
Fodrie, F.J. and K.L. Heck Jr. 2011. Response of coastal fishes to the Gulf of Mexico oil disaster. PLoS ONE 6:
e21609. doi:10.1371/journal.pone.0021609. Cited in Mendelssohn, I.A., Anderson, G.L., Baltz, D.M.,
Caffey, R.H., Carman, K.R., Fleeger, J.W., Joye, S.B., Lin, Q., Maltby, E., Overton, E.B. and L.P. Rozas.
2012. Oil impacts on coastal wetlands; implications for the Mississippi River Delta ecosystem after the
Deepwater Horizon oil spill. BioScience 62: 562-574.
Foght, J. 2006. Potential for Biodegradation of Sub-Littoral Residual Oil by Naturally Occurring Microorganisms
Following the Lake Wabamun Train Derailment. Submitted to Alberta Environment and Environment
Canada. Dept. Biological Sciences, University of Alberta.
Fu, J., Y. Gong, Y., Zhao, X., O’Reilly, S.E. and D. Zhao. 2014. Effects of oil and dispersant on formation of
marine oil snow and transport of oil hydrocarbons. Environmental Science and Technology
48:14392−14399
Fulford, R.S., Griffit, R.J., Brown-Peterson, N.J., Perry, H. and G. Sanchez-Rubio. 2014. Impacts of the Deepwater
Horizon oil spill on blue crab, Callinectes sapidus, larval settlement in Mississippi. In: Alford, J. B.,

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 474
Peterson, M. S. and C. C. Green (eds.) Impacts of Oil Spill Disasters on Marine Habitats and Fisheries in
North America. CRC Press, Boca Raton, FL. Pp. 253-264.
Garcia-Pineda, O., I. MacDonald, C. Hu, J. Svejkovsky, M. Hess, D. Dukhovskoy, and S.L. Morey. 2013. Detection
of floating oil anomalies from the Deepwater Horizon oil spill with synthetic aperture radar. Oceanography
26(2):124–137, http://dx.doi.org/10.5670/oceanog.2013.38.
GENIVAR. 2013. Évaluation Environnementale Stratégique sur la Mise en Valeur des Hydrocarbures dans les
Bassins d’Anticosti, de Madeleine et de la Baie des Chaleurs. Rapport de GENIVAR au Ministère des
Ressources Naturelles. Cited in WSP and SL Ross. 2014a. Risk Assessment for Marine Spill in Canadian
Waters. Phase 1: Oil Spills South of 60th Parallel. Final Study Report prepared for Transport Canada.
http://wcel.org/sites/default/files/file-downloads/131-17593-00_ERA_Oil-Spill-South_150116_pp1-
124.pdf.
George-Ares, A. and J. R. Clark. 2000. Aquatic toxicity of two Corexit dispersants. Chemosphere 40: 897-906.
Gilde, K., and J. Pinckney. 2012. Sublethal effects of crude oil on the community structure of estuarine
phytoplankton. Estuaries and Coasts 35: 853-861.
Goldberg, H. 2011. Pine River 2011 Fisheries Update: Status of Recovery Post-2000 Pipeline Rupture. Report
Submitted to Enbridge, c/o Paul Anderson, Director of Environment Northern Gateway Project. Response
to Haisla Nation IR 1.52. National Energy Board Northern Gateway Application.
Goldsmith, B.J., Waikem, T.K. and T. Franey. 2014. Environmental damage liability regimes concerning oil spills –
A global review and comparison. In: Proceedings of the 2014 International Oil Spill Conference.
http://ioscproceedings.org/doi/pdf/10.7901/2169-3358-2014.1.2172
Golder Associates Ltd. 2007 Lakewater and Aquatics Long-Term Monitoring –2005-2006 Interpretive Report,
Wabamun Lake Derailment Site. CanadianNational Railway Company PIN 2401257 Wabamun, Alberta.
Report prepared for CN Railway Co. by Golder Associates Ltd., North Vancouver, B.C. Cited in Birtwell,
I.K. 2003. Comments on the Effects of Crude Oil on Fish and Their Habitat in the Pine River, British
Columbia. Fisheries and Oceans Canada, West Vancouver, BC.
Gordon, D.C.Jr., Dale, J.and P.D. Keizer. 1978. Importance of sediment working by the deposit-feeding polychaete
Arenicola marina on the weathering rate of sediment-bound oil. Journal of the Fisheries Research Board of
Canada 35(5): 591-603.
Government of Alberta. 2015. Alberta Guide to Sportfishing Regulations. http://albertaregulations.ca/fishingregs-
pdfs-2015.html
Graham, W. M., Condon, R. H. Carmichael, R. H., D’Ambra, I., Patterson, H.K. Linn, L.J. and F. J. Hernandez.
2010. Oil carbon entered the coastal planktonic food web during the Deepwater Horizon oil spill.
Environmental Research Letters 5: 045301 (6p).
Green, J., Postlewaite, L. and P. Anderson. 2015. Context for assessing effects on the biophysical and human
environment – the pipeline ecological and human health risk assessment for the Northern Gateway Pipeline
Project. In: Proceedings of the Thirty-Eight AMOP Technical Seminar. Environment Canada, Ottawa, ON,
pp 526-535.
Hamilton, S. 2015. Presentation to the Royal Society Expert Panel on the Behaviour and Environmental Impacts of
Crude Oil Released in Aqueous Environments. April, 2015, Halifax, NS.
Harwell, M.A. and J.H. Gentile. 2006. Ecological significance of residual exposures and effects from the Exxon
Valdez oil spill. Integrated Environmental Assessment and Management 2: 204-246.
Hauge, K.H., Blanchard, A., Anderson, G. Boland, R. Grøsvik, B.E. Howell, D. Meier, S. Olsen, E. and F. Vikebø.
2014. Inadequate risk assessments –A study on worst-case scenarios related to petroleum exploitation in the
Lofoten area. Marine Policy 44: 82-89
Hazen, T.C., Dubinsky, E.A., Anderson, G.L., Piceno, Y.M., Singh, N., Jasson, J.K., Probst, A., Borglin, S.E.,
Fortney, J.L., Stringfellow, W.T., Bill, M., Conrad, M.S., Tom, L.M., Chavarria, K.L., Alusi, T.R.,
Lamendella, R., Joyner, D.C., Spier, C., Baelum, J., Auer, M., Zemla, M.L., Chakraborty, R., Sonnenthal,
E.L., D’haeseleer, P., Holman, H.-Y.N., Nostrand, J.D.V., Deng, Y., Zaou, J., and O.U. Mason. 2010.
Deep-sea oil plume enriches indigenous oil-degrading bacteria. Science 330: 204-208.
Hodson, P.V. 2008. Report on the Toxicity of Oil to Fish. Submitted to Fisheries and Oceans Canada. DFO Contract
# F2471-080006.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 475
Hodson. P.V. and J. Martin. 2012. Review of Technical Data Report. Ecological and human health risk assessment
for pipeline spills: Enbridge Northern Gateway Project. Prepared for the Haisla Nation Council, Kitimaat
Village, BC.
Hodson, P.V., Shaw, W., Abudulai, N. and R.S. Brown. 2007. Were Fish in L. Wabamun Exposed to Spilled Oil
After the Train Derailment in 2005? Platform presentation, 28th Arctic and Marine Oil spill Program
(AMOP) Technical Seminar, Edmonton, AL, June 4 7, 2007.
Hollebone, B.P. 2008. Physical and Chemical Fate and Behaviour of Oil Spilled in Wabamun Lake, August 2005.
Environmental Science and Technology Centre, Science and Technology Branch, Environment Canada. 45
pp.
Hollebone, B.P., Fieldhouse, B., Sergey, G., Lamter, P., Wang, Z., Yang, C. and M. Landriaut. 2011. The behaviour
of heavy oil in fresh water lakes. In: Proceedings of the 34th Arctic and Marine Oilspill Program (AMOP)
Technical Seminar on Environmental Contamination and Response. Environment Canada, Ottawa. Pp.
668-709.
Horn, M. and D. French-McCay. 2015. Trajectory and fate modeling with acute effects assessment of hypothetical
spills of diluted bitumen into rivers. In: Proceedings of the Thirty-Eight AMOP Technical Seminar.
Environment Canada, Ottawa, ON, pp 549-581.
Humphrey, B., P.D. Boehm, M.C. Hamilton and R.J. Nordstrom. 1987a. The fate of chemically dispersed and
untreated crude oil in Arctic benthic biota. Arctic 40 (Supp. 1): 149-161.
Humphrey, B., D.R. Green, B.r. Fowler, D. Hope and P.D. Boehm. 1987b. The fate of oil in the water column
following experimental oil spills in the Arctic marine nearshore. Arctic 40 (Supp. 1): 124-132.
