You are on page 1of 267

State of Knowledge Review of Fate and Effect of Oil

in the Arctic Marine Environment




2011


A report prepared for the National Energy Board of Canada

by


K. Lee
1
, M. Boudreau
2
, J. Bugden
1
, L. Burridge
3
, S.E. Cobanli
1
, S. Courtenay
2
,
S. Grenon
4
, B. Hollebone
5
, P. Kepkay
1
, Z. Li
1
, M. Lyons
2
, H. Niu
1
, T.L. King
1
,
S. MacDonald
5
, E.C. McIntyre
1
, B. Robinson
1
, S.A. Ryan
1
and G. Wohlgeschaffen
1







1
Centre for Offshore Oil, Gas and Energy Research (COOGER), Fisheries and Oceans Canada, Bedford
Institute of Oceanography, P.O. Box 1006, Dartmouth, Nova Scotia, Canada, B2Y 4A2
2
Fisheries and Oceans Canada, Gulf Fisheries Centre, P.O. Box 5030, Moncton, New Brunswick, E1C
9B6
3
Fisheries and Oceans Canada, St. Andrews Biological Station, 531 Brandy Cove Road, St. Andrews,
New Brunswick, Canada, E5B 2L9
4
Triox Environmental Emergencies, 4839 Garnier, Montral, Qubec, Canada, H2J 3S8
5
Emergencies Science and Technology Section, Environment Canada, 335 River Road, Ottawa, Ontario,
Canada, K1A 0H3


i
TABLE OF CONTENTS

List of Figures................................................................................................................................ iv
List of Tables .................................................................................................................................. v
List of Acronyms ........................................................................................................................... vi
Executive Summary........................................................................................................................ 1
1. Introduction................................................................................................................................. 4
1. Introduction................................................................................................................................. 4
2. Characterization and Classification of Crude Oil ....................................................................... 9
2.1 Physical and Chemical Properties......................................................................................... 9
2.2 Significance of Oil Properties in Oil Spill Response.......................................................... 12
2.3 Canadian Arctic Offshore Crude Oils................................................................................. 13
3. Oil Spills in Arctic Waters........................................................................................................ 17
3.1 International Concerns and Governance............................................................................. 17
3.2 Behaviour and Fate of Oil................................................................................................... 18
Oil in Ice-Free Waters .......................................................................................................... 19
Oil in Ice Covered Waters..................................................................................................... 22
3.3 Factors Influencing Oil Behaviour...................................................................................... 24
Spreading in Broken Ice........................................................................................................ 24
Movement on Ice................................................................................................................... 27
Movement Under Ice............................................................................................................. 27
Movement Through Ice ......................................................................................................... 29
Adsorption to Snow............................................................................................................... 31
3.4 Factors Influencing Oil Fate (Weathering) ......................................................................... 32
Evaporation........................................................................................................................... 32
Dissolution............................................................................................................................ 33
Dispersion............................................................................................................................. 34
Emulsification ....................................................................................................................... 35
Photo-Oxidation.................................................................................................................... 38
Biodegradation ..................................................................................................................... 39
Formation of Oil-Mineral Aggregates.................................................................................. 44
Sedimentation........................................................................................................................ 46
4. Oil Spills in Canada from Offshore Oil and Gas Activities...................................................... 47
4.1 Oil Types from Vessel Operations...................................................................................... 50
4.2 Oil Types from Oil Platform Operations ............................................................................ 51
Crude Oil from Drilling Activities ........................................................................................ 52
Petroleum Products used in Operations............................................................................... 52
4.3 Overview of Oil Spill Risks in the Canadian Arctic........................................................... 52
Defining Risk......................................................................................................................... 52
Incidents from Vessel Operation........................................................................................... 54
Incidents from Oil Platform Operations............................................................................... 57
4.4 Modelling Spill Scenarios and Oil Behaviour .................................................................... 62
Scenario for Spills of Crude Oil............................................................................................ 63
Scenario for Spills of Intermediate Fuel Oil......................................................................... 66
Scenario for Spills of Marine Gasoil .................................................................................... 67
Scenario for Spills of Aviation Fuel (Jet Fuel) ..................................................................... 69
Recommendations from Modelling Exercises....................................................................... 70
5. Offshore Arctic Oil Spill Response Options............................................................................. 72
5.1 Development of Policies, Regulations and Capacity.......................................................... 72
5.2 Mechanical Containment and Recovery............................................................................. 74
5.3 In situ Burning .................................................................................................................... 76
5.4 Chemical Dispersion........................................................................................................... 79
5.5 Oil-Mineral Aggregates ...................................................................................................... 82
5.6 Bioremediation.................................................................................................................... 85
5.7 Natural Attenuation............................................................................................................. 89
6. Biological Effects of Oil ........................................................................................................... 91
6.1 Oil Toxicity in the Arctic.................................................................................................... 91
6.2 Routes of Exposure............................................................................................................. 93
6.3 Bioaccumulation, Biomonitoring and Toxicity Assessment .............................................. 95
6.4 The Arctic Food Web.......................................................................................................... 98
6.5 Effects on Arctic Sea Ice Communities............................................................................ 100
6.6 Effects of Oil on Arctic Biota ........................................................................................... 103
Bacteria............................................................................................................................... 103
Phytoplankton and Macroalgae.......................................................................................... 104
Salt Marsh Vegetation......................................................................................................... 107
Zooplankton ........................................................................................................................ 108
Fish ..................................................................................................................................... 112
Benthic Invertebrates.......................................................................................................... 126
Mammals............................................................................................................................. 137
7. Arctic Oil Spill Field Trials .................................................................................................... 150
7.1 Balaena Bay experiment 1974 - 1975 (Norcor)................................................................ 150
Balaena Bay Revisited: 1981.............................................................................................. 152
7.2 Baffin Island Oil Spill Experiment (BIOS) ...................................................................... 154
Nearshore Study.................................................................................................................. 155
Shoreline Study ................................................................................................................... 158
BIOS Revisited .................................................................................................................... 160
7.3 Field Trials in Svalbard, Norway...................................................................................... 162
Early Field Studies on Oil Bioremediation......................................................................... 162
In Situ Treatment of Oiled Sediment Shorelines (ITOSS) Program................................... 162
7.4 Field Trial Projects on Arctic Oil Spills in Ice ................................................................. 165
7.5 Field Trials on Enhanced Oil Dispersion with Mineral Fines .......................................... 169
7.6 Monitoring Arctic Offshore Oil and Gas Operations ....................................................... 169
7.7 Future Oil Spill Field Trials in the Arctic......................................................................... 170
8. Operational Waste Discharges................................................................................................ 174
8.1 Produced Water................................................................................................................. 174
Chemical Composition........................................................................................................ 175
Petroleum Hydrocarbons.................................................................................................... 176
Environmental Concerns over Discharges ......................................................................... 177
Fate following Discharge into the Ocean........................................................................... 179
Environmental Effects of Discharges.................................................................................. 179
Effects on Water-Column Organisms ................................................................................. 180
Accumulation and Effects in Sediments .............................................................................. 180
ii
Aquatic Toxicity .................................................................................................................. 181
Bioaccumulation and Biomarkers as Evidence of Exposure .............................................. 183
Alteration of Trophic Level Dynamics by Produced Water................................................ 186
Ecological Risk of Produced Water Discharges................................................................. 187
Produced Water Treatment................................................................................................. 188
Environmental Effects Monitoring and Research Needs .................................................... 189
8.2 Drilling Muds.................................................................................................................... 191
Arctic Marine Food Webs and Toxicity.............................................................................. 193
9. Case Study: The Exxon Valdez Oil Spill................................................................................. 195
9.1 Fate of the Oil ................................................................................................................... 197
9.2 Effects of the Oil ............................................................................................................... 200
Bioavailability..................................................................................................................... 200
Shoreline Flora and Fauna................................................................................................. 200
Invertebrates ....................................................................................................................... 202
Fish ..................................................................................................................................... 202
Birds.................................................................................................................................... 203
Marine Mammals................................................................................................................ 204
9.3 Present Status of Injured Resources and Services ............................................................ 205
9.4 Lessons Learned and Issues to Consider .......................................................................... 208
10. Future Research Needs ......................................................................................................... 209
10.1 Oil Detection................................................................................................................... 209
10.2 Oil Fate and Behaviour ................................................................................................... 209
10.3 Biological Effects............................................................................................................ 210
10.4 Mechanical Recovery................................................................................................. 210
10.5 In Situ Burning................................................................................................................ 211
10.6 Enhanced Dispersion ...................................................................................................... 211
10.7 Biodegradation and Natural Attenuation ........................................................................ 211
10.8 Development of Predictive Models ................................................................................ 212
10.9 Field Trials...................................................................................................................... 212
11. Acknowledgements............................................................................................................... 213
References................................................................................................................................... 214
Appendix 1.................................................................................................................................. 257


iii
List of Figures

Figure 1 Since 1980 industry has increased the recovery of oil from the offshore sector to meet
global demand as land-based oil reserves declined. There is a greater risk of spills and
damage as operations expand into the frontier regions including the deep waters off the
continental shelf and the Arctic (Sandrea and Sandrea, 2007). .............................................. 4
Figure 2 The significance of the oil and gas industry to Canadas GDP in 2006........................... 5
Figure 3 Norman Wells crude oil dynamic viscosity as a function of temperature; data from
Environment Canadas Oil Properties database (Environment Canada, 2001a). ................. 11
Figure 4 Physical, chemical and biological processes affecting the fate and behaviour of spilled
oil (ITOPFL, 2002). .............................................................................................................. 20
Figure 5 Oil and ice interaction processes (Bobra and Fingas, 1986). ......................................... 23
Figure 6 Canadian Arctic (source: Solar Navigator, 2011, http://www.solarnavigator.net). ....... 47
Figure 7 Exploration parcels of Cairn Energy in the Greenland offshore (Cairn, 2011).............. 49
Figure 8 Accidents by ship type; OBO = other bulk operations (source: Transportation Safety
Board of Canada). ................................................................................................................. 54
Figure 9 Types of accidents (Source: Transportation Safety Board of Canada). ......................... 55
Figure 10 Canadian Arctic shipping routes. ................................................................................. 56
Figure 11 Number of spills per year in Nova Scotia from offshore operations............................ 58
Figure 12 Number of spills per year in Newfoundland-Labrador from offshore operations........ 58
Figure 13 Total spill volume in litres per year off Nova Scotia. .................................................. 59
Figure 14 Total spill volume in litres per year off Newfoundland. .............................................. 60
Figure 15 Oil budget for a spill of 4000 m
3
of Amauligak crude oil as calculated by ADIOS2. . 64
Figure 16 Predicted change in viscosity for Amauligak crude oil as calculated by ADIOS2. ..... 65
Figure 17 Predicted change in density for Amauligak crude oil as calculated by ADIOS........... 65
Figure 18 Oil budget for a spill of 1000 m
3
of IFO 180 as calculated by ADIOS2. .................... 66
Figure 19 Predicted change in viscosity for IFO 180 as calculated by ADIOS2. ........................ 67
Figure 20 Oil budget for a spill of 100 m
3
of MGO as calculated by ADIOS2............................ 68
Figure 21 Oil budget for a spill of 10 m
3
of jet fuel as calculated by ADIOS2............................ 69
Figure 22 The Arctic Food Web (ACIA, 2004)............................................................................ 98
Figure 23 Energy flow above trophic level 1 from phytoplankton and pelagic detritus in red, or
benthic detritus in blue, and proportional shades in between. Top panel: eastern Bering Sea
shelf; bottom panel: western Bering Sea shelf. Box and text size are proportional to log
10
of
biomass for the compartment; area of each link proportional to volume of flow (Aydin et
al., 2002). .............................................................................................................................. 99

iv
List of Tables

Table 1 Canadian Arctic oils (Drummond, 2006). ...................................................................... 15
Table 2 Vessel types operating in Arctic waters and oil types carried onboard. .......................... 50
Table 3 Reported vessel accidents in the Canadian Arctic. .......................................................... 57
Table 4 Historical large oil spills in barrels (bbl) from offshore well blowouts (source: Oil Spill
Intelligence Report database)................................................................................................ 61
Table 5 Valued ecosystem components (VEC) of various Arctic regions at risk from oil spills
(INAC, 2010; Word and Perkins, 2011). .............................................................................. 92
Table 6 Food web functional group and acute toxicity LC
50
(95% confidence interval) using 2-
methyl naphthalene for co-inhabiting Arctic species (Camus et al., 2010; Carroll et al.,
2010). .................................................................................................................................... 96
Table 7 Experimental crude oil spills of a few barrels to hundreds of barrels conducted in sea ice,
regardless of latitude (Dickins, 2011). ................................................................................ 166
Table 8 Spreading comparison for a 1600 m
3
(10,000 bbl) crude oil spill (SL Ross
Environmental Research Ltd. et al., 2010). ........................................................................ 167
Table 9 Concentration ranges (mg/L or parts per million) of several classes of naturally-
occurring metals and organic chemicals in produced water world-wide (Neff, 2002)....... 176
Table 10 National permissible concentrations of total oil and grease in produced water destined
for ocean disposal (Veil, 2006). .......................................................................................... 188
Table 11 Summary of produced water treatment systems used by three Arctic oil and gas
installations; adapted from Hawboldt et al. (2010)............................................................. 189
Table 12 Status in 2010 of resources and services injured by the Exxon Valdez oil spill in 1989.
Human services are those which were negatively impacted because of their connection with
impacted resources (EVOS Trustee Council, 2010). .......................................................... 207
Table 13 Sources of peer-reviewed biological effects data from Camus et al. (Camus et al., 2008)
with additional references added. ....................................................................................... 257