Iverson, S. A. and D. Esler. 2010. Harlequin duck population injury and recovery dynamics following the 1989
Exxon Valdez oil spill. Ecological Applications 20:1993–2006. Cited in Ballachey, B.E., Bodkin, J.L.,
Esler, D. and S.D. Ricer. 2014. Lessons from the 1989 Exxon Valdez oil spill: A biological perspective. In:
Alford, J. B., Peterson, M. S. and C. C. Green (eds.) Impacts of Oil Spill Disasters on Marine Habitats and
Fisheries in North America. CRC Press, Boca Raton, FL. Pp. 181-197.
Johansen, Ø, Brandvik, P. J. and U. Farooq. 2013. Droplet breakup in subsea oil releases—Part 2: Predictions of
droplet size distributions with and without injection of chemical dispersants. Marine Pollution Bulletin
73:327-335.
Kane, L. 2015. Vancouver fuel spill underscores gap in research after federal cuts: aquarium. Global and Mail April
16, 2015. http://www.theglobeandmail.com/news/british-columbia/vancouver-fuel-spill-underscores-gap-
in-research-after-federal-cuts-aquarium/article23999926/
Kinder Morgan. 2013. Trans Mountain Pipeline (ULC) Transmountain Expansion Project. Volume 7 – Risk
Assessment and Management of Pipeline and Facility Spills. + Appendices B 18-2, B18-3, B18-4, B18-5,
B18-6, B18-7, B18-8, B18-12. Submitted to the National Energy Board. https://docs.neb-one.gc.ca/ll-
eng/llisapi.dll?func=ll&objId=2393783&objAction=browse&viewType=1.
Kinder Morgan. 2014. Trans Mountain Pipeline ULC Trans Mountain Expansion Project. NEB Hearing Order H-
001-2014. Responses to Information Request from Village of Belcarra. 2007 Oil Spill Event: ‘Post
Mortem’, ‘Lessons Learned’ & ‘Corrective Actions’. Trans Mountain Response to Village of Belcarra IR
No.2. http://www.belcarra.ca/reports/TMEP_Response_to_Belcarra_Information_Request_No.2.pdf.
Kleindienst, S., Seidel, M., Ziervogel, K., Grim, S., Loftis, K., Harrison, S., Malkin, S.Y., Perkins, M.J., Field, J.,
Sogin, M.L., Dittmar, T.,Passow, U., Medeiros, P.M. and S.B. Joye. 2015. Chemical dispersants can
suppress the activity of natural oil-degrading microorganisms. Proceedings National Academy of Sciences
(USA) 112: 14900-14905.
Kostka, J.E., Prakash, O., Overholt, W.A., Green, S.J., Freyer, G., Canion, A., Delgardio, J., Norton, N., Hazen, T.C.
and M. Huettel. 2011. Hydrocarbon-degrading bacteria and the bacterial community response in Gulf of
Mexico beach sands impacted by the Deepwater Horizon oil spill. Applied and Environmental
Microbiology 77: 7962-7974.
Kujawinski, E.B., Kido Soule, M.C., Valentine, D.L., Boysen, A.K., Longnecker, K. and M.C. Redmond. 2011. Fate
of dispersants associated with the Deepwater Horizon spill. Environmental Science and Technology 45:
1298-1306.
Lake Wabamun Residents Committee. 2007. The Lake Wabamun Disaster: A Catalyst for Change. Report to the
Railway Safety Act Review Advisory Panel. August 31, 2007. 12 pp.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 476
Landis, W.G. 2007. The Exxon Valdez oil spill revisited and the dangers of normative science. Integrated
Environmental Assessment and Management 3: 439-441.
Lee, K., Wohlgeschaggen, G.D., Tremblay, G.H., Vandermeulen, J.H., Mossman, D.C., Doe, K.G., Jackman, P.M.,
Wilson, J.E.H., Prince, R.C., Garrett, R.M. and C.E. Haith. 1999. Natural recovery reduces impact of the
1970 Arrow oil spill. In: Proceedings of the 1999 International Oil Spill Conference. Pp. 1075-1078.
Lee, K., Prince, R.C., Greer, C.W., Doe, K.G., Wilson, J.E.H., Cobanli, S.E., Wohlgeschaffen, G.D., Alroumi, D.,
King, T. and G.H. Tremblay. 2003. Composition and toxicity of residual Bunker C fuel oil in intertidal
sediments after 30 years. Spill Science and Technology Bulletin 8: 187-199.
Lee,K. M. Boudreau, J. Bugden, L. Burridge, S.E. Coblani, S. Courtenay, S. Grenon, B. Hollebone, P. Kepkay, Z.
Li, M. Lyons, H. Niu, T.L. King, S. MacDonald, E.C. McIntyre, B. Robinson, S.A. Ryan and G.
Wohlgeschaffen. 2011. State of knowledge review of fate and effect of oil in the Arctic marine
environment. National Energy Board of Canada. 259 p. Available: https://docs.neb-one.gc.ca/ll-
eng/llisapi.dll/fetch/2000/90463/621169/700096/704342/A2A8R2_-
_NEB_State_of_Knowledge_Review_of_Fate_and_Effect_of_Oil_DFO_COOGER.pdf?_gc_lang=en&nod
eid=704343&vernum=0
Lee, K., Nedwed, T., Prince, R.C. and D. Palandro. 2013. Laboratory tests on the biodegradation of chemically
dispersed oil should consider the rapid dilution that occurs at sea. Marine Pollution Bulletin 73: 314–318.
La Peyre, J, S. Casa and S. Miles. 2014. Oyster responses to the Deepwater Horizon oil spill across coastal
Louisisana: examining oyster health and hydrocarbon bioaccumulation. In: Alford, J. B., Peterson, M. S.
and C. C. Green (eds.) Impacts of Oil Spill Disasters on Marine Habitats and Fisheries in North America.
CRC Press, Boca Raton, FL. Pp. 269-288
Lemly, A.D. 1993. Metabolic stress during winter increases the toxicity of selenium to fish. Aquatic Toxicology 27:
133-158. Cited in Birtwell, I.K. 2003. Comments on the Effects of Crude Oil on Fish and Their Habitat in
the Pine River, British Columbia. Fisheries and Oceans Canada, West Vancouver, BC.
Levy, E.M. 1972. Evidence for the recovery of the waters off the east coast of Nova Scotia from the effects of a
major oil spill. Water, Air and Soil Pollution 1: 144-148.
Lilwal, S. and E. Fitzpatrick. 2015. Wabamun Lake oil spill: A decade later, disaster still fresh in resident’s minds.
CBC News August 3, 2015. http://www.cbc.ca/news/canada/edmonton/wabamun-lake-oil-spill-a-decade-
later-disaster-still-fresh-in-residents-minds-1.3177556
Maritime Museum of the Atlantic. 2015. Marine Heritage Database. “On the Rocks: Find a Wreck”. Arrow – 1970.
Online. Arrow Shipwreck - Maritime Museum of the Atlantic fact sheet.htm
http://novascotia.ca/museum/wrecks/wrecks/shipwrecks.asp?ID=468
McCleneghan, K. 2008. Analysis of Rsponse to the Lake Wabamun Oil Spill. Prepared for Fisheries and Oceans
Canada, Edmonton, Alberta. 101 pp.
McDonald, B.G., deBruyn, A. M.H., Wernick, B.G., Patterson, L., Pilerin, N. and P.M. Chapman. 2007. Design and
application of a transparent and scalable weight-of-evidence framework: an example from Wabamun Lake,
Alberta, Canada. Integrated Environmental Assessment and Management 3: 476-483.
McKenzie, J.F., Schieble, C., Smith, P.W. and M.T. O’Connell. 2014. Occurrence of lemon sharks (Negaprion
brevirostris) at Chandeleur Islands, Louisiana, before and after the 2010 Deepwater Horizon disaster. In:
Alford, J. B., Peterson, M. S. and C. C. Green (eds.) Impacts of Oil Spill Disasters on Marine Habitats and
Fisheries in North America. CRC Press, Boca Raton, FL. Pp. 295-309.
Mendelssohn, I.A., Anderson, G.L., Baltz, D.M., Caffey, R.H., Carman, K.R., Fleeger, J.W., Joye, S.B., Lin, Q.,
Maltby, E., Overton, E.B. and L.P. Rozas. 2012. Oil impacts on coastal wetlands; implications for the
Mississippi River Delta ecosystem after the Deepwater Horizon oil spill. BioScience 62: 562-574.
Millsap, S., Williams, L., Haas, J., Hanshue, S., Wesley, J., Taft, W., Walterhouse, M., Winter, J., Williamson, R.T.,
Beltman, D., Ebbets, A., Ritter, K., Tilleitt, D.E., Papoiulias, D. Nicks, D. and P. Badra. 2012. Assessing
Natural Resource Impacts from the Enbridge Pipeline Spill into the Kalamazoo River. Presentation.
https://crrc.unh.edu/sites/crrc.unh.edu/files/media/docs/Workshops/oil_sands/Winter_naturalresourceimpac
ts.pdf
NORCOR Engineering & Research Ltd. 1975. The interaction of crude oil with Arctic sea ice. Beaufort Sea Project
Technical Report No. 27. Canadian Department of Environment, Victoria, BC. 201 pp.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 477
NRC (National Research Council). 2013. An Ecosystem Services Approach to Assessing the Impacts of the
Deepwater Horizon Oil Spill in the Gulf of Mexico. National Research Council of the National Academies.
National Academies Press, Washington, DC. http://www.nap.edu/catalog.php?record_id=18387.
Neff JM, Owens EH, Stoker SW, McCormick DM. 1995. Shoreline oiling conditions in Prince William Sound
following the Exxon Valdez oil spill. In: Wells PG, Butler JN, Hughes JS, editors. Exxon Valdez oil spill:
Fate and effects in Alaskan waters. Philadelphia (PA): American Society for Testing and Materials. ASTM
STP 1219. p 312–345.
Noble, S. 2012. Two years after the Enbridge pipeline rupture. EPA Situation Report. Presentation. Michigan
Department of Environmental Quality. Enbridge Response Unit. http://www.mi-
wea.org/docs/Two%20Year%20after%20the%20Enbridge%20Pipeline%20Rupture%20MWEA.pdf
North Saskatchewan Watershed Alliance. 2015. Riparian Health Assessment of Wabamun Lake: An Aerial
Assessment Using an Unmanned Air Vehicle (UAV). Prepared for the Wabamum Watershed Management
Council, Edmonton, AB.
https://www.nswa.ab.ca/sites/default/files/Riparian%20Health%20Assessment%20Wabamun.pdf.
NTSB (National Transportation Safety Board). 1990. Marine Accident Report: Grounding of the U.S. tankship
Exxon Valdez on Bligh Reef, Prince William Sound near Valdez, Alaska. March 24, 1989. U.S. National
Transportation Safety Board. NSB/MAR-90/04. PB90-916405.
http://docs.lib.noaa.gov/noaa_documents/NOAA_related_docs/oil_spills/marine_accident_report_1990.pdf
NTSB. 2012. Accident Report: Enbridge Incorporated Hazardous Liquid Pipeline Rupture and Release, Marchall,
Michigan. July 25, 2010. U.S. National Transportation Safety Board. NTSB/PAR-12/01. Washington, DC.
http://www.ntsb.gov/investigations/AccidentReports/Reports/PAR1201.pdf.
OSAT. 2011. Operations Science Advisory Team (OSAT) Unified Area Command Summary Report for Sub-Sea
and Sub-Surface Oil and Dispersant Detection: Ecotoxicity Addendum. Prepared for Julia A. Hein, CPTN,
U.S. Coast Guard, Federal On-Scene Coordinator Deepwater Horizon MC252. July 8, 2011.
Owens, E.H. 1978. Mechanical dispersion of oil stranded in the littoral zone. Journal of the Fisheries Research
Board of Canada 35: 563-572.
Owens, E.H., J.R. Harper, W. Robson and P.D. Boehm. 1987. Fate and persistence of crude oil stranded on a
sheltered beach, Arctic 40 (Supp. 1): 109-123.
Owens EH. 1991. Shoreline conditions following the Exxon Valdez spill as of fall 1990. In: Proceedings of the 14th
Arctic and Marine Oilspill Program Technical Seminar; 1991 June 12–14; Vancouver, BC. Ottawa (ON):
Environment Canada. p 576–606.
Owens, E.H., G.A. Sergy, B.E. McGuire and B. Humphrey. 1993. The 1970 Arrow soil spill – What remains on the
shoreline 22 years later?. Proceedings of the 16th Arctic and Marine Oilspill Programme (AMOP) Technical
Seminar. Environment Canada, Ottawa, On. p1149-1167.
Owens, E.H., Prince, R.C. and R.B. Taylor. 2008. Natural attenuation of heavy oil on a coarse sediment beach:
results from Black Duck Cove, Nova Scotia, Canada over 35 years following the Arrow oil spill. In:
Proceedings of the 31st AMOP technical seminar on environmental contamination and response.
Owens, E.H. 2010. Shoreline response and long-term oil behaviour studies following the 1970 “Arrow” spill in
Chedabucto Bay, NS. Proceedings of the 33rd Arctic and Marine Oilspill Programme (AMOP) Technical
Seminar. Environment Canada, Ottawa, On.
Paris, C. B., Le Hénaff, M., Aman, Z.M., Subramaniam, A., Helgers, J., Wang, D. -P.,Kourafalou, V.H. and A.
Srinivasan. 2012. Evolution of the Macondo Well blowout: Simulating the effects of the circulation and
synthetic dispersants on the subsea oil transport. Environmental Science & Technology 46(24): 13293-
13302.
Parker-Hall, H.A. and E.H. Owens. 2006. Lake Wabamun Derailment: Fate and Persistence of the Spilled Oil.
Polaris Applied Sciences. Prepared for CN. 38 pp.
Peterson, C.H., Rice, S.D., Short,J.W. Esler, D., Bodkin, J.L., Ballachey, B.E. and D.B. Irons. 2003. Long-term
ecosystem response to the Exxon Valdez oil spill. Science 302: 2082-2086.
Peterson, C. H., Anderson, S.S., Cherr, G.N., Ambrose, R.F., Anghera, S. Bay, S., Blum, M., Condon, R., Dean,
T.A., Graham, M. Guzy, M., Hampton, S. Joye, S., Lambrinos, J., Mate, B., Meffert, D., Powers, S.P.,

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 478
Somasundaran, P., Spies, R.B., Taylor, C.M., Tjeerdema, R. and E.E. Adams. 2012. A tale of two spills:
novel science and policy implications of an emerging new oil spill model. Bioscience 62: 461-469.
Platt, J.F. and R. G. Ford. 1996. How many seabirds were killed by the Exxon Valdez spill? American Fisheries
Society Symposium 18: 712-719.
Prepas, E. and P. Mitchell. 1990. Atlas of Alberta Lakes. University of Alberta Press. Digital edition available at:
http://sunsite.ualberta.ca/Projects/Alberta-Lakes/preface.php
Prince, R.C. 2015. Oil spill dispersants; bone or bane? Environmental Science & Technology 49: 6376-6384.
Prince, R.C., Garrett, R. M., Bare, R.E., Grossman, M.J., Townsend, T., Suflita, J.M., Lee, K., Owens, E.H., Sergy,
G.A., Braddock, J.F., Lindstrom, J.E. and R. R. Lessard. 2003. The roles of photooxidation and
biodegradation in long-term weathering of crude and heavy fuel oils. Spill Science and Technology Bulletin
8: 145-156.
Pritchard, P.H., Mueller, J.G., Rogers, J.C., Kremer, F.V. and J.A. Glaser. 1992. Oil spill bioremediation:
experiences, lessons and results from the Exxon Valdex oil spill in Alaska. Biodegradation 3: 315-335.
Reddy, C. M., Arey, J. S., Seewald, J.S., Sylva, S.P., Lemkau, K.L., Nelson, R.K., Carmichael, C.A., McIntyre,
C.P., Fenwick, J. and G. T. Ventura. 2011. Composition and fate of gas and oil released to the water
column during the Deepwater Horizon oil spill. Proceedings of the National Academy of Sciences of the
United States of America 109:20229-20234.
Redmond, M.C. and D.L. Valentine. 2012. Natural gas and temperature structured a microbial community response
to the Deepwater Horizon oil spill. Proceedings of the National Academy of Sciences 109: 20292-20297.
Reich, D.A., Balouskus, R., French-McCay, D., Fontenault, J. Rowe, J., Singer-Leavitt, Z., Schmidt Erkin, D.,
Michel, J., Nixon, Z., Boring, C., McBrien, M. and B. Hay. 2014. Assessment of Marine Oil Spill Risk and
Environmental Vulnerability for the State of Alaska. Submitted to the National Oceanic and Atmospheric
Administration, Seattle, WA. NOAA Contract Number: WC133F-11-CQ-0002.
https://alaskafisheries.noaa.gov/habitat/restoration/oilspill/riskreport.pdf
Rico-Martinez, R., Snell, T.W., and T.L. Shearer. 2013. Synergistic toxicity of Macondo crude oil and dispersant
Corexit 9500A® to the Brachionus plicatilis species complex (Rotifera). Environmental Pollution 173: 5-
10.
Schindler, D.W., Anderson, A.-M., Brzustowski, J., Donahue, W.F., Goss, G., Nelson, J., St. Louis, V., Sullivan, M.
and S. Swanson. 2004. Lake Wabamun:A Review of Scientific Studies and Environmental Impacts.
Submitted to the Minister of Alberta Environment, Edmonton, AB.
http://www3.gov.ab.ca/env/water/reports/wabamun/Wabamun_Report_Dec04.pdf.
Sempels, J.-M. 1982. Coastlines of the Eastern Arctic. Arctic 35: 170-179.
Sergy, G.A and P.J. Blackall. 1987. Design and conclusions of the Baffin Island Oil Spill Project. Arctic 40 (Supp
1): 1-9.