v
List of Acronyms
AhR aryl hydrocarbon receptor
AL Arabian light, crude oil
ANS Alaska North Slope, crude oil
API American Petroleum Institute
BaP benzo[a]pyrene
BIOS Baffin Island Oil Spill Project
BTEX benzenes, toluenes, ethylbenzenes, xylenes
CCG Canadian Coast Guard
CEWAF chemically enhanced water accommodated fraction
CNLOPB Canada-Newfoundland and Labrador Offshore Petroleum Board
CNSOPB Canada-Nova Scotia Offshore Petroleum Board
CYP1A
cytochrome P4501A, an enzyme system used as a biomarker for detecting biological
effects of xenobiotics
DE dispersant effectiveness
DGGE denaturing gradient gel electrophoresis
DOR dispersant-to-oil ratio
dpm disintegrations per minute (a measure of radioactivity)
DWH BP Macondo MC 252 Deepwater Horizon spill in the Gulf of Mexico, April 2010
EROD ethoxyresorufin-O-deethylase
EVCO Exxon Valdez crude oil; the oil specifically from this spill
EVOS Exxon Valdez oil spill
GC/MS gas chromatography-mass spectrometry
GST glutathione S-transferase
IFO intermediate fuel oil
IMO International Maritime Organization
ISB in situ burning
ITOSS In situ Treatment of Oiled Sediment Shorelines
JIP Joint Industry Program
LC
50

lethal concentration 50: the concentration of a toxicant that kills 50% of the test organisms
in an acute toxicity test
MESA medium South American, crude oil
MGO marine gasoil or marine diesel
OBM oil-based mud
OMA oil-mineral aggregates
PAH polycyclic aromatic hydrocarbon
PEC predicted environmental concentration
PNEC predicted no effect concentration
ppt parts per thousand; often used as units of salinity
PWS Prince William Sound
ROS reactive oxygen species
vi
SBM synthetic-based mud
SPM suspended particulate matter
TOSC total oxyradical scavenging capacity
TPAH total polycyclic aromatic hydrocarbons
TPH total petroleum hydrocarbons
UV ultraviolet light (10 nm to 400 nm)
VEC valued ecosystem component
WAF water accommodated fraction
WBM water-based drilling muds

vii
1
Executive Summary

The improvement of policies and regulations for spill response/remediation technologies, and
contingency plans for marine environmental protection to address the anticipated growth of
Canadas offshore petroleum industry in the North will require the best available information on
the factors and processes influencing the fate and effects of oil released into the marine
environment. As the result of past interests in the development of oil and gas resources in the
Canadian Arctic, as well as studies by other northern countries, a vast amount of basic
information exists to fulfill our need for science based advice to support environmental risk
assessments. However, this review of emerging environmental concerns has also identified a
number of knowledge gaps that should be addressed to ensure the protection of our marine
habitat and its living resources within the Arctic.

The risk of having an oil spill in Canadian Arctic waters is anticipated to increase because of
community growth that will increase marine traffic and industrial development including
offshore oil exploration and production. A range of petroleum hydrocarbon fluids from crude
oils to refined products will be transported within the Arctic. Of these, crude oil will likely be
the largest source of petroleum hydrocarbons transported. While the quantities of a spill from a
tanker may be large, the probability of a spill occurring is low based on the current information
available from historical records. Nevertheless, response measures must be considered within
contingency plans for a worst-case challenging situation such as a deep-well blowout in the
Arctic in the presence of ice.

A combination of laboratory, mesocosm and field studies have shown that the physico-chemical
properties of oil, temperature and the presence or absence of ice will influence the fate and
behaviour of oil spilled in the environment as well as the effectiveness of spill response
operations. For example, due to reduced rates of evaporative loss under cold temperature
conditions, oil will retain its viscosity and remain more persistent in Arctic waters. On the other
hand, there can also be some advantages to consider when oil is spilled in ice infected waters.
The decrease in oil evaporation may retain an oils flash point and viscosity providing an ideal
environment for in situ burning (ISB). The results of recent field tests have also demonstrated
that the inhibition of oil-weathering processes (natural, physical and chemical processes that
oil undergoes following its release into the environment) in ice and cold temperatures prolonged
the window of opportunity for the application of chemical oil dispersants as a spill response
strategy. In addition to active oil spill response strategies, there is also a renewed interest in the
potential rates of natural recovery in the Arctic following oil spill events. This is largely due to
advances in the application of biotechnology techniques in microbial ecology that have
highlighted the significance of natural oil biodegradation rates by indigenous bacteria and the
influence of suspended particulate material on the dispersion and biodegradation rates of residual
oil.

In terms of the development of predictive models on the fate, behaviour and effects of oil on
various components of the Arctic ecosystem, while there is a considerable amount of existing
data, the results from experimental studies are largely anecdotal or empirical in nature. As a
result, there is limited data of use for the development of integrated risk assessment models that
fully take into account the numerous physical, chemical and biological processes within the
Arctic ecosystem.

A multitude of biological effects have been observed in toxicological studies with oil with a
range of biota covering multiple trophic levels. In the Arctic, seasonal aggregations of animals,
such as marine mammals in open areas of sea ice, seabirds at breeding colonies or feeding sites,
or fish at spawning time may be particularly vulnerable to oil spills. For example, an oil spill in
the spawning areas of polar cod could severely reduce a year-class of the population. Appendix
1, Table 13, provides peer-reviewed biological effects data from an extensive review by Camus
et al. (2008), to which additional references have been added. There has been a shift in
biological effect studies from acute studies focused on mortality as the end-point to that of
chronic responses associated with much lower exposure levels and their effect on the long-term
health, growth and reproduction of the target organisms. With the implementation of ecosystem
based management by regulators, future studies must include consideration of biological effects
on population, and on community structure and function.

2
Interpretation of the data collected for environmental risk assessments is challenging as the
exposure conditions in past scientific studies (e.g. dosage and exposure time) are frequently
outside of the range observed in the field following actual spill events. Furthermore, as illustrated
by a case study following the Exxon Valdez spill in Prince William Sound, Alaska, a consensus
on the levels of environmental impacts have not been achieved due to a number of confounding
factors including different approaches to natural resource damage assessment, the lack of pre-
spill baseline information, and reported high levels of natural variation in population numbers
and community structure.

To address the knowledge gaps identified in this review, additional scientific research in
Canadian Arctic waters, including the conduct of large-scale field trials, should be conducted on:
1) the behaviour, transport and fate of oil spilled in the Canadian Arctic; methodologies to
monitor acute and chronic biological effects and recovery on multi-trophic level valued
ecosystem components; and 3) the development, application and validation of oil spill
countermeasures including natural recovery (natural attenuation). To optimize the use of
scientific expertise and resources, the reseach program would contribute towards an international
pan-Arctic global effort involving both government and non-governmental organizations
including academia and the private sector.





Suggested Citation:
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. 2011. 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.

3
1. Introduction

Despite rapid advances in the development of ocean renewable energy technologies, such as
offshore wind and tidal energy, Canada and the world will remain dependent on petroleum
hydrocarbons to meet future energy needs over the next few decades. With traditional
hydrocarbon reserves in the Western Canada Basin and other areas being depleted, exploration
and production operations within the oil and gas industry have shifted towards frontier regions in
the offshore and Arctic (Figure 1). Based on current analysis, it is evident that the bulk of our
newly discovered petroleum reserves and the best prospects for future discoveries will lie under
water rather than on land. The future of Canadas offshore oil and gas production may also rely
to a substantial extent on finds in deeper more distant locations on the outer continental shelf.

Figure 1 Since 1980 industry has increased the recovery of oil from the offshore sector to meet global demand
as land-based oil reserves declined. There is a greater risk of spills and damage as operations expand into the
frontier regions including the deep waters off the continental shelf and the Arctic (Sandrea and Sandrea,
2007).

Ocean sector activities related to Canadas oil and gas industry provide major socio-economic
benefits to Canada; $17.7B in direct GDP in 2006. This was linked to the direct generation of
over 171,340 jobs (Figure 2).
4

Figure 2 The significance of the oil and gas industry to Canadas GDP in 2006.

Among the greatest uncertainties in future energy supply, and a subject of considerable
environmental concern, is the amount of oil and gas yet to be found in the Arctic. The Arctic is
estimated to contain between 44 and 157 billion barrels of undiscovered recoverable oil. In
addition, the Arctic is gas-prone with an estimated 770 to 2990 trillion cubic feet of undiscovered
conventional natural gas (Bird et al., 2008). By using a probabilistic, geology-based
methodology, the United States Geological Survey examined the area north of the Arctic Circle
and concluded that about 30% of the worlds undiscovered gas and 13% of the worlds
undiscovered oil may be found there, mostly offshore under less than 500 metres of water.
Undiscovered natural gas is three times more abundant than oil in the Arctic and is largely
concentrated in Russia (Gautier et al., 2009). Thus, the Arctic continental shelves constitute one
of the worlds largest remaining prospective areas. To date, the remoteness and technical
difficulty presented by the Arctic, coupled with abundant low-cost petroleum in other regions of
the world, ensured that little exploration of the Arctic offshore reserves occurred. Even where
offshore wells have been drilled, in the Mackenzie Delta, the Barents Sea, the Sverdrup Basin,
and offshore Alaska, most of the resulting discoveries remain undeveloped.

Previous exploration activities by Imperial Oil Limited, Dome Petroleum Limited, Gulf Canada
Resources Limited, and Panarctic Oils Limited between 1960 and 1980 have verified the
presence of significant oil and gas finds in Canadas Arctic waters (SL Ross Environmental
5
Research Ltd. et al., 2010). The Drake F76 program (Panarctic Oils Limited) at a water depth of
55 m, approximately 1200 m offshore of Drake Point on the Sabine Peninsula of Melville Island
was the first offshore floating ice platform within Canada. The well at this facility with under-
the-ice subsea tree and subsea diverless connections of flowlines and controls (Bomba and
Brown, 2011) has proven the feasibility of offshore gas production in the Arctic region. The
Drake F76 program has been a showcase for advances in technologies and continues production
today.

With the rising prices and increased global demand for oil has come renewed interest in the
extraction of petroleum hydrocarbons in the Arctic, with the attendant concerns over sovereignty,
energy security, and advances in technology. Indeed, consideration has been recently given to
exploration and production in the deepwater Arctic environment of the U.S. Beaufort Sea (Pilisi
et al., 2011) by the use of winterized drill-ships constructed of material able to withstand the ice,
or an icebreaker converted into a drilling vessel. Increases in oil and gas exploration and
production activities in frontier regions would result in an increased risk of operational and
accidental releases of petroleum hydrocarbons due to the expansion of drilling operations,
marine support and shipping operations including that of pipelines. This risk is further
compounded in the Arctic environment due to environmental challenges including the
interference of ice, cold temperatures, isolated locations, high winds, and low visibility
especially during the winter when there are limited daylight hours. It is also important to note
that the availability of oil spill response personnel, and logistics for waste containment and
disposal in spill response operations, are issues in the Arctic. Furthermore, residents of
communities in Canadas Arctic are concerned over the effects of oil spills on indigenous species
and their habitat, as well as the effectiveness of existing oil spill response strategies that were not
originally designed for use in the North.

Due to the anticipated increase in both offshore oil and gas activities and onshore developments
(e.g. community growth, mining industry, etc.) in the Canadian Arctic, and the recent BP
Macondo MC 252 Deepwater Horizon spill (DWH) in the Gulf of Mexico in April 2010, more
stringent government policies and regulations for environmental protection are being considered.
A primary emphasis of environmental oversight is that of protecting the environment from
6
accidental spills. This includes all spills ranging from the small, repeated discharges linked to
routine exploration, production, processing and transport operations, to that of a catastrophic
spill of national or international significance. Within this context there is a demand to
understand the factors and processes influencing the fate and effects of oil released into the
marine environment. This information is essential for the development of improved spill
response and remediation technologies, and contingency plans.

To support a public review of Arctic safety and environmental offshore drilling requirements the
National Energy Board of Canada is looking for the best available information. This report has
been generated by the Centre for Offshore Oil, Gas and Energy Research, which is a Centre of
Expertise within the Science Branch of the Department of Fisheries and Oceans Canada, to
provide scientific facts and information on the fate and effects of oil in the Arctic marine
environment. The scope of the project includes:
review of the properties of oil that have been, or are likely to be discovered in the Arctic,
and the identification of chemical parameters that influence oil fate and effects in the
Canadian Arctic offshore or nearshore environment;
reporting on the fate and effect of oil that could be released as a result of accident, mishap
(including oil or fuel that would be in support of operational offshore oil and gas drilling
activities), or a well that becomes out of control;
review of relevant reports providing knowledge about the fate and effect of oil spills in
Arctic waters and information on the fate and effects of the Exxon Valdez spill in the Gulf of
Alaska;
incorporation of experience and knowledge from previous field studies relevant to offshore
oil and gas drilling in the Canadian Arctic;
identification of knowledge gaps that may influence the safety of offshore drilling
operations and the protection of the environment; and
recommendations for future research.