Shigenaka, G. 2014. Twenty-Five Years After the Exxon Valdez Oil Spill: NOAA’s Scientific Support, Monitoring,
and Research. Seattle: NOAA Office of Response and Restoration.
http://response.restoration.noaa.gov/sites/default/files/Exxon_Valdez_25YearsAfter_508_0.pdf
Short, J. W., M. R. Lindeberg, P. A. Harris, et al. 2004. Estimate of oil persisting on beaches of Prince William
Sound, 12 years after the Exxon Valdez oil spill. Environmental Science & Technology 38:19–25.
Short, J.W. 2008. An Evaluation of Oil Fate, Persistence and Toxic Potential to Fish Following the August 2005
Canadian National Derailment Spill into Lake Wabamun, Alberta. August 27, 2008. https://docs.neb-
one.gc.ca/ll-eng/llisapi.dll/Open/2586376.
Short, J.W. 2015. Presentation to the Royal Society Expert Panel on the Behaviour and Environmental Impacts of
Crude Oil Released in Aqueous Environments. April, 2015, Halifax, NS.
Silliman, B.R., van de Koppel, J., McCoy, M.W., Diller, J., Kasozi, G.N., Earl, K., Adamns, P.N. and A.R.
Zimmerman. 2012. Degradation and resilience in Louisiana salt marshes after the BP-Deepwater Horizon
oil spill. Proceedings of the National Academy of Sciences 109: 11234-11239.
Spies, R.B., Rice, S.D., Wolfe, D.A. and B.A. Wright. 1996. The effect of the Exxon Valdez oil spill on Alaskan
coastal environment. In: Proceedings of the 1993 Exxon Valdez Oil Spill Symposium. American Fisheries
Society, Bethesda, MD. Pp. 1-16
SL Ross. 2014. Guideline Document for Measuring the Biological Effects of Accidental Oil Spills. Submitted to
Fisheries and Oceans Canada, National Contaminants Advisory Group, Ottawa, ON.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 479
Socolosky, S.A., Adams, E.E. and C.R. Sherwood. 2011. Formation dynamics of subsurface hydrocarbon intrusions
following the Deepwater Horizon blowout. Geophysical Research Letters 38: 1-6.
Stanbury, M., Hekman, K., Wells, E., Miller, C., Smolinske, S. and J. Rutherford. 2010. Acute Health Effects of the
Enbridge Oil Spill. Michigan Department of Community Health, Lansing, MI.
http://www.michigan.gov/documents/mdch/enbridge_oil_spill_epi_report_with_cover_11_22_10_339101_
7.pdf.
Stantec. 2012. Summary of Clean Up and Effects of the 2007 Spill of Oil from Trans Mountain Pipeline to Burrard
Inlet. http://www.transmountain.com/uploads/pages/1374960772-2012-Summary-2007-Spill-Cleanup---
Effects-REV2.pdf
State of Alaska. 1993. Shoreline treatment techniques. In: The Exxon Valdez Oil Spill: Final Report, State of Alaska
Response. Alaska Department of Environmental Conservation. Pp. 61-87.
Stephenson, M., Henderson, J., Mazzacco, P. and A. St-Amand. 2015. Ecological risk assessment of hypothetical
spills of diluted bitumen in rivers. In: Proceedings of the Thirty-Eight AMOP Technical Seminar.
Environment Canada, Ottawa, ON, pp. 607-630.
Sumners, J. 2001. July 31 2000 Pine River Oil Spill Physical Fish Habitat Impact Statement. Submitted to John
Cliffe, Crown Counsel, Department of Justice.
Tarnecki, J.H. and W.F. Patterson III. 2015. Changes in red snapper diet and trophic ecology following the
Deepwater Horizon oil spill. Marine and Coastal Fisheries 7: 135-147
Thomas, M.L. 1978. Comparison of oiled and unoiled intertidal communities in Chedabucto Bay, Nova Scotia.
Journal of the Fisheries Research Board of Canada 35(5): 707-716.
Thormann, M.N. and S.E. Bayley. 2008. Impacts of the CN Rail Oil Spill on Softstem Bulrush-Dominated
Lacustrine Marshes in Wabamun Lake. Report submitted to Alberta Environment, Edmonton, AB.
TSB (Transportation Safety Board of Canada). 2007. Pipeline Investigation Report: Crude oil Pipeline – Third-Party
Damage. Trans Mountain Pipeline L.P. 610-millimetre-diameter crude oil pipeline. Kilometre post 3.10,
Westridge Dock Transfer Line, Burnaby, BC. Report P07H0040.
USGS. 2015. Research, Response for Future Oil Spills: Lessons Learned from Deepwater Horizon.
http://www.usgs.gov/newsroom/article.asp?ID=3469&from=rss#.VeUu9ZWFOpo
Valentine, D. L. Fisher, G.B., Bagby, S.C., Nelson, R.K., Reddy, C.M., Sylva, S.P. and M.A. Woo. 2014. Fallout
plume of submerged oil from Deepwater Horizon. Proceedings of the Nationa. Academy of Sciences USA.
http://www.pnas.org/cgi/doi/10.1073/pnas.1414873111
Valentine, D.L., Kessler, J.D., Redmond, M.C., Mendes, S.D., Heintz, M.B., Farwell, C., Hu, L., Kinnaman, F.S.,
Yvon-Lewis, S., Du, M.R., Chan, E.W., Tigreros, F.G. and C.J. Villanueva. 2010. Propane respiration
jump-starts microbial response to a deep oil spill. Science 330: 208-211.
Vancouver Sun. 2013. CN’s crude oil-by-rail plan increases risk of spills, fatalities, say opponents. Vancouver Sun
February 1, 2013. http://www.canada.com/story.html?id=bed39d62-7e95-498b-be36-b8f8e0913089
Vandermeulen, J.H. and J.G. Singh. 1994. Arrow oil spill, 1970-90: persistence of 20-yr weathered Bunker C fuel
oil. Canadian Journal of Fisheries and Aquatic Sciences 51: 845-855.
Venn-Watson S, Colegrove KM, Litz J, Kinsel M, Terio K, Saliki J, et al. (2015) Adrenal Gland and Lung Lesions
in Gulf of Mexico Common Bottlenose Dolphins (Tursiops truncatus) Found Dead following the
Deepwater Horizon Oil Spill. PLoS ONE 10(5): e0126538. doi:10.1371/journal.pone.0126538
Wade, T.L., Sericano, J.L,, Sweet, S., Knap, A.H. and N.L. Guinasso, Jr. 2016. Spatial and temporal distribution of
water column total polycyclic aromatic hydrocarbons (pah) and total petroleum hydrocarbons (tph) from
the Deepwater Horizon (Macondo) incident. Marine Pollution Bulletin (In Press). Manuscript Number:
MPB-D-15-00523R1
Wang, Z., M.F. Fingas and G. Sergy. 1994. Study of 22-Year-Old ARROW Oil Samples Using Biomarker
Compounds by GC/MS. Environmental Science and Technology, 28:1733-1746.
Water Survey of Canada. 2015. Discharge Graph for Pine River at East Pine (07FB001).
https://wateroffice.ec.gc.ca/report/report_e.html?mode=Graph&type=h2oArc&stn=07FB001&dataType=D
aily&parameterType=Flow&year=2012&y1Max=1&y1Min=1&y1Mean=1&scale=normal
Weinberg, A. 1972. Science and trans-science. Science 21:209-22.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 480
Wernick, B., deBruyn, A.H., Patterson, L. and P.Chapman. 2009. Effects of an oil spill on the regrowth of emergent
vegetation in a Northern Alberta lake. Archives of Environmental Contamination and Toxicology 57: 697-
706.
White, H.K, Hsing, P-Y., Cho, W., Shank, T.N., Cordes, E.E., Quattrin, A.M., Nelson, R.K., Camilli, R.,
Demopoiulos, A.W.J., German, C.E., Brooks, J.M., Roberts, H.H., Shedd, W., Reddy, C.M. and C.R.
Fisher. 2012. Impacts of the Deepwater Horizon oil spill on a deep-water coral community in the Gulf of
Mexico. Proceedings of the National Academy of Sciences 109: 20303-20308.
Wolfe DA, Hameedi MJ, Galt JA, Watabaytashi G, Short J, O’Clair C, Rice S, Michel J, Payne JR, Braddock J,
Hanna S, Sale D. 1994. The fate of the oil spilled from the Exxon Valdez. Environ Sci Technol 28:561–
568.
WSP and SL Ross. 2014a. Risk Assessment for Marine Spill in Canadian Waters. Phase 1: Oil Spills South of 60th
Parallel. Final Study Report prepared for Transport Canada. http://wcel.org/sites/default/files/file-
downloads/131-17593-00_ERA_Oil-Spill-South_150116_pp1-124.pdf.
WSP and SL Ross. 2014b. Risk Assessment for Marine Spills in Canadian Waters. Phase 2, Part B: Spills of Oil and
Select Hazardous and Noxious Substances (HNS) Transported in Bulk North of the 60th Parallel North.
Final Report prepared for Transport Canada.

Appendix A

ASTM (American Society of Testing and Materials). 2011. Method D7678 – 11. Standard Test Method for Total
Petroleum Hydrocarbons (TPH) in Water and Wastewater with Solvent Extraction Using Mid-IR Laser
Spectroscopy. DOI: 10.1520/D7678-11. http://www.astm.org/Standards/D7678.htm.
Brown, C. W., Lynch, P. F. and M. Ahmadjian. 1975. Applications of infrared spectroscopy in petroleum analysis
and oil spill identification. Applied Spectroscopy Reviews 9: 223-248.
Butt, J.A., Duckworth, D. and S.G. Perry. 1986. Characterization of Spilled Oil Samples: Purpose, Sampling,
Analysis and Interpretation. Institute of Petroleum. John Wiley & Sons, London.
de Andrade, D. F., Fernandes, D. R. and J.L. Miranda. 2010. Methods for the determination of conjugated dienes in
petroleum products: A review. Fuel 89: 1796-1805.
Eide, I. and K. Zahlsen. 2005. A novel method for chemical fingerprinting of oil and petroleum products based on
electrospray mass spectrometry and chemometrics. Energy and Fuels 19: 964-967.
Fernández-Varela, R., Gómez-Carracedo, M. P., Fresco-Rivera, P., Andrade, J. M., Muniategui, S. and D. Prada.
2006. Monitoring photooxidation of the Prestige's oil spill by attenuated total reflectance infrared
spectroscopy. Talanta 69: 409-417.
Fish, R. H., Komlenic, J. J. and B.K. Wines. 1984. Characterization and comparison of vanadyl and nickel
compounds in heavy crude petroleums and asphaltenes by reverse-phase and size-exclusion liquid
chromatography/graphite furnace atomic absorption spectrometry. Analytical Chemistry 56: 2452-2460.
Gaines, R. B., Frysinger, G. S.; Hendrick-Smith, M. S. and K.D. Stuart. 1999. Oil spill source identification by
comprehensive two-dimensional gas chromatography. Environmental Science & Technology 33: 2106-
2112.
Gautam, K., Jin, X. and M. Hansen. 1998. Review of spectrometric techniques for the characterization of crude oil
and petroleum products. Applied Spectroscopy Reviews 33: 427-443.
Kenkel, J. 2002. Analytical Chemistry for Technicians. CRC Press, London.
Khanmohammadi, M., Garmarudi, A.B. and M. de la Guardia. 2012. Characterization of petroleum-based products
by infrared spectroscopy and chemometrics. TrAC Trends in Analytical Chemistry 35: 135–149
Kim, M., Hong, S. H., Won, J., Yim, U. H., Jung, J. H., Ha, S. Y., An, J. G., Joo, C., Kim, E., Han, G. M., Baek, S.,
Choi, H. W. and W. J. Shim. 2013. Petroleum hydrocarbon contaminations in the intertidal seawater after
the Hebei Spirit oil spill – Effect of tidal cycle on the TPH concentrations and the chromatographic
characterization of seawater extracts. Water Research 47: 758-768.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 481
Li, K. and M. Fingas. 1997. An Evaluation of Bench Top Liquid Chromatography/Solid Probe Mass Spectrometry
for Emergency Spill Sample Analysis. Environment Canada, Ottawa.
Lynch, P. F. and C.W. Brown. 1973. Identifying source of petroleum by infrared spectroscopy. Environmental
Science & Technology 7: 1123-1127.
Pasadakis, N., Gaganis, V. and N. Varotsis. 2001. Accurate determination of aromatic groups in heavy petroleum
fractions using HPLC-UV-DAD. Fuel 80: 147-153.
Payne, J.R., Driskell, W.B., Short, J.W. and M.L. Larsen. 2008. Long term monitoring for oil in the Exxon Valdez
spill region. Marine Pollution Bulletin 56: 2067-2081.
Rogers, R.P. and A.M. McKenna. 2011. Petroleum analysis. Analytical Chemistry. 83: 4665-4687.
Rostad, C. E. 2006. Differentiation of commercial fuels based on polar components using negative electrospray
ionization/mass spectrometry. Environmental Forensics 7: 5-14.
Roussis, S.G. and W.P. Fitzgerald. 2000. Gas chromatographic simulated distillation-mass spectrometry for the
determination of the boiling point distributions of crude oils. Analytical Chemistry 72: 1400-1409.
Simanzhenkov, V. and R. Idem. 2003. Crude Oil Chemistry. Marcel Dekker, New York.
Speight, J. G. and J. Speight. 2002. Handbook of Petroleum Product Analysis. Wiley-Interscience, New Jersey.
Steffens, J., Landulfo, E., Courrol, L.C. and R. Guardani. 2011. Application of fluorescence to the study of crude
petroleum. Journal of Fluorescence 21: 859-864.
Sun, P. Y., Bao, M. T., Li, G. M., Wang, X. P., Zhao, Y. H., Zhou, Q. and L.X. Cao. 2009. Fingerprinting and
source identification of an oil spill in China Bohai Sea by gas chromatography-flame ionization detection
and gas chromatography-mass spectrometry coupled with multi-statistical analyses. Journal of
Chromatography A 1216: 830-836.
Tzing, S. H., Chang, J. Y., Ghule, A., Chang, J. J., Lo, B. and Y.C. Ling. 2003. A simple and rapid method for
identifying the source of spilled oil using an electronic nose: confirmation by gas chromatography with
mass spectrometry. Rapid Communications in Mass Spectrometry 17: 1873-1880.
Wang, Z., Li, K., Fingas, M., Sigouin, L. and L. Ménard. 2002. Characterization and source identification of
hydrocarbons in water samples using multiple analytical techniques. Journal of Chromatography A 971:
173-184.
Yim, U. H., Ha, S. Y., An, J. G., Won, J. H., Han, G. M., Hong, S. H., Kim, M., Jung, J. H. and W.J. Shim. 2011.
Fingerprint and weathering characteristics of stranded oils after the Hebei Spirit oil spill. Journal of
Hazardous Materials 197: 60-69.

Appendix B

Aske, N., Kallevik, H. and J. Sjöblom. 2002. Water-in-crude oil emulsion stability studied by critical electric field
measurements. Correlation to physico-chemical parameters and near-infrared spectroscopy. Journal of
Petroleum Science and Engineering 36: 1-17.
ASTM (The American Society for Testing and Materials). 2005. ASTM D5002-99: Standard Test Method for
Density and Relative Density of Crude Oils by Digital Density Analyzer. DOI: 10.1520/D5002-99R05.
ASTM. 2007. ASTM D1310 – 01: Standard Test Method for Flash Point and Fire Point of Liquids by Tag Open-
Cup Apparatus. DOI: 10.1520/D1310-01R07.
ASTM. 2012a. ASTM D97 – 12: Standard Test Method for Pour Point of Petroleum Products. DOI:
10.1520/D0097-12.
ASTM. 2012b. ASTM D287 – 12: Standard Test Method for API Gravity of Crude Petroleum and Petroleum
Products (Hydrometer Method). DOI: 10.1520/D0287-12B.
ASTM. 2013. ASTM D7566 - 13 Standard Specification for Aviation Turbine Fuel Containing Synthesized
Hydrocarbons. DOI: 10.1520/D7566-13
ASTM. 2015. ASTM D445-15 Standard Test Method for Kinematic Viscosity of Transparent and Opaque Liquids
(and Calculation of Dynamic Viscosity). DOI: 10.1520/D0445-15

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 482
Beal, C. 1946. The viscosity of air, water, natural gas, crude oil and its associated gases at oil field temperatures and
pressures. Transactions of the AIME 165: 94-115.
Beggs, H. D. and J. Robinson, J. 1975. Estimating the viscosity of crude oil systems. Journal of Petroleum
Technology 27: 1140-1141.
du Noüy, P. L. 1925. An interfacial tensiometer for universal use. The Journal of General Physiology 7: 625-631.
Fingas, M., Hollebone, B. and B. Fieldhouse. 2006. The density behaviour of heavy oils in freshwater: The example
of the Lake Wabamun spill. In: Proceedings of the 29th Arctic and Marine Oilspill Program. Environment
Canada, Ottawa, ON. pp. 57-77.
Fingas, M. F. 2010. Review of the North Slope Oil Properties Relevant to Environmental Assessment and
Prediction. Report prepared for Prince William Sound Regional Citizens' Advisory Council (PWSRCAC).
Anchorage, Alaska.
Lambert, P. 2003. A literature review of portable fluorescence-based oil-in-water monitors. Journal of Hazardous
Materials 102(1): 39-55.
Morrison, R. D. and B.L. Murphy. 2006. Environmental Forensics: Contaminant Specific Guide. Academic Press.