To meet the above study objectives, this review is structured in several sections. Chapter 1,
following the Introduction, is focused on the characterization and classification of crude oils
based on physical properties that influence the significance of oil fate and behavior in the
7
environment as well as the success of oil spill response strategies. A section of the chapter
describes the characteristics of the crude oils which we anticipate to be recovered within the
offshore Arctic regions of Canada during production operations in the future. Chapter 2 provides
an overview of current observations and scientific findings on the fate and behavior of oil spilled
in Arctic waters. Detailed information is provided on the influence of environmental conditions
such as ice-cover. Chapter 3 covers the subject of oil spills in the Canadian Arctic. The
potential sources of accidental and operational releases in Canadian Arctic waters are described
with risk analysis based on the probability and consequences of various spill scenarios that may
occur during anticipated future operations. Chapter 4 provides an overview of the current oil
spill response technologies available for use in the Arctic environment. Emphasis is given to
emerging technologies currently being developed for specific use in ice-infested waters. Chapter
5 provides a comprehensive review on the biological effects of oil spilled in Arctic waters. This
chapter covers topics such as routes of exposure, bioaccumulation, environmental effects
monitoring, potential changes on community structure and trophic level dynamics. Sub-sections
cover the effects of oil on various groups of organisms ranging from bacteria at the base of the
food web to the top predators such as polar bears and whales. Information is provided on the
environmental significance of valued ecosystem components (VECs) in Canadian Arctic waters
(e.g. polar cod) and their sensitivity to contaminant hydrocarbons. Chapter 6 covers the sources
of the various types of petroleum hydrocarbons that may be released into the Arctic marine
environment from operational activities. The potential environmental impacts associated with
the discharge of production waters and drilling muds are discussed. Chapter 7 describes the
results of experimental field trials conducted in the Arctic which have provided essential
information on the fate and behavior of oil in the environment as well as a platform for the
development and validation of oil spill countermeasures. A discussion on the need for additional
field trials to advance our scientific and technical knowledge is given. Chapter 8 provides a case
study on the impact of the oil spilled in Prince William Sound (PWS) from the grounding of the
tanker Exxon Valdez, oil spill response operations and the current controversy among various
parties over the extent of habitat recovery. Based on the information in this report, a list of
future research needs is presented in the final chapter to address the knowledge gaps that have
been identified.

8
2. Characterization and Classification of Crude Oil

2.1 Physical and Chemical Properties

Crude oils are complex mixtures comprised of hundreds to thousands of individual compounds.
However, the hydrocarbon content of crude oil can be separated into four main classes: saturates,
aromatics, resins (includes waxes), and asphaltenes (Fingas, 2010; Hannisdal et al., 2007). For
quantitative analysis, gas chromatography-mass spectrometry (GC/MS) can be used to
characterize the individual components of saturates, aromatics and biomarkers in crude oil, while
the resin and asphaltene content is usually measured using thin layer chromatography coupled to
flame ionization detection, or TLC-FID (Obermajer et al., 2010).

The chemical composition of crude oil influences its physical properties, which impact its fate
and transport properties on release into the marine environment. For example, a high content of
resins and asphaltenes will increase the viscosity making it less mobile. Crude oil chemical
composition can also affect the rate of biodegradation where microbial attack has generally been
ranked in the following order of decreasing susceptibility: saturates > aromatics > resins >
asphaltenes (Leahy and Colwell, 1990). Crude oils also contain a number of compounds referred
to as biomarkers, such as steranes and hopanes, which are persistent and less susceptible to
microbial attack. Since every crude oil has a unique biomarker profile, quantification of these
compounds provides a means for the identification of the origin of unknown oils in spill response
operations, provided that a database of oil biomarker profiles exists (Wang and Fingas, 1995). In
addition, other chemical components that are susceptible to biodegradation can be normalized to
these biomarkers to determine if changes in oil composition are linked to physical processes or
biodegradation (Prince et al., 2003b).

Physical properties, which are dependant on the chemical properties, can be used to characterize
crude oils. The main oil properties to take into consideration in an Arctic oil spill scenario, since
they are temperature dependant, include the American Petroleum Institute (API) gravity values,
viscosity, pour point, distillation characteristics, surface tension, flash point, and weathering
9
(ITOPFL, 2002; Payne et al., 1991; Shata, 2010). Definitions for these parameters are provided
here to better illustrate their significance in describing the physical properties of oil.
The specific gravity of oil is its density compared to seawater which is on average 1.025
g/mL. Most oils have a specific gravity <1.0 g/mL; therefore, they float on seawater.
The API gravity scale is used to describe the specific gravity of crude oil and it is
determined as (ASTM International, 2006; ITOPFL, 2002):
API = 141.5/(specific gravity) - 131.5
Crude oils are classified as light, medium or heavy according to their API gravity (ITOPFL,
2002). Light crudes have an API gravity >31, medium oils have an API gravity between
22 and 31, and heavy oils have an API gravity <22.
Viscosity of oil is its resistance to flow (Shata, 2010). A viscometer measures in units of
centipoises (cP) for dynamic viscosity, or centistokes (cSt) for kinematic viscosity.
Viscosity is governed by the chemical composition of oil. The higher the content of the
lighter components such as saturates and a lesser amount of asphaltenes, the lower the
viscosity (Fingas, 2010). When oil is spilled at sea it will spread over the sea surface. Oil
that flows readily at high temperatures can become a viscous mass at low temperatures
(Fingas, 2010). Spreading is greatly dependant on the viscosity of the oil, where viscosity
increases thus: light oil > medium oil > heavy oil. Temperature can affect viscosity (Figure
3). As the temperature falls, oil becomes more viscous (less mobile) and persistent in the
marine environment (Brandvik and Leirvik, 2008). In some cases, the rheological properties
(deformation and flow) of heavy oil can be altered using lighter fuels (Elasheva et al., 2001;
Schmidt et al., 2005). This process increases the commercial value of heavy oils and
decreases the viscosity, which improves mobility of the oil during transport (Elasheva et al.,
2001; Schmidt et al., 2005).
10
0
1
2
3
4
5
6
7
8
9
10
0 5 10 15 20 25 30
Temperature ( C)
C
e
n
t
i
P
o
i
s
e

Figure 3 Norman Wells crude oil dynamic viscosity as a function of temperature; data from Environment
Canadas Oil Properties database (Environment Canada, 2001a).

Pour point is the temperature below which oil will not flow (ITOPFL, 2002). The waxy
(resins) and asphaltenic components affect the pour point of oil. As the temperature of the
oil decreases, the wax components crystallize. This is often referred to as the cloud point.
This process hinders the flow of oil until it eventually changes from a liquid to a semi-solid
at the pour point (ITOPFL, 2002). The pour point is used to evaluate the flow of crude oil at
low temperatures (Zhang and Liu, 2008).
Distillation characteristics (evaporation) of crude oils describe their volatility (ITOPFL,
2002). As the temperature increases the low boiling point components, i.e. benzene, toluene,
ethylbenzene and xylenes, begin to evaporate or distil. Evaporation loss by weight or
volume is logarithmic with time for multi-component mixtures (Bobra, 1992; Fingas, 1994;
Fingas, 1999). The waxy and asphaltenic components of crude oil will not distil under
ambient conditions and remain persistent for extended periods in the environment.
Evaporation can change the physical and chemical composition of fresh crude oil.
Surface tension is the amount of pressure necessary to break the surface of a liquid. The
oil/water interfacial tension is the force of attraction between the surface molecules of the
oil and the water (Shata, 2010). The lower the interfacial tension at the seawater-oil
interface, the greater the extent of oil spreading (Fingas, 2010; Fingas and Hollebone, 2003;
11
The flash point is the lowest temperature at which an oil will ignite (Brandvik and Leirvik,
2008; Shata, 2010). This property is dependant on oil composition, in particular the volatile
components. The volatiles are generally lost due to evaporation; however, this process is
temperature dependant. In cold water, evaporation of oil is slow and the volatiles remain.
Fresh crude oils normally have a low flash point from -40 to 30C (Brandvik and Leirvik,
2008). If the flash point of crude oil is close to or lower than sea temperature then there is a
risk of fire or explosion hazard (Brandvik and Leirvik, 2008).
Weathering describes a series of natural physical and chemical changes that oil undergoes
following its release into the environment. Weathering of oil depends on the type of oil
(physical and chemical properties), environmental conditions (wind, waves, temperature and
sunlight), the properties of seawater (salinity, temperature) and the presence of biodegrading
microbes such as bacteria (Brandvik and Leirvik, 2008). The main environmental processes
that encourage weathering are oxidation, dispersion, dissolution and sedimentation, and
evaporation which lead to the disappearance of oil from the sea surface (see Section 3.2
Behaviour and Fate of Oil).

2.2 Significance of Oil Properties in Oil Spill Response

Oil properties will directly affect the fate and behaviour of oil spilled in the environment as well
as the effectiveness of spill response operations. Crude oil would be presumably more persistent
in Arctic waters because evaporation is slow, and spilled oil can become trapped under ice
making it less accessible to oil degrading bacteria and decreased weathering (Leahy and Colwell,
1990; Shata, 2010). At low temperatures oil becomes more viscous, the toxic short chain
volatiles remain intact and their solubility decreases, slowing the biodegradation process (Leahy
and Colwell, 1990). Environmental conditions in the Arctic may hamper the efficacy of current
oil spill response operations. The presence of ice may dampen the wave energy within ice floes
to a level below that required for effective chemical dispersion, precluding the use of chemical
dispersants as a spill response option (Deshpande et al., 2005; Shata, 2010). In addition, the
physical properties controlling dispersant effectiveness, namely the lowering of the surface
12
tension of seawater permitting oil to be broken into small droplets with wave energy, are less
effective in Arctic waters since surface tension and oil viscosity both increase with decreasing
temperature. The result is a thicker oil slick requiring more wave energy to disperse (Deshpande
et al., 2005; Glover and Dickins, 1999). On the other hand, there can also be some advantages to
consider when oil is spilled in ice infested waters (Shata, 2010). In areas where ice cover
provides boundaries, oil dispersants may be used as herding agents where they are sprayed
around the periphery of thin oil slicks, thus contracting the oil into a thicker slick (SL Ross
Environmental Research, 2010). The colder waters and increased thickening (viscosity) of the oil
decreases evaporation with the volatile components remaining in the oil, which maintains a low
flash point (Glover and Dickins, 1999; Shata, 2010). This provides an ideal environment for in
situ burning (Shata, 2010). In cold environmental conditions, oil persists longer, increasing the
time window for mechanical oil spill recovery operations and the application of other
remediation technologies (Shata, 2010). The formation of emulsions, where seawater becomes
suspended in oil, occurs slowly under ice cover due to damping of the waves (Shata, 2010).
Where emulsions occur, the selection of the appropriate spill response countermeasure will be
largely dependent on the process that stabilizes the emulsion (Friberg, 2007). In the event of a
deep sea well-head blowout such as the DWH, both pressure and temperature would influence
the subsequent fate and effect of the crude oil released (Leahy and Colwell, 1990).

2.3 Canadian Arctic Offshore Crude Oils

Northern Canada is potentially rich in oil and gas resources. Indeed, oil and gas has been
discovered in Baffin Bay, the Arctic Islands, the Mainland Territories, and the Beaufort-
Mackenzie Basin (Drummond, 2006). A total of 1544 wells were drilled in the Canadian Arctic
prior to 2004, and of these, 944 were exploratory (Drummond, 2006). There is an enormous
global demand for energy resources, and given our national demand for energy and potential
economic gains for Canada, the recovery of petroleum resources in the Arctic is now a priority
issue shared by government and industry. In order to assess the potential environmental impacts
from the production and transport of crude oil in Arctic waters, it is important to identify and
classify crude oils from the area to provide the information required for risk assessment and spill
response operations in the event that an accidental spill should occur.
13

In Northern Canada, a variety of oils have been identified for commercial production. These oils
are listed in Table 1. The origin of crude oils can greatly affect their physical and chemical
properties. Crude oils from the Arctic have been classified according to their API gravity values
with the aid of Environment Canadas database on oil properties (Environment Canada, 2001a;
Jokuty, 2001). Further information can be obtained from the database (Environment Canada,
2001a), which contains information on 450 oils. There is also a Spilltox database containing
toxicity data on over 30 crude oils (Environment Canada, 2001b). The hydrocarbon content of
crude oils can vary depending on the origin of the oil. Hydrocarbon content affects the physical
properties of Arctic oils, which for the most part determine the fate, effects, and transport of
spilled oil at sea.