Appendix C

American Fuel & Petrochemical Manufacturers. 2014. A Survey of Bakken Crude Oil Characteristics Assembled for
the U.S. Department of Transportation. http://www.ourenergypolicy.org/wp-
content/uploads/2015/06/Survey-of-Crude-Oil-Characteristics_FINAL-1.pdf (accessed June, 2015)
BP (British Petroleum) Gulf Science Data MC-252 Oil Characterization Data File. Reference No. O-01v01-01. Last
modified November 12, 2013. http://gulfsciencedata.bp.com/go/doctype/6145/178706 (accessed June,
2015)
Crude Quality Inc. 2015. crudemonitor.ca. http://crudemonitor.ca. (accessed June 2015)
Enbridge Pipelines Inc. 2014 Crude Characteristics No. 45. Enbridge Energy Partners, L.P.
http://www.enbridge.com/~/media/www/Site%20Documents/Delivering%20Energy/2014%20Mainline%20
Crude%20Characteristics.pdf?la=en
Environmental Technology Centre, Environment Canada. 2015. Oil Properties Database. http://www.etc-
cte.ec.gc.ca/databases/oilproperties/. (accessed June 2015)
Hollebone, B.P. 2015. Appendix A. The oil properties data appendix. In: M.F. Fingas, M.F.( ed.) Handbook of Oil
Spill Science and Technology. Wiley & Sons, Inc., Hoboken, NJ.
Indian and Northern Affairs Canada. 1995. Petroleum Exploration in Canada. Chapter 4 – Mackenzie Delta and
Beauford Sea. http://www.aadnc-aandc.gc.ca/eng/1321378288752/1321378450630 (accessed June 2015)
MSDS product data sheets https://www.msdsonline.com/msds-search
North Dakota Petroleum Council. No date. Bakken Crude Properies. http://www.ndoil.org/resources/bkn/. (accessed
June 2015)
National Research Council. 2013. TRB Special Report 311: Effects of Diluted Bitumen on Crude Oil Transmission
Pipelines. http://www.nap.edu/read/18381/chapter/1 (accessed Oct 2015)
Wikipedia. 2015. Western Canadian Select. http://en.wikipedia.org/wiki/Western_Canadian_Select. (accessed June
2015)

Appendix D

Adams, G.G., Klerks, P.L., Belanger, S.E. and D. Dantin. 1999. The effect of the oil dispersant Omni-Clean® on the
toxicity of fuel oil no. 2 in two bioassays with the sheepshead minnow Cyprinodon variegatus.
Chemosphere 39: 2141-2157.
Adams, J., Sweezey, M., and P.V. Hodson. 2014a. Oil and oil dispersant do not cause synergistic toxicity to fish
embryos. Environmental Toxicology and Chemistry 33: 825-835.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 483
Adams, J.E., Munno, K., Bornstein, J., King, T., Brown, R.S., Hollebone, B.P. and P.V. Hodson. 2014b.
Identification of compounds in heavy fuel oil that are chronically toxic to rainbow trout embryos through
effects-driven chemical fractionation. Environmental Toxicology and Chemistry 33: 825-835.
Al-Ayed, M.I. 2001. Effect of crude oil on some hematological parameters of the freshwater fishes, Oreochromis
niloticus. Saudi Journal of Biological Sciences 8: 26-40.
Akaishi, F., de Assis, H., Jakobi, S., Eiras-Stofella, D., St-Jean, S., Courtenay, S., Lima, E., Wagener, A., Scofield,
A. and C. Ribeiro. 2004. Morphological and neurotoxicological findings in tropical freshwater fish
(Astyanax sp.) after waterborne and acute exposure to water soluble fraction (WSF) of crude oil. Archives
of Environmental Contamination Toxicology 46: 244-253.
Amat, A., Burgeot, T., Castegnaro, M. and A. Pfohl-Leszkowicz. 2006. DNA adducts in fish following an oil spill
exposure. Environmental Chemistry Letters 4: 93-99.
Anderson, J.W., Neff, J.M., Cox, B.A., Tatem, H.E., and Hightower, G.M. 1974. Characteristics of dispersions and
water-soluble extracts of crude and refined oils and their toxicity to estuarine crustaceans and fish. Marine
Biology 27: 75-88.
Balk, L., Hylland, K., Hansson, T., Berntssen, M.H.G., Beyer, J., Jonsson, G., Melbye, A., Grung, M., Torstensen,
B.E, Borseth, J.F., Skarphedinsdottir, H. And J. Klungsoyr. 2011. Biomarkers in natural fish populations
indicate adverse biological effects of offshore oil production. PLoS ONE 6: 1-10.
Barber, W.E., McDonald, L.L., Erickson, W.P. and M Vallarino. 1995. Effect of the Exxon Valdez oil spill on
intertidal fish - A field study. Transactions of the American Fisheries Society 124: 461-476.
Barron, M., Carls, M., Short, J. and S. Rice. 2003. Photoenhanced toxicity of aqueous phase and chemically
dispersed weathered Alaska North Slope crude oil to Pacific herring eggs and larvae. Environmental
Toxicology and Chemistry 22: 650-660.
Bhattacharyya, S., Klerks, P. And J. Nyman. 2003. Toxicity to freshwater organisms from oils and oil spill chemical
treatments in laboratory microcosms. Environmental Pollution 122: 205-215.
Bilbao, E., Raingeard, D., de Cerio, O.D., Ortiz-Zarragoitia, M., Ruiz, P., Izagirre, U., Orbea, A., Marigomez, I.,
Cajaraville, M.P. and I. Cancio. 2010. Effects of exposure to prestige-like heavy fuel oil and to
perfluorooctane sulfonate on conventional biomarkers and target gene transcription in the thicklip grey
mullet Chelon labrosus. Aquatic Toxicology 98: 282-96.
Birtwell, I.K., Fink, R., Brand, D., Alexander, R., and C.D. McAllister. 1999. Survival of pink salmon
(Oncorhynchus gorbuscha) fry to adulthood following a 10-day exposure to the aromatic hydrocarbon
water-soluble fraction of crude oil and release to the Pacific Ocean. Canadian Journal of Fisheries and
Aquatic Sciences 56: 2087-2098.
Boudreau, M., Sweezey, M., Lee, K., Hodson, P.V. and S.C. Courtenay. 2009. Toxicity of Orimulsion-400 to early
life stages of Atlantic herring (Clupea harengus) and mummichog (Fundulus heteroclitus). Environmental
Toxicology and Chemistry 28: 1206-1217.
Brand, D.G., Fink, R., Bengeyfield, W., Birtwell, I.K. and C.D. McAllister. 2001. Salt water-acclimated pink
salmon fry (Oncorhynchus gorbuscha) develop stress-related visceral lesions after 10-day exposure to
sublethal concentrations of the water-soluble fraction of North Slope crude oil. Toxicologic Pathology 29:
574-584.
Brannon, E.L., Collins, K.M., Brown, J.S., Neff, J.M., Parker, K.R. and W.A. Stubblefield. 2006a. Toxicity of
weathered Exxon Valdez crude oil to pink salmon embryos. Environmental Toxicology and Chemistry 25:
962-972.
Brannon, E.L., Maki, A.W., Moulton, L.L., and P.R. Parker. 2006. Results from a sixteen year study on the effects
of oiling from the Exxon Valdez on adult pink salmon returns. Marine Pollution Bulletin 52(8): 892-899.
Carls, M.G., Heintz, R.A., Marty, G.D., and R.D. Rice. 2005. Cytochrome P4501A induction in oil-exposed pink
salmon Oncorhynchus gorbuscha embryos predicts reduced survival potential. Marine Ecology Progress
Series 301: 235-265.
Carls, M.G., Holland, L., Larsen, M., Collier, T.K., Scholz, N.L. and J.P. Incardona. 2008. Fish embryos are
damaged by dissolved PAHs, not oil particles. Aquatic Toxicology 88(2): 121-127.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 484
Carls, M.G., Holland, L., Larsen, M., Lum, J., Mortensen, D., Wang, S. and A. Wertheimer. 1996. Growth, feeding,
and survival of pink salmon fry exposed to food contaminated with crude oil. American Fisheries Society
Symposium 18: 608-618.
Carls, M.G., Hose, J., Thomas, R. and S. Rice. 2000. Exposure of Pacific herring to weathered crude oil: Assessing
effects on ova. Environmental Toxicology and Chemistry 19: 1649-1659.
Carls, M.G., Marty, G,D., Meyers, T.R,, Thomas, R.E. and S.D. Rice. 1998. Expression of viral hemorrhagic
septicemia virus in pre-spawning pacific herring (Clupea pallasi) exposed to weathered crude oil. Canadian
Journal of Fisheries and Aquatic Sciences 55: 2300-2309.
Carls, M.G., Rice, S. And J. Hose. 1999. Sensitivity of fish embryos to weathered crude oil: Part I. low-level
exposure during incubation causes malformations, genetic damage, and mortality in larval Pacific herring
(Clupea pallasi). Environmental Toxicology and Chemistry 18: 481-493.
Colavecchia, M.V., Backus, S.M., Hodson, P.V. and J.L. Parrott. 2004. Toxicity of oil sands to early life stages of
fathead minnows (Pimephales promelas). Environmental Toxicology Chemistry 23: 1709-1718.