There is considerable interest in defining the composition of a reference surrogate oil for use in
predictive numerical models on the fate, transport and effects of oil spills in the Arctic for risk
assessment purposes. Most of the crude oils harvested in the Arctic Region have an API gravity
greater than 22; therefore, it is recommended that a light or medium grade oil be selected from
the list in Table 1 to model the environmental impacts of an accidental oil spill in cold Arctic
waters. Based on the chemical composition of the crude oils found in the Arctic Region,
Atkinson crude has the highest aromatic content (Table 1). Unfortunately, as noted in the table,
the level of information on the chemical content for many of the crude oils recovered in the
Arctic is limited. Crude oil toxicity to marine life is associated with its aromatic content, namely
the polycyclic aromatic hydrocarbon (PAH) compounds containing 1-3 rings in the chemical
structure, and their alkylated homologues (Miller et al., 1982; Van den Heuvel Greve and
Koopmans, 2007). Based on the toxicity classification scheme by Van den Heuvel Greve and
Koopmans (2007), the three crude oil classes (light, medium and heavy) can be defined as
follows.
1. Light crude oil (low density, high toxicity): API gravity > 40
2. Medium crude oil (medium density, medium toxicity): 28 < API gravity < 40
3. Heavy crude oil (high density, low toxicity): API gravity < 28
14
Table 1 Canadian Arctic oils (Drummond, 2006).
Hydrocarbon Groups (%weight)
1
Origin Well API
1

gravity
Class
S A R A
Beaufort Sea Adgo 16.8 heavy 89 19 1 1
Beaufort Sea Amauligak 27.4 medium 90 9 1
Beaufort Sea Atkinson 23.7 medium 46 36 15 3
Beaufort Sea Issungnak 35.0 light 92 3 0 0
Beaufort Sea Koakoak 29.5 medium NA NA NA NA
Beaufort Sea Kopanoar 25.7
a
medium NA NA NA NA
Beaufort Sea Nektoralik 24.5 medium NA NA NA NA
Beaufort Sea Nerlerk 23.9 medium NA NA NA NA
Beaufort Sea Tarsiut 30.1 medium 92 7 0 0
Beaufort Sea Ukalerk 45.7 light NA NA NA NA
Beaufort Sea Uviluk 29.4 medium NA NA NA NA
Northwest
Territories
Bent Horn 41.3 light 94 5 7 0
Northwest
Territories
Bent Horn A-02 41.5 light NA NA NA NA
Northwest
Territories
Norman Wells 38.4 light 86 11 3 1
Beaufort Sea Adlaktok P-09
b
30 light NA NA NA NA
MacKenzie
Delta
Garry P-04
b
45 light NA NA NA NA
MacKenzie
Delta
Hansen G-07
b
22-57 light -
medium
NA NA NA NA
Beaufort Sea Isserk I-15
b
20-21 heavy NA NA NA NA
MacKenzie
Delta
Kugpik O-13
b
45-49 light NA NA NA NA
MacKenzie
Delta
Mayogiak J-17
b
33 light NA NA NA NA
Beaufort Sea Nipterk L-19
b
18-21 heavy NA NA NA NA
Beaufort Sea Nipterk P-32
b
29 medium
Beaufort Sea Pitsiulak A-0
b
30 light -
medium
NA NA NA NA
MacKenzie
Delta
Tuk J-29
b
32 light NA NA NA NA
MacKenzie
Delta
W. Atkinson L-17
b
27 medium NA NA NA NA
1
Environment Canada oil properties database, 2001
a
calculated from specific gravity
b
peer communication AEB
NA - not available



In this case, it is important to note that reported API gravity values used to classify the oils do
not conform to that defined by the American Petroleum Institute (ITOPFL, 2002). SINTEF
critically reviewed the classification provided and reported that it was not fit for operational
purposes as the heavy oil was more toxic than light oil (Van den Heuvel Greve and Koopmans,
15
2007). In contrast, data taken from Environment Canadas Spilltox database supported this
classification system (Van den Heuvel Greve and Koopmans, 2007). While API gravity values
aid in the classification of crude oils, the chemical composition of the oil is crucial for
environmental risk analysis since a high aromatic content may be an indicator of toxicity. Both
parameters provide essential information for the assessment of the potential impacts that a spill
will have in Arctic marine waters. Based on current information, the fate, effects and transport
of an accidental oil spill in cold Arctic waters can be adequately assessed using either light or
medium grade oils, such as Atkinson crude, as a surrogate to assess environmental impacts and
the efficacy of various oil spill remediation technologies.
16

17
3. Oil Spills in Arctic Waters

3.1 International Concerns and Governance

For the past 40 years the potential for pollution of the sea by accidental releases of oil has been a
concern of many national governments and other organizations. Despite advances in technology
and the implementation and enforcement of appropriate regulations and good working practices,
an expansion of shipping activity, and oil exploration and production will increase the risk of an
accidental oil spill. While we can minimize this risk, we cannot totally eliminate it. The recent
Arctic Marine Shipping Assessment Report completed under the leadership of the Arctic Council
Ministers Working Group on Protection of the Arctic Marine Environment has provided an
overview of ships and their infrastructure needs, and impacts in the Arctic Ocean (Fretheim et
al., 2011).

Contrary to that anticipated for offshore oil and gas exploration and production facilities, the
majority of marine traffic operations will occur in Arctic waters that are either permanently or
seasonally open. Nearly all current shipping activities take place on the periphery of the Arctic
Ocean, away from permanent or drifting ice. In other areas of the Arctic which have seasonal ice
cover, nearly all vessel activity occurs when and where the ice has melted or is melting so that
icebreakers are not required for assistance. The changing climate in the Arctic region is also
leading to a reduction in ice coverage that may increase the frequency of marine shipping traffic
and the risk of accidental oil spills.

The Arctic Marine Shipping Assessment reaffirmed the Arctic states view that the United
Nations Convention on the Law of the Sea remains the legal framework that influences and
guides current and future governance of the Arctic Ocean and also acknowledged that the
International Maritime Organization is the lead, and appropriate, UN body that can focus on
marine safety and environmental protection measures for the global maritime industry, including
operations in the Arctic. The Behaviour of oil and other Hazardous Substances in Arctic
waters project under the auspices of the Arctic Council has gathered and synthesized the current

knowledge and expertise on the behaviour of oil and other hazardous and noxious substances
released into Arctic waters resulting from accidental spills (Bjerkemo, 2011). The project
promoted the development and use of technologies and working methods to improve our
international capability to respond to accidents involving the spill of potential toxic
contaminants.

3.2 Behaviour and Fate of Oil

Various chemical and physical processes govern the fate and behaviour of oil following a spill in
the environment. It is important to note that the presence of seasonal ice has a significant effect
on oil weathering in the Arctic. It is thus important to distinguish the fate and behaviour of oil
during the winter season when ice is generally present, from when ice is absent or less
concentrated during the spring-summer season.

Weathering processes will change the oil and affect its properties and behaviour. Oil type and
environmental conditions such as temperature, sea state, winds, and other factors play an
important role in how the oil will weather over time. In some cases, these processes will
contribute to the natural removal of oil from the environment while in others they will contribute
towards the persistence of residual oil. Knowledge of oil behaviour is essential for the
identification of efficient response strategies at the time of a spill.

Work through the 1970s and 1980s largely focussed on questions related to oil behaviour and the
capture of oil under ice following a large blowout. Comprehensive reviews noted that during this
period, six full-scale field trials, twelve small-scale laboratory studies of oil and gas under static
ice, six studies of water current transport of oil under ice, and nine studies of oil on new ice-
growth were conducted (Dickins, 1994; Fingas and Hollebone, 2003). The majority of these
studies produced empirical relationships of oil spreading behaviour. The primary results were
that the oil spreading under ice was much slower than on the water surface alone and that the
extent was governed mostly by the shape of the under-ice surface. During the last two decades,
laboratory and mesocosm (including wave tank) studies, field tests and modelling efforts, have
been expanded to provide insights on the interactions that occur when oil, and oil and gas
18

mixtures are discharged in waters where ice is present. Much has also been learned from spill
response operations following accidental spills in ice-infested environments. A number of
literature reviews on the fate and behaviour of oil in icy waters have been compiled over the past
20 years (Fingas, 1992; Dickins, 1994; Reed et al., 1999; Fingas and Hollebone, 2003; Gjsteen
et al., 2003; Yapa and Dasanayaka, 2006; Brandvik, 2007; Khelifa, 2010; SL Ross
Environmental Research, 2010; Dickins, 2011). This review and that of Drozdowski et al.
(2011) summarize the current state-of-knowledge from studies to date, and how they might be
used to predict the behaviour of oil spills in the Arctic.

Oil in Ice-Free Waters

Ice coverage in the Arctic in most areas is seasonal. Generally, ice formation will take place
around September to October, and will start melting and breaking up around April to May.
During the summer months many areas are either free of ice or with ice concentrations of less
than 10% (drift ice). There are many possible scenarios in which an oil spill can occur, such as
sub-sea blowout, sub-sea pipeline leaks, oil tanker grounding or collision incidents, crude oil
leaking during loading and unloading, and bunker fuel releases. The fate and behaviour of oil
spilled in open water (ice-free) conditions has been extensively studied and described in many
forums and review papers (Huang, 1983; IMO (International Maritime Organisation), 1988;
ITOPFL, 1986; Reed et al., 1999; Spaulding, 1988; Yapa and ASCE Task Committee on
Modelling of Oil Spills, 1996).

Oil behaviour in ice-free Arctic waters is similar to oil behaviour at lower latitudes. The main
processes affecting oil behaviour in ice-free waters are summarized in Figure 4. As soon as oil is
spilled into the environment, it starts spreading on the sea surface, forming a thin layer. The
speed and extent at which the oil spreads will depend largely on its viscosity and on the quantity
spilled. Low viscosity oil will spread faster than high viscosity oil. Most crude oils discovered
so far in the Canadian Arctic are of medium viscosity. Low Arctic summer temperatures might
contribute to elevated oil viscosity and reduced spreading.

19

Once oil spreads and forms thin oil slicks, these start to fragment into smaller slicks or narrow
bands (windrows) under the influence of winds, waves and currents. At the same time, volatile
components evaporate into the atmosphere. Warm temperatures, high wind speeds and rough
seas will increase the rate of evaporation. Spreading and evaporation share a close relationship.
As oil spreads, the surface area of the slick will increase, increasing the evaporation rate.
Generally, refined products of low viscosity such as diesel or aviation fuel will evaporate more
rapidly than most crude oils. During the season in which Arctic waters are free of ice, low water
temperature still plays a significant role in determining the oil weathering rate. Colder
temperatures increase the viscosity of the oil, which affects spreading and dispersion but also
reduces the evaporation rate. In these conditions, an oil spill in ice-free water will behave as it
would in warmer conditions, but with a reduced weathering rate.


Figure 4 Physical, chemical and biological processes affecting the fate and behaviour of spilled oil (ITOPFL,
2002).

Natural dispersion will occur when turbulence generated by waves is sufficient to break up the
oil slick into small droplets of various sizes. These become entrained into the water column
where they mix in the upper layer and become diluted by the turbulent energy of the sea.
Eventually, these dispersed droplets are biodegraded by microbial organisms. However, in order
for this process to occur on a significant scale, oil droplets must average < 70 m diameter in
size, as larger droplets will recoalesce, come out of suspension and form a slick, which reduces
20

the surface area of the oil available to microbial attack. This process is largely dependent on oil
viscosity and sea state. Natural dispersion is more likely to happen with low oil viscosity (usually
Group 2 oil, API gravity > 40) and a sea state generating small breaking waves.

Biodegradation is the process by which microorganisms already present in the sea will use oil as
a carbon source and metabolize petroleum compounds. Biodegradation is a significant
weathering process when the oil is either dispersed into small droplets or broken up in a very thin
film. It is not a significant process when oil is highly viscous, remaining in thick slicks, or when
emulsification has taken place. It mainly occurs with light to medium oils in which hydrocarbon
chains are relatively short. Heavier oils with high wax or asphaltene content are difficult to
biodegrade. Biodegradation is a relatively slow process that necessitates the presence of
microorganisms, nutrients and oxygen. The rates are usually assumed to be slow in cold water
environments based on the majority of past microbiological studies. However, recent studies
under low temperature conditions including the Arctic, ice infested waters, and at deep ocean
depths (Hazen et al., 2010; Lee et al., 2011c; Lee et al., 2011d) have also shown significantly
rapid oil biodegradation rates.

Emulsification is another process that is likely to occur as oil slicks are drifting on the sea
surface. With the turbulent energy from waves, some oils will take up water droplets to form an
emulsion of water in oil. Emulsification usually forms in oil with a nickel or vanadium
concentration greater than 15 ppm, or an asphaltene content of more than 0.5%. However, the
most significant factor will be the presence of waves usually generated by a sea state of Beaufort
3 or above. Emulsions are usually very viscous and increase in volume up to four times the
volume of the original oil. They are also very persistent since other weathering processes are
greatly reduced once they are formed.

The weathering processes summarized in the preceding paragraphs occur as soon as oil is spilled.
Their relative importance varies mainly according to oil type and weather conditions. Oxidation,
dissolution and sedimentation can also take place during an oil spill, but in most cases, these
particular weathering processes play a limited role in the ultimate fate of oil. Sedimentation is
more likely to happen in the nearshore environment where oil becomes associated with
21

suspended particulate material including sediments which would facilitate its transport to the
seabed. Dissolution is likely to happen with light oils or refined products as these products have
a higher concentration of light compounds that are soluble.