Colavecchia, M.V., Hodson, P.V., and J.L. Parrott. 2006. CYP1A induction and blue sac disease on early life stages
of white suckers (Catostomus commersoni) exposed to oil sands. Journal of Toxicology and Environmental
Health Part A 69: 1-28.
Colavecchia, M.V., Hodson, P.V. and J.L. Parrott. 2007. The relationships among CYP1A induction, toxicity and
eye pathology in early life stages of fish exposed to oil sands. Journal of Toxicology and Environmental
Health 70: 1542-1555.
Couillard, C.M. 2002. A microscale test to measure petroleum oil toxicity to mummichog embryos. Environmental
Toxicology 17: 195-202.
Couillard, C.M., Lee, K., Legare, B. and T.L. King. 2005. Effect of dispersant on the composition of the water-
accommodated fraction of crude oil and its toxicity to larval marine fish. Environmental Toxicology and
Chemistry 24(6): 1496-1504.
Debruyn, A.M.H., Wernick, B.G., Stefura, C., McDonald, B.G., Rudolph, B.-L., Patterson, L. and P.M. Chapman.
2007. In situ experimental assessment of lake whitefish development following a freshwater oil spill.
Environmental Science and Technology 41: 6983 -6989.
Deér, A., Henczová, M., Banka, L., Varanka, Z., and J. Nemcsók. 2010. Effects of crude oil and oil fractions on the
liver P450-dependent monooxygenase activities and antioxidant defence system of different freshwater fish
species. Acta Biologica Hungarica 61: 262-
Fuller, C., Bonner, J., Page, C., Ernest, A., McDonald, T. and S. McDonald. 2004. Comparative toxicity of oil,
dispersant, and oil plus dispersant to several marine species. Environmental Toxicology and Chemistry 23:
2941-2949.
Gulec, I. And D. Holdway. 2000. Toxicity of crude oil and dispersed crude oil to ghost shrimp Palaemon serenus
and larvae of Australian bass Macquaria novemaculeata. Environmental Toxicology 15: 91-98.
Heintz, R.A., Rice, S.D., Wertheimer, A.C., Bradshaw, R.F., Thrower, F.P., Joyce, J.E., and J.W. Short. 2000.
Delayed effects on growth and marine survival of pink salmon Oncorhynchus gorbuscha after exposure to
crude oil during embryonic development. Marine Ecology Progress Series 208: 205-216.
Heintz, R.A., Short, J.W. and S.D. Rice. 1999. Sensitivity of fish embryos to weathered crude oil: Part II. increased
mortality of pink salmon (Oncorhynchus gorbuscha) embryos incubating downstream from weathered
Exxon Valdez crude oil. Environmental Toxicology and Chemistry 18: 494-503.
Hicken, C.E., Linbo, T.L., Baldwin, D.H., Willis, M.L., Myers, M.S., Holland, L., Larsen, M., Stekoll, M.S., Rice,
S.D., Collier, T.K., Scholz, N.L. and J.P. Incardona. 2011. Sublethal exposure to crude oil during
embryonic development alters cardiac morphology and reduces aerobic capacity in adult fish. PNAS
108(17): 7086-7090.
Hose, J.E., McGurk, D., Marty, G.D., Hinton, D.E., Brown, E.D. and T.T. Baker. 1996. Sublethal effects of the
Exxon Valdez oil spill on herring embryos and larvae: Morphological, cytogenetic, and histopathological
assessments, 1989-1991. Canadian Journal of Fisheries and Aquatic Sciences 53: 2355-2365.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 485
Huovinen, P.S., Soimasuo, M.R., and Oikari, A.O.J. 2001. Photoinduced toxicity of retene to Daphnia magna under
enhanced UV-B radiation. Chemosphere 45:683-691.
Incardona, J.P., Carls, M.G., Day, H.L., Sloan, C.A., Bolton, J.L., Collier, T.K. and N.L. Scholz. 2009. Cardiac
arrhythmia is the primary response of embryonic Pacific herring (Clupea pallasi) exposed to crude oil
during weathering. Environmental Science &Technology 43: 201-207.
Incardona, J.P., Vines, C.A., Linbo, T.L., Myers, M.S., Sloan, C.A., Anulacion, B.F., Boyd, D., Collier, T.K.,
Morgan, S., Cherr, G.N. and N.L. Scholz. 2012b. Potent phototoxicity of marine bunker oil to translucent
herring embryos after prolonged weathering. PLoS ONE 7(2): e30116.
Jung, J., Kim, M., Yim, U.H., Ha, S.Y., An, J.G., Won, J.H., Han, G.M., Kim, N.S., Addison, R.F., and W.J. Shim.
2011. Biomarker responses in pelagic and benthic fish over 1 year following the Hebei Spirit oil spill
(Taean, Korea). Marine Pollution Bulletin 62: 1859-66.
Kazlauskiene, N., Svecevicius, G., Vosyliene, M., Marciulioniene, D. and D. Montvydiene. 2004. Comparative
study on sensitivity of higher plants and fish to heavy fuel oil. Environmental Toxicology 19: 449-51.
Kazlauskiene, N., Vosyliene, M.Z. and E. Ratkelyte. 2008. The comparative study of the overall effect of crude oil
on fish in early stages of development. NATO Science for Peace and Security 1: 307-316.
Khan, R.A. 1995. Histopathology in winter flounder, Pleuronectes americanus, following chronic exposure to crude
oil. Bulletin of Environmental Contamination and Toxicology 54: 297-301.
Khan, R.A. 2012. Effects of polycyclic aromatic hydrocarbons on sexual maturity of Atlantic cod, Gadus morhua,
following chronic exposure. Environment and Pollution 2(1): 1-10.
Kocan, R.M., Hose, J.E., Brown, E.D., and T.T. Baker. 1996a. Pacific herring (Clupea pallasi>) embryo sensitivity
to Prudhoe Bay petroleum hydrocarbons: laboratory evaluation and in situ exposure at oiled and unoiled
sites in Prince William Sound. Canadian Journal of Fisheries and Aquatic Sciences 53: 2366-2375.
Kocan, R.M., Marty, G.D., Okihiro, M.S., Brown, E.D. and T.T. Baker. 1996b. Reproductive success and
histopathology of individual Prince William Sound Pacific herring 3 years after the Exxon Valdez oil spill.
Canadian Journal of Fisheries and Aquatic Sciences 53: 2388-2393.
Koyama, J. and A. Kakuno. 2004. Toxicity of heavy fuel oil, dispersant, and oil-dispersant mixtures to a marine fish,
Pagrus major. Fisheries Science 70: 587-594.
Le Bihanic, F., Morin, B., Cousin, X., Le Menach, K., Budzinski, H., and J. Cachot. 2014. Developmental toxicity
of PAH mixtures in fish early life stages. Part I: adverse effects in rainbow trout. Environmental Science
and Pollution Research 21: 13720-13731.
Lee, K., R.C. Prince, R.C., Greer, C.W., Doe, K.G., Wilson, J.E.H., Cobanli, S.E., Wohlgeschaffen, G.D., Alroumi,
D., King. T. and G.H. Tremblay. 2003. Composition and toxicity of residual Bunker C fuel oil in intertidal
sediments after 30 years. Spill Science and Technology Bulletin 8: 187-199.
Little, E., Cleveland, L., Calfee, R. And M. Barron. 2000. Assessment of the photoenhanced toxicity of a weathered
oil to the tidewater silverside. Environmental Toxicology and Chemistry 19: 926-932.
Liu, B., Romaire, R., Delaune, R. and C. Lindau. 2006. Field investigation on the toxicity of Alaska North Slope
crude oil (ANSC) and dispersed ANSC crude to gulf killifish, eastern oyster and white shrimp.
Chemosphere 62: 520-526.
Lockhart, W.L., Duncan, D.A., Billeck, B.N., Danell, R.A. and M.J. Ryan. 1996. Chronic toxicity of the ‘water-
soluble fraction’ of Norman Wells crude oil to juvenile fish. Spill Science and Technology Bulletin 3: 259-
262.
Lotufo, G.R. and Fleeger, J.W. 1997. Effects of sediment-associated phenanthrene on survival, development and
reproduction of two species of meiobenthic copepods. Marine Ecology Progress Series 151: 91-102
Martin, J.D., Adams, A., Hollebone, B.P., King, T., Brown, R.S., and P.V. Hodson. 2014. Chronic toxicity of heavy
fuel oils to fish embryos using multiple exposure scenarios. Environmental Toxicology and Chemistry 33:
677-687.
Marty, G.D., Hose, J.E., McGurk, M.D., Brown, E.D., and D.E. Hinton. 1997a Histopathology and cytogenetic
evaluation of Pacific herring larvae exposed to petroleum hydrocarbons in the laboratory or in Prince

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 486
William Sound, Alaska, after the Exxon Valdez oil spill. Canadian Journal of Fisheries and Aquatic
Sciences 54(8): 1846-1857.