Oil in Ice Covered Waters

All of the scenarios that occur in open water conditions could occur in locations where ice is
present; therefore, oil could be released under ice cover from a sub-sea leak or blowout, or onto
the ice surface (for example from a platform blowout), or in waters with ice floes. Numerous
field and laboratory research projects have been conducted since the 1980s to study oil
weathering in ice-covered waters. These have shown that oil is subjected to the same weathering
processes mentioned in the previous section. Because of the cold temperatures, oil viscosity will
generally increase for all oil types; however, interaction of oil with ice will affect the rate at
which these processes are taking place. Natural dispersion and emulsification are largely
dependent on wave energy in order to take place. The presence of ice will have a significant
impact, as short waves will be damped by ice floes, greatly reducing the rates of natural
dispersion and emulsification. In water with an ice coverage greater than 60%, the ice may
effectively contain and reduce oil movement, and thus protect sensitive shoreline resources.
Evaporation and biodegradation still take place in ice-covered water, but the low temperatures
usually associated with the presence of ice reduce the rate at which they occur.

Direct interaction of oil with ice affects its behaviour. Figure 5 summarizes the multiple ways oil
can interact with ice. The presence of ice greatly complicates the possible fate of oil spilled in
Arctic waters. The behaviour of oil in ice is complex, and the difficulties in modelling the
physics of ice formation and movement on scales of metres are magnified when the uncertainties
of oil behavior are added. If oil is spilled on ice, evaporative loss will be the main weathering
process as spreading will be limited by ice surface roughness and by absorption into snow.
Because of this, oil accumulations on the ice surface are expected to be limited in area and fairly
thick. For lighter oil types, absorption by snow could be a significant process depending on the
quantity of snow on the ice surface.

22


Figure 5 Oil and ice interaction processes (Bobra and Fingas, 1986).

Oil spilled under the ice is subject to various processes (Figure 5). In most cases, oil will drift
under the ice layer because of the currents and movement of the ice floes and will eventually
accumulate in naturally formed reservoirs due to ice roughness. These will vary in size but
significant quantities of oil could be trapped under the ice in this manner. Some of the oil could
be dislodged by currents and continue to drift under the ice. Several studies have set the
threshold for movement of oil under ice at 0.5 kt (26 cm/s). Oil 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 h) once oil is trapped under the ice. It is important to note
that oil weathering processes are inhibited by the encapsulation of oil. In springtime with
melting and warming of the ice sheet, encapsulated and trapped oil can migrate vertically
through brine channels and reach the surface of the ice to form pools of fresh oil. The rate of
vertical migration largely depends on the number of brine channels in the ice sheet and the oil
viscosity.

The lower oil weathering rate generally observed in ice-covered waters could represent an
advantage for response effectiveness in some spill scenarios. Reduced spreading in icy
23

conditions could increase the window of opportunity for certain response techniques (chemical
oil dispersant applications, for example) and be a significant advantage in the Arctic, since
residual oil would remain in an unweatherd state longer than it would under ice-free conditions.
However, the presence of ice in itself represents a great operational difficulty that can offset the
advantages provided by the reduced weathering rate.

3.3 Factors Influencing Oil Behaviour

Numerous field, laboratory and tank studies have been conducted over the past 30 years on the
behaviour of crude oils and refined products (including fuel oils) spreading in, on, and under ice.
A good understanding of the basic processes controlling the behaviour of fresh and weather oil,
and the development of numerical models to describe their time-series changes have been gained
from these efforts, and are reviewed by Lewis (2000), Fingas (2003), Yapa et al. (2006) and
Drozdowski et al. (2011).

Spreading in Broken Ice

An oil spill on a calm water surface spreads by gravity and is resisted by inertia, viscosity, and
surface tension until the slick reaches a thickness of approximately 0.1 mm (NRC, 2005). As
noted previously, the spreading speed is determined by many factors. Light crude oil spreads
much faster than heavy fuel oil, and wind, waves and currents can significantly increase the rate
of spreading. A rough sea with high mixing energy will significantly enhance this effect.

In cold regions, the spreading of oil can be significantly affected by the presence of snow and
ice. In broken ice, oil is assumed to move at the water surface, flowing around any ice present.
Oil can pile up and thicken around ice flows, and if sufficiently confined, may begin to flow
under the ice at the ice-water interface. It has been noted that for ice concentrations less than
30%, oil behaves as in open water (Deslauriers, 1979; ITG, 1983; SL Ross Environmental
Research Ltd. and DF Dickins Associates LLC., 1987; Venkatesh et al., 1990). For ice
concentrations greater than 30%, the oil is found to drift with ice. The equilibrium oil thickness
in slush or brash ice (accumulations of the wreckage of other forms of ice made up of fragments
24

not more than 2 m across) is nearly four times that on cold water, which itself is very different
from that on warm water. As a result, the oil-contaminated area at higher ice concentration is
several orders of magnitude smaller.

In tank tests, Martin et al. (1976) observed that both diesel oil and Prudhoe Bay crude oil
surfaced readily when poured into grease ice and slush, and they calculated an equilibrium
thickness of 1 mm for diesel oil in a mixture of pancake and grease ice. Sayed and Lset (1993),
studying the spreading of oil on water and among brash ice in moderate to high concentrations,
showed that the ice confined the oil to between 5/10 and 8/10 of surface coverage, increasing the
slick thickness and limiting the final equilibrium extent of the oil. A modified Fay equation for
spreading based on the results of small field spills represents the most suitable analysis to date.
In brash ice, spreading rates decrease with increasing ice concentrations and the presence of
slush ice strongly reduces the spreading, but the effect is small for ice concentrations below 20
30% (Gjsteen and Lset, 2004). Increased motion in the ice cover resulted in increased
spreading rates, and this effect was especially pronounced in the presence of slush. Decreased
spreading rates due to increased oil viscosity were also observed.

In terms of developing predictive models, Gjsteen and Lset (2004) measured spreading rates
of marine fuel oils in slush and frazil ice with different wave energies. It was clear that the
presence of broken ice significantly slowed oil spreading. Gjsteen (2004) developed a
mathematical model using Newtonian viscosity to predict the spreading of oils over water and
slush ice, and found good agreement with the lab data of Sayed and Lset (1993). The model
was reportedly coupled to a discrete-element ice model. Hara et al. (2008) developed a model
for flow of oil through a broken ice field, and included a specific term for the critical thickness of
oil which will begin to flow under the ice, rather than through the broken ice field, dependent on
the oil and water densities and the oil-ice-water interfacial tension. This model has not yet been
tested.

Lewis (2000) concluded that at lower ice coverage (< 30%), the oil and ice will move at different
rates under the influence of wind. The influence of wind on an oil slick is to increase the
movement of the slick by approximately 3%, while ice floes tend to move the slick faster (4.5 to
25

6%). At ice concentrations greater than 50%, oil drifts with ice at speeds from 4 to 7% and the
rate decreases as the ice concentration increases.

Field experimental data on oil spreading under various broken ice conditions (4/10, 5/10 and >
9/10 coverage) has been reported (Buist and Bjerkelund, 1986). With increased coverage of
brash ice, oil spreading was dramatically reduced. A phenomenon described as lead pumping
was reported to redistribute oil from the water to the ice surface under dynamic conditions. The
effects of lead closure rates on the vertical movement of oil were studied by MacNeil and
Goodman (1987) in an outdoor basin. It was observed that oil was forced under the ice when a
lead closed slowly. At high closure rates (above 12 cm/s) most of the oil was forced to the top
surface of the ice.

Weerasuriya and Yapa (1993) and Yapa and Belaskas (1993) experimented with spreading of oil
under and over simulated broken ice fields in small tanks. Based on the laboratory observations it
was concluded that the behaviour of oil spilled under a fragmented ice cover depends on the type
of ice cover. While oil may penetrate completely through one type of cover, it may not penetrate
another.

In 1993, numerous authors reported on test tank experiments and a subsequent experimental spill
of North Sea crude in the Barents Sea marginal ice zone off the coast of Norway (Jensen, 1994;
Reed and Aamo, 1994; Singsaas et al., 1994). High concentrations of pack ice (90% initially,
declining to 75% at the end of the experiment) kept the oil thick and immobile for days which, in
combination with cold temperatures and the damping of wave action by the ice, significantly
slowed oil weathering processes (evaporation, natural dispersion and emulsification). Brandvik
(2006a) presents a comparison between the results obtained from the experimental spill in pack
ice with a similar experimental spill in open water.




26

Movement on Ice

The spreading of oil on ice is more similar to spreading of oil on land than on the sea, and the
slick is much thicker (Lewis, 2000). The spreading rate for oil on ice is determined by the oil
density and viscosity as well as the roughness of the surface over which the oil is spreading.
Discrete ice deformation (rafting, rubble, pressure ridge) may lead to a contained, thick, oil pool.
Glaeser and Vance (1971) is probably the earliest reference to oil spills on ice. It was found in
their study that the ice surface adsorbed the oil to a saturation level of about 25% and the degree
of spreading is a function of spill volume and spreading time. Spreading is also affected by
temperature (Chen et al., 1976). There was no spreading below -19
o
C and warm oil spread more
rapidly. Another early study on oil spreading on ice concluded that gravity is the only important
spreading force and the final radius is only a function of time (McMinn and Golden., 1973).
Additional new papers have been written on oil spill spreading on ice (Brandvik and Faksness,
2009; Dickins et al., 2011; Jaraula et al., 2008; Peishi et al., 2011; Wang et al., 2008).

Movement Under Ice

Oil pooling immediately underneath the ice layer has been observed in laboratory experiments
(Payne et al., 1991) and field trials (Dickins et al., 2008). A site specific study (Ramseier and
Rene, 1973) showed that with water currents near the North Pole, the oil spreading rate under ice
was 800 cm/day. Chen et al. (1976) observed the spreading of oil under a freshwater ice sheet in
a small test tank and found that the radius of oil is affected by the water density, water or oil
viscosity, time, and volume rate of oil flow to the slick. If oil is in direct contact with the
underside of the ice, oil viscosity is important. The spreading rate was proportional to the
power of the elapsed time in the absence of current. In the presence of strong current, oil droplets
travelled some distance before rising, and many droplets did not adhere to the ice (Fingas and
Hollebone, 2003).

It was reported that the effect of an ambient current on oil transport under ice is much greater
than effects due to forces such as buoyancy, viscosity, and interfacial tension (Uzuner et al.,
27

1979). However, some have argued that buoyancy and viscosity are the dominant forces for oil
spreading under ice (Yapa and Belaskas, 1993; Yapa and Chowdhury, 1990).

Buist et al. (1981) studied the behaviour of oil and gas release under ice in the Beaufort Sea. It
was observed that oil broke into droplets as the oil and gas mixture left the discharge pipe due to
turbulence. As the gas rose out of the jet stream it quickly collected in a pocket under the ice. For
the under-ice oil and gas injection study at a gas to oil ratio of 150:1, it was observed that oil rose
in the form of pendent drops of 1 or 2 cm in diameter. With currents of 14 cm/s in the flume
tank, oil began to migrate downstream. Another study involving gas as well as oil under ice is
the work done by Purves (1978). For a 60:1 ratio of gas to oil released under ice in a test tank, it
was found that the presence of gas helped the oil to spread more rapidly and thinner.

It was summarized by Fingas and Hollebone (2003) that in low currents, the oil released under
ice will spread upon reaching the under ice surface due to combined actions of buoyancy,
viscosity and surface tension. A number of force-balance models have been developed to predict
spreading under a smooth ice bottom (reviewed in Fingas and Hollebone 2003). However, in
practice, sea-ice is characterized by significant under-ice roughness, and the final under-ice
configuration is determined by the under-ice roughness and topography. Oil has been observed to
spread systematically, filling the nearest under-ice depressions first before overflowing into the
next depression. Malcolm and Cammaert (1981) calculated the pool thickness of crude oil under
ice in sea water, and the thickness ranged from 4.27 to 10.63 mm for interfacial tension ranging
from 10 to 30 dynes/cm. Local ice conditions are much more important to the final oil
disposition than microscale spreading behaviour. A volumetric analysis is considered to be the
most effective approach for predicting the spread of large oil and gas discharges under an ice
sheet, and several general spreading models have utilized this method. The key parameters are
oil and gas volumes, under-ice storage capacity, and potential for gas expansion.

Oil under stationary or land-fast ice spreads according to not just the fluid property factors, but is
also governed by currents and the under-ice topography. Several field observations have noted
that, in the absence of small currents, the deposition and spreading of oil under continuous ice is
largely governed by the shape of the under-ice surface. An early approach to estimating the oil
28

volume holding capacity of ice for oil was to estimate the volume above the mean draft level of
the under-ice surface. Based on measurements of under-ice topography (Goodman et al., 1987;
Kovacs et al., 1981), this geometric filling approach implied that oil spills under ice would be
confined to a much smaller area than spills in open water (Dickens et al., 2008). Note that these
models considered the final geometric holding capacity of the under-ice topography only, and do
not consider the oil spreading dynamics.

Early field observations of oil spilled under ice in Beaufort Sea had indicated that oil under a
smooth continuous ice surface would spread to an equilibrium thickness of approximately 1 cm
(Dome Petroleum Ltd., 1981; NORCOR Engineering & Research Ltd, 1975). This quality was
confirmed in a series of observations of oils spilled under smooth ice (Rytkonen et al., 1998): the
oil spreading speed was dependent on oil viscosity and the oils spread to an equilibrium
thickness of 7 mm to 14 mm, independent of viscosity. This minimum equilibrium thickness
concept was key to a new model for the spreading of oil under ice proposed by Wilkinson et al.
(2007a; 2007b), which included the oil-filling dynamics driven by gravity-controlled motion
under the ice cover. They found that the oil pooled in much lower areal concentrations than
geometric volume factors predicted, spreading to cover an area 14 times greater than predicted
by a mean-volume filling approach. This new theoretical approach did not incorporate
hydrodynamic currents or oil plume spreading during rise to the ice sheet, which would be
expected to further increase spill area. While the approach provides a physics-based process
model for spreading under continuous ice, it remains to be validated by experimental
observations.

In terms of knowledge gaps, a particular area that requires much more research is the movement
of spilled oil under old, multi-year ice. Old ice would be encountered in deeper water and off
Greenland which are locations for future oil and gas exploration (Cairn, 2011; Dickins, 2011).

Movement Through Ice

Oil released under a first year ice sheet rapidly became encapsulated and remained in place until
February and March when it began to migrate through former brine channels to the surface
29

(NORCOR Engineering & Research Ltd, 1975). The rate of encapsulation depends on the air-
ice-water temperature gradient, the under-ice topography, and the volume and properties of the
trapped oil (Fingas and Hollebone, 2003). The observed vertical movement length in March was
only about 20 cm but increased significantly in April. Similarly, it was observed that oil started
appearing on the surface in early June from a test release done in December (Buist et al., 1981).
In anecdotal data reported by early field trials, NORCOR Engineering and Research Ltd. (1975),
Dome Petroleum Ltd. (1981) and Buist et al. (1983) all reported that oil or emulsions were
trapped within 24 hours, and sometimes within four hours (Buist et al., 1983). Numerical models
based on heat flow across the oil and gas lens generally predict that the ice growth beneath the
lens will be reduced (as the thermal conductivity of the oil is less than that of the ice for most
field situations). However, no measurable difference in thickness between oiled and unoiled ice
was observed in the NORCOR or Dome Petroleum field spills. This may be attributable to
natural variation in snow cover which produces variations in ice thickness greater than those
which could be induced by the oil. Unless very thick pools are involved, the effect of
encapsulated oil on subsequent ice growth beneath the lens will probably be minimal.

The available field and laboratory test data show that the encapsulated oil will be released in the
spring as the ice sheet deteriorates. Oil escapes from the ice sheet by a combination of two
general processes: vertical rise of the oil through the brine channels in the ice, and ablation of the
ice surface down to the oil lens in the ice. For a combined oil and gas spill, the gas will be
released in advance of the oil. Both release processes have been observed to occur in the field.
The relative quantities of oil released depend on several factors including the depth of the oil lens
in the ice, the rate of brine channel opening and the configuration of the oil in the ice, whether as
discrete droplets or pools of oil.

Of the field spills conducted to date under first-year sea ice, encapsulated oil was released in the
next melt season. Data on release from multi-year ice is limited to one series of small-scale field
spills. This study indicated oil rises quickly to the surface through cracks, but persists for at least
two melt seasons and possibly five on the surface of the ice because multi-year ice is thicker and
the brine channels may be discontinuous.

30

Russian researchers conducted experiments on oil spilled under ice floes similar to those found
in the Sea of Okhotsk (Ohtsuka et al., 2001; Ohtsuka et al., 1999). These authors found that oil
progressively filled the cavities on the bottom of the ice floes. Less than 1% of the oil permeated
through a 7-10 cm thick floe of pancake ice. Faskness and Brandvik (2005) report a series of
experiments involving freezing lenses of oil at differing depths to examine the migration of
PAHs, the water-soluble components from six different crude oils, over a winter field season.
They found that oil compounds (but not the whole oil) diffused through the ice sheet forming a
concentration gradient to the bottom of the sheet, but at concentrations that were very low (less
than 6 ppb) at the water interface.

The movement of oil through multi-year ice has been reported by Comfort and Purves (1982). It
was observed that most of the oil placed under multi-year ice migrated to the surface after one
year and no oil remained at the study site after five years.

Adsorption to Snow

The earliest study of oil spill processes in snow can be traced back to McMinn (1972) who
conducted experimental spills of Prudhoe Bay crude oil on Arctic snow and ice. Oil adsorbed to
snow to form a fairly stable mulch (Buist et al., 1987; McMinn, 1972). The adsorbed amount was
20% and 25% in the case of McMinn (1972) and Buist et al. (1987) respectively.

Oil spilled onto snow may also flow downward to the ice layer, and then creep onward beneath
the snow layer. A comparison of theoretical equations (Belore and Buist, 1988) with
experimental data (Glaeser and Vance, 1971; McMinn, 1972) indicates that the initial spreading
of instantaneous spills on snow or ice is governed by the gravity-inertia regime, but spreading
reverts to the gravity-viscosity regime. The horizontal flow of oil in snow at an impermeable
layer was a function of snow properties (void ratio, specific permeability), dynamic viscosity of
the oil, and time. Nelson and Allen (1982) observed that oil soaks into snow more thoroughly if
the air temperature is higher (4
o
C) and the snow softer. Bech and Sveum (1991) observed that oil
soaks through the snow and spreads out. The containment of diesel fuel by snow is reported by
Allen (1978).
31


3.4 Factors Influencing Oil Fate (Weathering)

Changes to oil properties following its release into the environment have received considerable
attention as it influences the behaviour of the oil. The properties that are altered during oil
exposure, thus defining the fate of oil include density, viscosity, surface and interfacial tensions,
and water content. Typical weathering processes include evaporation to the air, dissolution in
water, emulsification with water to form water-in-oil gels, photodegradation and dispersion of oil
into water.

Singsaas et al. (1994) found that the rates of evaporation, dispersion and emulsification were all
significantly retarded in the broken ice found in leads. They reported that the primary factors
reducing the weathering rate when compared to open water and temperate conditions were wave-
damping, limited spreading of oil bounded by sea ice, and temperature.

Evaporation

Evaporation is the most important and rapid of all weathering processes, and it accounts for the
loss of 20-50% of many crude oils, 75% or more of refined petroleum products, and 10% or less
of residual fuel oils (NRC, 2003; NRC, 2005). Immediate surface evaporation results in losses of
small alkanes (C
5
-C
10
), and monoaromatic compounds like BTEX (benzenes, toluenes,
ethylbenzenes, xylenes). Several experiments with refined gasoline and diesel (Ivanov et al.,
2005; Serova, 1992) have found nearly complete evaporation on the surface of ice during the
summer season.

Payne et al. (1991) observed that under spring and summer conditions simulated in wave tank
systems, all compounds with vapour pressures greater than C
11
-alkanes were lost within nine
days. The evaporation of oil on ice in winter is slower, but eventually the oil will evaporate to
approximately the same degree as it would if spilled on the water in summer. Since evaporation
is a surface phenomenon, the evaporation rate can be significantly influenced by the oil film
32

thickness as a result of spreading. The rate of evaporation is also increased by strong winds,
rough seas and higher air temperature.

The effects of snow on evaporation have been discussed by Belore and Buist (1988). It was
observed in their experiments that oil in a control tray (no snow) experienced the greatest
evaporative loss. Of the ones with snow, the uncompacted snow had the highest evaporative loss.
Increasing snow density and thickness reduced evaporation.

Brandvik and Faksness (2009) studied evaporative loss of a light oil (fresh Statfjord crude) under
three different ice coverage levels at various surface current and wave height conditions with
different air temperatures (-15 to about -5
o
C). They reported that evaporative loss for open water,
30% ice coverage, and 90% ice coverage was 30%, 25%, and 19%, respectively. The difference
in evaporative loss was mainly due to the different film thicknesses. Ice concentration also
affects pour point and flash point. They also found that the viscosity change due to evaporation
was minor, but changes in viscosity due to water-in-oil emulsion were significant.

In freezing waters during ice formation, layers of ice platelets can form to cover the oil surface
and cause encapsulation of oil among ice crystals. In this case, the spill will remain entrapped
within the columnar ice until a thaw cycle is initiated. During and after oil encapsulation within
ice in a flow-through system, the concentration of dissolved components and dispersed oil
droplets in the water column was effectively diluted to background levels within several days
(Payne et al., 1991).

Dissolution

Compounds of low molecular weight in an oil slick are subject to two competing processes:
dissolution into the surrounding water and evaporation into the atmosphere. For surface spills,
true dissolution of individual components from the slick is not considered significant in terms of
the overall mass balance of an oil spill (NRC, 2003; NRC, 2005). Dissolution of individual
components is important, however, when considering the potential for biological impacts.
33

Dissolution of volatile and semi-volatile components of oil from a deepwater blowout oil spill
can be significant.

The dissolution of water soluble components through sea ice has been studied by performing
laboratory experiments and field trials dealing with oil encapsulated in ice (Faksness and
Brandvik, 2008a; Faksness and Brandvik, 2008b). By freezing six oils (five crude oils and a
heavy fuel oil) into sea ice in winter and quantifying the migration of water soluble components
from the oil into the ice, it was found that the content of water soluble compounds in the ice
decreased with ice depth, and the concentrations changed as a function of time, confirming that
the water soluble components had been transported from the overlying oil through the brine
channels in the ice to the underlying water. The field experiments showed that not only ice
thickness, but also air temperature prior to an oil spill, are important for the distribution of water
soluble components in ice. At relatively high air temperatures, ice has higher porosity, and the
majority of the water soluble components will leak out more quickly. Conversely, if a spill
occurs during extremely cold conditions, the ice is less porous and water soluble components are
released at a slower rate, resulting in a longer exposure period of toxic components to ice-
dwelling (sympagic) organisms.

Dispersion

Under turbulent hydrodynamic regimes, a surface oil slick can be entrained in the water as small
oil droplets that are dissipated and diluted in the water column (Delvigne and Sweeney, 1988;
Fraser and Wicks III, 1995; Lee et al., 2001). This is a natural process that is beneficial to oil
spill cleanup, because oil in the form of small droplets with a higher surface to volume ratio is
biodegraded more rapidly. Furthermore, the formation of oil droplets enhances interaction with
suspended particulate material. Waves create eddies of different sizes cascading to the smallest
size class where the energy is dissipated ultimately into heat under the influence of viscosity. The
steep velocity gradients of the small eddies are of major importance for shearing off of droplets
from the slick and for the collision probability of droplets, whereas the large eddies are primarily
responsible for the diffusive transport of the oil droplets.

34

The entrainment rate of oil depends upon the properties of oil such as viscosity, the oil-water
interfacial tension, and the surface energy breaking rate. As the slick weathers, the oils physico-
chemical properties change. Viscosity, in particular, can increase by orders of magnitude during
the uptake of water into the slick to form a water-in-oil emulsion. The entrainment rate also
depends upon the fraction of the sea surface hit by breaking waves per unit time, which in turn is
a function of wind speed. Stochmal and Gurgul (1992) found that oil in water dispersions
separated 3.5 times more slowly at near-freezing temperatures (1C) than at 15C.

Emulsification

Water-in-oil emulsion, commonly called chocolate mousse, is formed by dispersion inversion or
by water drop entrainment. Emulsification competes with oil-in-water dispersion (dispersed oil).
Water-in-oil emulsion is an important weathering process, especially in the early stages of an oil
spill, which alters the oil composition and behaviour. Emulsification increases the volume of the
slick and reduces the biodegradation rate. The formation of stable water-in-oil emulsions has a
negative impact on the success of various oil spill cleanup efforts, including increased volume
for containment, lowered combustibility for ISB, and increased viscosity so that chemical
dispersion is more difficult. Rapid emulsion formation also diminishes the rates of evaporation
and dissolution, retaining more toxic low molecular aromatic compounds in the residual oil.

The rate and extent of the formation of water-in-oil emulsion are dependent upon oil properties,
the turbulence energy level, and environmental conditions. Weathering caused by evaporation in
water-in-oil emulsions of crude oil occurs to a significant extent in a matter of 12 to 24 h at sea
surface temperatures above 10C (Daling et al., 1997; Lewis et al., 1995). Weathering of Brage
crude oil, as water-in-oil emulsions in a meso-scale flume at 13C, reached a maximum water
content (80%) within 4 h and emulsion viscosity steadily increased during the initial 70 h (Lewis
et al., 1995). The corresponding chemical dispersability of the weathered oil dropped drastically
within the first 12 to 24 h.

When oils are released into ice-covered waters, ice formation may dramatically alter the rates
and relative importance of the different weathering processes. Oil released into a growing slush
35

ice field during late fall or early winter may be subject to stranding on the upper ice pan surfaces
and rapid water-in-oil emulsification, followed by partial density-mediated submersion and
incorporation into the ice canopy (Payne et al., 1991). For the oil on the upper ice surface of
pans and smaller floes, diffusion-controlled evaporation predominates. For the portion of oil that
remains in the ice-water matrix, emulsification may occur rapidly before significant evaporation
and dissolution can occur, affecting the overall mass balance of the oil slick. Oil from a
subsurface release under an existing ice canopy is subject to encapsulation before evaporative
weathering, although dissolution has been demonstrated.