Marty, G.D., Okihiro, M.S., Brown, E.D., Hanes, D. and D.E. Hinton. 1999. Histopathology of adult Pacific herring
in Prince William Sound, Alaska, after the Exxon Valdez oil spill. Canadian Journal of Fisheries and
Aquatic Sciences 56: 419-426.
Marty, G.D., Short, J.W., Dambach, D.M., Willits, N.H., Heintz, R.A., Rice, S.D., Stegeman, J.J. and D.E. Hinton.
1997b. Ascites, premature emergence, increased gonadal cell apoptosis, and cytochrome P4501A induction
in pink salmon larvae continuously exposed to oil-contaminated gravel during development. Canadian
Journal of Zoology 75: 989-1007.
McGurk, M.D. and E.D. Brown. 1996. Egg-larval mortality of Pacific herring in Prince William Sound, Alaska,
after the Exxon Valdez oil spill. Canadian Journal of Fisheries and Aquatic Sciences 53: 2343-2354.
McIntosh, S., King, T., Wu, D. and P.V. Hodson. 2010. Toxicity of dispersed crude oil to early life stages of
Atlantic herring (Clupea harengus). Environmental Toxicology and Chemistry 29: 1160-1167.
Middaugh, D.P., Chapman, P.J., Shelton, M.E., McKenney, C.L. and L.A. Courtney. 2002. Effects of fractions from
biodegraded Alaska North Slope crude oil on embryonic inland silversides, Menidia beryllina. Archives of
Environmental Contamination and Toxicology 42: 236-243.
Middaugh, D.P., Chapman, P.J. and M.E. Shelton. 1996. Responses of embryonic and larval inland silversides,
Menidia beryllina, to a water-soluble fraction formed during biodegradation of artificially weathered
Alaska North Slope crude oil. Archives of Environmental Contamination and Toxicology 31: 410-419.
Milinkovitch, T., Kanan, R., Thomas-Guyon, H. and S. Le Floch. 2011a. Effects of dispersed oil exposure on the
bioaccumulation of polycyclic aromatic hydrocarbons and the mortality of juvenile Liza ramada. Science of
the Total Environment 409: 1643-1650.
Milinkovitch, T., Ndiaye, A., Sanchez, W., Le Floch, S. and H. Thomas-Guyon. 2011b. Liver antioxidant and
plasma immune responses in juvenile golden grey mullet (Liza aurata) exposed to dispersed crude oil.
Aquatic Toxicology 101: 155-64.
Moles A, and B.L. Norcross. 1998. Effects of oil-laden sediments on growth and health of juvenile flatfishes.
Canadian Journal of Fisheries and Aquatic Sciences 55: 605-610.
Nakayama, K., Kitamura, S., Murakami, Y., Song, J., Jung, S., Oh, M., Iwata, H. and S. Tanabe. 2008.
Toxicogenomic analysis of immune system-related genes in Japanese flounder (Paralichthys olivaceus)
exposed to heavy oil. Marine Pollution Bulletin 57: 445-452.
Navas, J.M., Babin, M., Casado, S., Fernandez, C. and J.V. Tarazona. 2006. The Prestige oil spill: A laboratory
study about the toxicity of the water-soluble fraction of the fuel oil. Marine Environmental Research 62:
352-355.
Ndimele, P.E., Jenyo-Oni, A. and C.C. Jibuike. 2010. Comparative toxicity of crude oil, dispersant and crude oil-
plus-dispersant to Tilapia guineensis. Research Journal of Environmental Toxicology 4: 13-22.Neff, J.,
Ostazeski, S., Gardiner, W. and I. Stejskal I. 2000. Effects of weathering on the toxicity of three offshore
Australian crude oils and a diesel fuel to marine animals. Environmental Toxicology and Chemistry 19:
1809-1821.
Norcross, B.L., Hose, J.E., Frandsen, M., and Brown, E.D. 1996. Distribution, abundance, morphological condition,
and cytogenetic abnormalities of larval herring in Prince William Sound, Alaska, following the Exxon
Valdez oil spill. Canadian Journal of Fisheries and Aquatic Science. 53: 2376-2387.
Pilcher, W., Miles, S., Tang, S., Mayer, G., and A. Whitehead. 2014. Genomic and genotoxic responses to
controlled weathered-oil exposures confirm and extend field studies on impacts of the Deepwater Horizon
oil spill on native killifish. PLoS ONE 9: e106351.
Pollino, C.A. and D.A. Holdway. 2002. Toxicity testing of crude oil and related compounds using early life stages of
the crimson-spotted rainbowfish (Melanotaenia fluviatilis). Ecotoxicology and Environmental Safety 52:
180-189.
Ramachandran, S.D., Hodson, P.V., Khan, C.W. and K Lee. 2004. Oil dispersant increases PAH uptake by fish
exposed to crude oil. Ecotoxicology and Environmental Safety 59: 300-308.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 487
Ramachandran, S.D., Sweezey, M.J., Hodson, P.V., Boudreau, M., Courtenay, S., King, T., Dixon, J.A. and K. Lee.
2006. Influence of salinity and fish species on PAH uptake from dispersed MESA crude oil. Marine
Pollution Bulletin 52: 1182-1189.
Schein, A., Scott, J.A., Mos, L. and P.V. Hodson. 2009. Oil dispersion increases the apparent bioavailability and
toxicity of diesel to rainbow trout (Oncorhynchus mykiss). Environmental Toxicology and Chemistry 28:
595-602.
Seilor, T.-B., Best, N., Fernqvist, M.M.l., Hercht, H., Smith, K.E.C., Braunbeck, T., Mayer, P., and Hollert, H. 2014.
PAH toxicity at aqueous solubility in the fish embryo test with Danio rerio using passive dosing.
Chemosphere 112: 77-84.
Shen, A., Tang, F. And X. Shen. 2010. Toxicity of crude oil and fuel oil on the embryonic development of the large
yellow croaker (Larmichthys crocea). In: Proceedings of 3rd International Conference on Biomedical
Engineering and informatics, Oct 16-18, 2010. Yantai, China.
Shukla, P., Gopalani, M., Ramteke, D.S. and S.R. Wate. 2007. Influence of salinity on PAH uptake from water
soluble fraction of crude oil in Tilapia mossambica. Bulletin of Environmental Contamination and
Toxicology 79: 601-605.
Sol, S.Y., Johnson, L.L., Horness, B.H. and T.K. Collier TK. 2000. Relationship between oil exposure and
reproductive parameters in fish collected following the Exxon Valdez oil spill. Marine Pollution Bulletin
40: 1139-1147.
Thomas, P. and L. Budiantara. 1995. Reproductive life-history stages sensitive to oil and naphthalene in Atlantic
croaker. Marine Environmental Research 39: 147-50.
Thomas, R.E., Carls, M.G., Rice, S.D. and L. Shagrun. 1997. Mixed function oxygenase induction in pre- and post-
spawn herring (Clupea pallasi) by petroleum hydrocarbons. Comparative Biochemistry and Physiology Part
C: Toxicology 116: 141-147.
Vignet, C., Le Menach, K., Lyphout, L., Guionnet, T., Frère, L., Leguay, D., Budzinski, H., Cousin, X., and M.L.
Bégout. 2014b. Chronic dietary exposure to pyrolytic and petrogenic mixtures of PAHs causes
physiological disruption in zebrafish – part II: behavior. Environmental Science and Pollution Research 21:
13818-13832.
Vignet, C., Le Menach, K., Mazurais, D., Lucas, J., Perrichon, P., Le Bihanic, F., Devier, M.H., Lyphout, L., Frère,
L., Bégout, M.L., Zambonino-Infante, J.L., Budzinski, H., and X. Cousin. 2014a. Chronic dietary exposure
to pyrolytic and petrogenic mixtures of PAHs causes physiological disruption in zebrafish – part I: survival
and growth. Environmental Science and Pollution Research 21: 13804-13817.
Vosyliene, M., Kazlauskiene, N. And K. Joksas. 2005. Toxic effects of crude oil combined with oil cleaner simple
green on yolk-sac larvae and adult rainbow trout Oncorhynchus mykiss. Environmental Science and
Pollution Research 12: 136-139.
Wang, Y., Zhou, Q., Peng, S., Ma, L. and X. Niu. 2009. Toxic effects of crude-oil-contaminated soil in aquatic
environment on Carassius auratus and their hepatic antioxidant defense system. Journal of Environmental
Sciences 21: 612-617.
Woodin, B.R., Smolowitz, R.M. and J.J. Stegeman. 1997. Induction of cytochrome P4501A in the intertidal fish
Anoplarchus purpurescens by Prudhoe Bay crude oil and environmental induction in fish from Prince
William Sound. Environmental Science and Technology 31: 1198-1205.
Wu, D., Wang, Z., Hollebone, B., McIntosh, S., King, T., and P.V. Hodson. 2012. Comparative toxicity of four
chemically-dispersed and undispersed crude oils to rainbow trout embryos. Environmental Toxicology and
Chemistry 31: 754-765.

282 Somerset Street West, Ottawa ON, K2P 0J6 • Tel: 613-991-6990 • www.rsc-src.ca | 488

You might also like