Brandvik and Faksness (2009) describe oil weathering experiments with a light naphthenic crude
in broken ice (0%, 30% and 90% coverage) in an outdoor race-track flume cut in the ice. They
reported that the weathering of spilled oil was reduced with increasing ice coverage. The
formation of water-in-oil emulsions was reduced by wave damping by the ice. Ice confined the
oil, increasing thicknesses and reducing available surface area. This geometric effect slowed
evaporation at the highest ice coverage and reduced emulsification. At low ice cover, little
difference in emulsification, monitored by viscosity increase and water-uptake, was observed
from oil in open water. The effect of ice on water-in-oil emulsification (38-41% water content)
between open water and 30% ice cover was found to be small and probably insignificant, but the
difference with 90% ice cover (20% water content) was large (Brandvik and Faksness, 2009). In
comparison, weathering of Statfjord oil in the open water without any slush ice can reach a water
uptake of 75-85%.

Buist et al. (1983) placed a water-in-oil emulsion under first year ice in the Southern Beaufort
Sea and the emulsion became encapsulated within 48 h. Unlike oil which can migrate to the
surface through brine channels (Buist et al., 1981), the emulsion only appeared on the surface
during ice melt.

The weathering and fate of water-in-oil emulsifications of four Alaskan crude oils, namely ANS,
Endicott, Kuparuk and Northstar, in pack ice meso-scale and large-scale wave tanks, were
evaluated in a three-year program of laboratory and test-tank trials (Buist et al., 2009). Empirical
relationships based on oil properties for evaporation of oil on water, in broken ice and in snow;
36

equilibrium spreading thickness on water; equilibrium spreading thickness on ice; the extent of
oil spreading on ice; the extent of oil spreading in snow; and the critical current velocity which
results in stripping of oil from under-ice surfaces were reported. The researchers were unable to
develop quantitative relationships for the rate or probability of emulsion formation in broken ice
and the rate of appearance of oil encapsulated in ice during spring melt. Very little emulsification
occurred with oils under the low wave energy that was tested. Emulsion formation was impeded
by wax precipitation in the test oils, interfering with the uptake of water to form a stable
emulsion. This caused the oils to become non-Newtonian fluids, which, as well as reducing
emulsion formation, also greatly reduced spreading of the oil on water, on ice and through snow.
The wax matrix of the oils, gelled by the low temperature, reduced the evaporation rate of the
oil, which Buist et al. (Buist et al., 2009) suggested could be modeled as a lowered mass transfer
term in the evaporation equation.

Most oils formed high water content (80%) emulsions under the high wave conditions that were
tested (Buist et al., 2009). The emulsions formed more rapidly in higher ice concentrations. This
is consistent with the conclusions reached by Payne et al. (1991), but in conflict with the results
of Brandvik & Faksness (2009), probably due to the differences of the surface energy levels and
the ice morphologies of these test systems.

Brandvik and Faksness (2009) hypothesized that weathering processes strongly influenced by
cold temperatures may extend the operational time window for several contingency methods in
comparison to similar treatment of oil spills in open water. Such processes include reduced oil
spreading (caused by ice coverage and weakened wave action), decreased water uptake (owing to
less oil-water interaction), less evaporation, and slower changes in viscosity and pour point in
dense ice conditions. All these factors can potentially open up a wider window of opportunity for
dispersant and in situ burn (ISB) treatments.

Brandvik et al. (2010a) subsequently reported improvements to the SINTEF oil weathering
model as a result of studies carried out under the joint industry program (JIP). In a series of
experiments conducted in a permanent race-track flume and in field trials in a temporary in-ice
flume similar to the design used in their previous work (Brandvik and Faksness, 2009), and in
37

trial spills into natural pack ice, they evaluated oil evaporation, emulsification, viscosity increase,
water uptake and natural dispersion for five oils from the North Sea. They found that the various
weathering processes were highly dependent on the initial oil composition and properties.
Similar to their earlier results, they reported that oil in open water and low ice conditions
weathered most quickly, while oil in high ice cover had only slight increases in viscosity, low
water uptake and small evaporative losses. They suggested that a combination of wave
dampening reducing the shear experienced by the oil, and herding by the ice, greatly reduces the
rate of change of an oil in Arctic ice-infested waters.

Photo-Oxidation

Photo-oxidation and biodegradation are two very important transformation processes that occur
with petroleum products that are released into the marine environment (Dutta and Harayama,
2000; Garrett et al., 1998; Prince et al., 2003b). Experiments with refined gasoline and diesel
(Ivanov et al., 2005; Serova, 1992) on the surface of ice during the summer season showed that,
due to longer exposure times, photo-oxidation is a much more important process for oil
degradation in the Arctic than in more temperate climates. Despite some reservations over the
toxicity of products formed (Lee, 2003; Maki et al., 2001), photo-oxidation of crude oil by
natural sunlight may provide an opportunity for the introduction of novel procedures for the
remediation of marine oil spills. Garret et al. (1998) demonstrated that photo-oxidation and
biodegradation have opposite effects on aromatic hydrocarbons in crude oils. Whereas microbial
degradation and evaporation/dissolution result in the depletion of unsubstituted aromatic
compounds relative to their alkylated homologues, photo-oxidation selectively degrades the
alkylated aromatic compounds (Prince, 1993; Prince and Clark, 2004). Nevertheless, since
neither photochemistry nor biodegradation affect the levels of hopanes and other biomarkers to a
significant extent, referencing the PAHs to hopane or a similar conserved marker can reveal the
extent of degradation.



38

Biodegradation

Many microorganisms that are ubiquitous in nature have the capacity to utilize petroleum
hydrocarbons as the sole source of carbon and energy (Atlas, 1981; Head et al., 2006). A wide
variety of bacterial and fungal genera exhibit the capacity to degrade hydrocarbons. Many
hydrocarbon degrading microorganisms have been isolated from a variety of marine
environments (Atlas, 1984; Leahy and Colwell, 1990; ZoBell, 1973). Recently, Prince et al.
(2010b; 2010a) published a list of 181 genera of bacteria, 163 genera of filamentous fungi and
yeast, and 22 genera of algae that are able to degrade hydrocarbons by metabolizing them in
order to grow. These findings are not surprising considering the fact that marine microorganisms
have long been exposed to significant quantities of petroleum hydrocarbons from natural
seepages.

From 1990 to 1999, approximately 600,000 tons of petroleum were released into the worlds
oceans per year from natural seepages alone (NRC, 2003; Stout and Wang, 2008).
Biodegradation by microbial communities is the major process controlling the weathering and
eventual removal of oil from natural seeps that enters the marine environment (Atlas, 1995; Atlas
and Bartha, 1992; Leahy and Colwell, 1990). Bacteria are the predominant hydrocarbon
degraders in the marine environment, with a global distribution from tropical environments to
extremely cold Antarctic and Arctic environments (Head et al., 2006; Venosa and Zhu, 2003).
They are normally present in small numbers in pristine environments, but these oil degrading
microbes can be selected and multiply rapidly upon the introduction of oil (Atlas, 1995).

Biodegradation rates have been shown to be the highest for saturates, followed by light
aromatics, with high-molecular-weight aromatics and polar compounds exhibiting extremely low
biodegradation rates (Prince, 2010c). Co-metabolism plays an important role in oil
biodegradation and may require microbial consortia or syntrophic interspecies cooperation
(McInerney et al., 2008). Many complex branched, cyclic, and aromatic hydrocarbons, which
otherwise would not be biodegraded individually, can be oxidized through co-metabolism in an
oil mixture due to the abundance of other substrates that can be metabolized easily within the oil
(Atlas, 1981).
39


Microorganisms are known to produce extracellular biosurfactants to promote the formation of
oil-in-water emulsions that aid in the uptake and subsequent degradation of hydrocarbons (Desai
and Banat, 1997). The hydrophilic and hydrophobic components within the biosurfactants
emulsify hydrophobic hydrocarbons, and allow for transport into the hydrophilic intracellular
space for biodegradation (Southam et al., 2001). In addition, the fatty acid moieties of
biosurfactants promote the growth of microorganisms on the surface of oil droplets (Rosenberg
et al., 1979). Nikolopoulou and Kalogerakis (2008) reported that the use of rhamnolipid
biosurfactants increased removal of weathered petroleum hydrocarbons (96% removal of C19
C34 n-alkanes within a period of 18 days) and reduced the lag phase prior to the onset of
biodegradation. Saeki et al. (2009) showed that addition of biosurfactant JE1058BS to seawater
stimulated the degradation of weathered crude oil (ANS 521) by stimulating the activity of the
indigenous marine bacteria and facilitating the removal of crude oil from the surface of
contaminated marine sediments.

Field studies have shown that active microorganisms living in low-temperature environments are
dominated by two groups: psychrophilic and psychrotolerant, which are sometimes called
psychrotrophic (Atlas, 1984). As defined by Morita (1975), psychrophiles experience optimum
growth at less than 15C, with a maximum growth temperature below 20C and a minimum
growth temperature at or below 0C. Recent studies on pelagic bacteria in the coastal Alaskan
Arctic, Chukchi Sea, showed that psychrophiles were numerically dominant over psychrotolerant
organisms regardless of the nutrient status or algal bloom conditions of their surrounding waters
(Connelly et al., 2006). Despite living at these low temperatures, psychrophiles often have
metabolic rates comparable to those displayed by the mesophiles adapted to more moderate
temperatures. For example, Delille et al. (2009) reported that a temperature of 4C in the
Antarctic had little effect on biodegradation efficiency and that the nutrients, nitrogen and
phosphorus, were the limiting factors. Results obtained by Siron et al. (1995) indicated that the
temperature threshold for observing significant oil biodegradation was around 0C.

Decreases in solubility associated with low temperatures were considered to be a causal factor
for the cases of observed recalcitrance of hydrophobic compounds in cold-water and Arctic
40

conditions. However, recent reports have indicated that some bacteria may have adapted to the
low solubility of hydrophobic environmental chemicals (Deppe et al., 2005; Wick et al., 2002).
Indeed there is now evidence that hydrocarbon-degrading microbes may have novel uptake
mechanisms that enable them to degrade hydrocarbons at rates that exceed their rates of
dissolution in the aqueous phase (Leahy and Colwell, 1990; Thomas et al., 1986).

Throughout the world, the salinity of seawater averages about 3.5%. Salinity variations, albeit
small, are mainly caused by such factors as melting of ice, inflow of river water, evaporation,
rain, snowfall, wind, wave action, and ocean currents that cause horizontal and vertical mixing of
the saltwater (Lagerloef et al., 1995). Most marine species have an optimum salinity range of
2.53.5% (ZoBell, 1973) and species living in the transition environments are well adapted to
fluctuations in salinity. Microorganisms requiring salt for growth are referred to as halophiles.
Whereas halophilic hydrocarbon-metabolizing bacteria perform well in this salinity range, there
have been reports of the isolation of bacteria capable of degrading hydrocarbons above a salinity
of 3.5%. Bertrand et al. (1990) reported the isolation from a salt march of an extremely
halophilic archaea bacterium capable of degrading hydrocarbons in 3.5 M NaCl (20.4%), but not
below 1.8 M NaCl (10.5%). Diaz et al. (2008) reported the isolation of a bacterial consortium,
which mainly included members of the genera Marinobacter, Erwinia and Bacillus, from a crude
oil sample from the Cormorant field in the North Sea. This consortium was able to metabolize
petroleum hydrocarbons in a salinity range from 0 to 22% NaCl. Total oil degradation ranged
from 48% to 75%, with the greater degradation occurring at the lower salinities.

At the sea surface, wind and wave action maintain a constant supply of oxygen, thus aerobic
catabolism of hydrocarbons is usually the preferred biochemical pathway (Leahy and Colwell,
1990). Oxygen may become limiting in subsurface sediments and anoxic zones of water
columns. Oxygen limitation is also a concern for most fine-grained marine shorelines, freshwater
wetlands, mudflats and salt marshes (Venosa et al., 2002; Venosa and Zhu, 2003). It is
commonly believed that biodegradation rates under anaerobic conditions are almost negligible,
while aerobic biodegradation of hydrocarbons occurs rapidly. However, the importance of
anaerobic biodegradation should not be underestimated as it has been shown to be a major
process under certain conditions. In anoxic marine sediments, reductions of sulphate, Mn(IV)
41

and Fe(III) are the primary terminal electron-accepting processes (Canfield et al., 2005; Finke et
al., 2007). Results obtained by Coates et al. (1997) revealed that hydrocarbon degradation
coupled with sulphate reduction prevails in marine anoxic sediments. Information on how to
enhance microbial growth and biodegradation performance in anoxic marine sediments is still
limited and requires further study.

McFarlin et al. (2011a; 2011c) set out to determine the ability of indigenous microbes to
biodegrade crude oil in marine waters of the Beaufort and Chukchi Seas. Fresh and weathered
Alaska North Slope crude oil was tested with and without Corexit 9500. Biodegradation was
significant in all treatments at -1 to 2C, and dispersal with Corexit enhanced it. Degradation at
2C of 10 mg/L fresh crude was 37%. When chemically dispersed with 5% Corexit it was 56%,
and chemically dispersed with the addition of 0.5% Bushnell Haas was 66%. In the case of 12
mg/L of 20% weathered crude at -1C with 0.5% Bushnell Haas and without Corexit,
biodegradation was 46%, whereas with chemical dispersion it was 55%. Clearly, fresh crude oil
was more effectively biodegraded than the weathered.

With recent advances in analytical methods such as genomics, we are now able to determine the
potential of whole microbial communities for oil biodegradation at low temperatures. New
evidence as a result of advances in biotechnology (i.e. metagenomics) are now suggesting that oil
in the Arctic may be degraded at higher rates by indigenous organisms (that are adapted to their
environment) than previously thought. This is not surprising based on the fact that natural oil
seeps occur in the worlds oceans including the Arctic. Studies have shown that elevated
concentrations of hydrocarbons in the environment increase catabolic-gene copy numbers among
the microbial community (Heiss-Blanquet et al., 2005; Stapleton and Sayler, 2000; Whyte et al.,
2002).

Although various chemical and microbiological aspects of petroleum oil degradation in marine
systems have been relatively well studied in Arctic regions (Boehm et al., 1987; Haines and
Atlas, 1982; Hodson et al., 2002; Owens, 1994), there is a general lack of knowledge concerning
the diversity or abundance of the oil-degrading bacteria in relation to the extant hydrocarbon
pollution. The recent DWH spill has highlighted the importance of oil degradation in the
42

recovery of marine ecosystems impacted by crude oil spills. Hazen et al. (2010) reported that the
disappearance of residual oil in the Gulf of Mexico from the DWH spill was associated microbial
degradation processes based on the results of metagenomics. Temperature did not appear to be a
major limiting factor as significant rates of oil degradation were observed within the subsea
plume of dispersed oil at a depth of 1300 m (temperature below 4
o
C). There is a need for studies
using advanced genomic techniques to assess the significance of natural oil degradation
processes in the Arctic marine environment. Compared to the amount of knowledge of the
diversity of oil catabolic genes and levels of expression in oil-contaminated soils (Luz et al.,
2004; Whyte et al., 1997; Whyte et al., 2002), our understanding of this diversity in Arctic
marine systems is practically nil.

Unlike oil spills occurring at the sea surface, during the DWH spill, petroleum hydrocarbons
experienced a prolonged, buoyancy-driven ascent through the 1500 m water column (Hazen et
al., 2010). Consequently a unique set of processes affected the released hydrocarbons during
their trajectory in the deep sea. Some oil and gas never reached the sea surface, but instead
formed hydrocarbon-rich plumes within the cold waters present at ~1100 m depth, supporting an
active deep-sea microbial community (Hazen et al., 2010; Valentine et al., 2010). This
demonstrates the importance of interwoven chemical, physical and biological processes in
regulating the transport and fate of hydrocarbons in the deep marine environment. It would be
difficult to disentangle the role of natural processes from the effects of countermeasures such as
the use of dispersants when considering such questions as, How might droplet size from
dispersant applications affect biodegradation? Nevertheless, oil in the water column and on the
surface can be evaporated, biodegraded, diluted and dispersed. Significant evidence has shown
that gases have already biodegraded (Hazen et al., 2010).

There are new papers on the biodegradation of oil spills in cold water (Brakstad and Bonaunet,
2006; Brakstad et al., 2004; Brakstad et al., 2008; McFarlin et al., 2011b), in Arctic soil and
shorelines (Greer et al., 2010; Labbe et al., 2007; Margesin et al., 2003; Whyte et al., 2001;
Whyte et al., 2002; Whyte et al., 1999; Yergeau et al., 2009), and the persistence of oil in Arctic
shorelines (Boehm et al., 2007a; Boehm et al., 2008; Boehm et al., 2007b; Boufadel and Bobo,
43

2011; Boufadel et al., 2010; Guo et al., 2010; Li and Boufadel, 2010; Neff et al., 2006a; Neff et
al., 2011; Page et al., 2010; Page et al., 2002; Page et al., 2006; Xia and Boufadel, 2010).

Formation of Oil-Mineral Aggregates

Biodegradation of oil can be stimulated when the surface area of oil available for microbial
attack is increased (Prince, 1993). Swannell and Daniel (1999) adapted a technique (Delvigne
and Sweeney, 1988) to measure droplet sizes that allow for the determination of microbial
colonization of small droplets. Their results showed that many dispersed oil droplets were
colonized by bacteria and that substantial biodegradation took place after only 28 days of
dispersant addition to the microcosms. The increase in surface area amenable to bacterial
colonization occurs by the formation of an oil-in-water emulsion, where the oil is dispersed in
small droplets into the water column, which is promoted by the use of dispersants (Desai and
Banat, 1997). There has been much interest in recent years in the fate of oil droplets suspended
in the water column (Khelifa and Brown, 2010; Payne et al., 2003b; Stoffyn-Egli and Lee, 2002).
The mechanisms for the dispersion of oil into water as fine droplets are not well understood and
are one of the large unknowns for predicting oil spill behaviour. In the natural environment, oil
particles do not remain for long as discrete particles. They either recombine or coalesce, rising
back to the surface, or they can attract and incorporate suspended particulate matter (SPM) from
the water column (Owens and Lee, 2003) to form oil-mineral aggregates (OMA). These
aggregations are known by many other names in the literature: oil-suspended particulate matter
aggregates, clay-oil flocs, OMAs, and oil-fines interactions, to name a few.

Huang and Elliot (1977) described a process of clay-oil flocculation in which very fine
particles (alumina, silica, kaolinite) tended to adhere to oil droplets and prevent further
coalescence of the oil. Bassin and Ichiye (1977) showed that oils and clays form spontaneous
association colloids or colloidal electrolytes in the presence of dissolved salts. An OMA is
composed of an oil droplet coated with these kinds of fine particles. Typical sizes range from a
few to several hundred microns. The formation of OMA stabilizes the oil-water interface, with
the SPM acting as a surfactant. As stable particles this thermodynamically favours droplet
formation and enhances dispersion of oil into the water column.
44


The typical SPM that encourage OMA formation are clay and silicate particles suspended in the
water column, either from surface water outflows or from sediment stirred up from the bottom,
such as in beach surf. As mineral particles are much more dense (typically 2-4 times) than the oil
to which they are attached, OMA are less buoyant than naked oil droplets, and tend to sink.
Because the OMA are stable, the oil droplets do not tend to shed the particles and thus cannot
recoalesce to form a surface slick. This process of stabilizing oil droplets with fine clay particles
has formed the basis of an oil spill remediation strategy based on the dispersion of oil trapped
within shoreline sediments (Bragg and Yang, 1995; Lee et al., 2003; Lunel et al., 1996; Owens,
1999). Nutrient-bearing treatment products that promote OMA formation to enhance oil
biodegradation are also being engineered for application to open-water marine oil spills (Kjeilen-
Eilertsen et al., 2011).

Recent studies for the Canadian Coast Guard (Cloutier et al., 2005; Khelifa et al., 2005a), have
shown that OMA can form at near-freezing temperatures in seawater. With high-mixing energy
applied, the OMA readily form more than half within the first 10 minutes. Within 40 minutes
of turbulence, most of the free oil had converted fully to OMA. These laboratory results were
reconfirmed on a larger scale in a meso-scale basin (Blouin and Lee, 2007; Cloutier and Doyon,
2008) containing brackish water (18) with slush ice and broken ice. With strong turbulent
mixing, OMA particles formed readily in slush (mixed snow) ice and broken ice. Almost all of
the resulting particles were less than 1 mm in diameter. A mixing time varying between 20 and
30 minutes was sufficient to disperse about 50% of the spilled oil.

These trials were reproduced at full scale in January 2008, by the Canadian Coast Guard
icebreaker Martha Black, in the St. Lawrence River near Matane, Quebec. The icebreakers
propeller was used to create OMA in real ice conditions (Lee et al., 2009c). Several trials of 200
litres of fuel oil were mixed with chalk fines by the propeller of the icebreaker. The oil was seen
to be dispersed into the water column and did not resurface. A control test with no added
particles was observed to produce significant resurfacing oil. Samples collected from the water
column were confirmed by optical fluoro-microscopy to have OMA particles present. Water
45

samples taken back to the laboratory for tracking biodegradation revealed that more than half of
the total petroleum hydrocarbons (TPH) had been degraded after 56 days of incubation at 0.5C.

In further support of the concept of enhanced oil degradation by the formation of OMA, other
studies have shown that the adhesion of microbial cells to surfaces or particles offers protection
against physico-chemical stress and promotes increased catabolic potential compared to
suspended populations (Baror, 1990; Diaz, 2008). Diaz et al. (2008) showed that when the
salinity of the culture medium was over 20 g/L, a consortium of bacteria immobilized on
polypropylene fibers significantly enhanced the biodegradation rate of crude oil compared with
free-living cells. The authors concluded that through diffusion limitation the biofilm offered a
protective effect against salinity stress and toxicity, and that the oleophilic carrier increased the
oil availability for microbial attack. Brakstad and Bonaunet (2006) showed that oil adsorbed to
hydrophobic Fluortex fabrics was quickly utilized by a bacterial biofilm that formed on the
fabrics in seawater at the freezing point (0C). With the use of these types of adsorbents, not only
is spilled oil physically eliminated, but also biodegradation in challenging environments is
stimulated when cells form a biofilm on the hydrophobic surfaces.

Sedimentation

Most oils do not sink as their specific gravity is less than that of seawater. However, when
suspended particulate material including fine sedimentary material interacts with oil, the
conglomerate formed may be negatively buoyant. This natural process may occur in nearshore
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 it is well known that anaerobic biodegradation of hydrocarbons is a slow, poorly
understood process.

46

4. Oil Spills in Canada from Offshore Oil and Gas Activities

The United States Geological Survey estimates the Arctic to contain nearly ninety billion barrels
of oil of which nearly 10 billion is in the Amerasia Basin, an area that includes the Beaufort Sea
(Figure 6). The focus on oil exploration in the Canadian Beaufort began in the late 1960s and
early 1970s following successful exploration of Alaskas Prudhoe Bay (Matthews, 2011), and
from 1972 to 1989, 33 wells were drilled from artificial islands in shallow water. Between 1976
and 1989, drill-ships established another 31 wells for a total of 86 in waters up to 60 m deep,
with the only oil production occurring near the end of this period amounting to about 300,000
barrels of oil (Matthews, 2011).

Sverdrup
Basin
Figure 6 Canadian Arctic (source: Solar Navigator, 2011, http://www.solarnavigator.net).

Devon Energy in December 2005 drilled the exploratory offshore natural gas well, Paktoa C-60,
in 13 m of water in the MacKenzie delta, and completed it in 2006 (Energytoday, 2010; Lunan,
2005); however, no information was released for competitive reasons. Since then, no exploratory
drilling has occurred in the Canadian Beaufort region, although BP has commenced seismic
47

testing of the sea floor (Energytoday, 2010). In the Beaufort Sea and MacKenzie delta a total of
295 wells have been drilled between 1965 and 2009 (Hawkins, 2010).

Resources in the Sverdrup Basin (Figure 6) are estimated at 19.8 trillion cubic feet of gas
(CGPC, 2006) and 1.9 billion barrels of oil (Chen et al., 2000). Climate change has already
impacted the ice cover in the Sverdrup Basin, such that it there might be a point in time when
neither the use of ice platforms nor conventional offshore drilling techniques would be feasible
(Meneley, 2008).

Baffin Bay and Davis Strait (Figure 6) are shared with Greenland and may hold up to 18 billion
barrels of oil (Edmonton Journal, 2010). Three wells were drilled in Canadian waters offshore of
southern Baffin Island: Hekja O-71, Raleigh N-l 7 and Gjoa G-37 (INAC, 1995). Five wells (dry
and abandoned) were drilled in Danish waters on the west Greenland Shelf in Davis Strait,
approximately 500 km to the northeast (Figure 6), at the southern entrance to Baffin Bay, in
1976-77 (INAC, 1995). One well was drilled in the Strait in 2000 (Cairn, 2011).

Cairn Energy PLC of Scotland has been drilling in the Greenland waters of Baffin Bay since the
summer of 2010 (CBC, 2011). Three exploratory wells were completed in 2010 in the Sigguk
block, offshore West Greenland in 300-500 m of water (Cairn, 2011). As of 3 June 2011, drilling
in Davis Strait has commenced on the AT-7 well, Atammik Block, about 160 kilometres offshore
of Nuuk (Figure 7), and the LF-7 well in the Lady Franklin Block about 300 km offshore of
Nuuk (Figure 7). The AT-7 and LF-7 prospects are at water depths of 905 and 989 m
respectively (Cairn, 2011). Another 2 are scheduled for the summer of 2011: one in each of the
Napariaq and Eqqua blocks in the southern Baffin Bay west Disko area (Figure 7). It is
anticipated that these 4 wells, located at subsurface depths between 2500 and 4750 m, will
produce 3.2 billion barrels of oil equivalent. It is hoped that oil will be the product, but it could
be gas. The uncertainty reflects the lack of data in this frontier (Cairn, 2011).
48


Figure 7 Exploration parcels of Cairn Energy in the Greenland offshore (Cairn, 2011).


As of June 2011, there are no reported marine oil spills in the Beaufort Sea, Sverdrup Basin,
Baffin Bay or Davis Strait. Over the past 40 years, there have been no recorded oil spills caused
by an offshore well blowout in the U.S. Alaskan or Canadian Arctic marine waters.

Increasing oil exploration activities in northern latitudes in the context of the recent blowout of
the DWH well have increased concerns over the safety of operations in the Arctic. Oil
exploration activities are complex and they require the involvement of various types of
equipment such as rigs, vessels and aircraft to support operations. The complexity and risk of
such operations is compounded in the Arctic by the presence of ice, cold temperatures and by the
lack of daylight during certain periods of the year. These difficult conditions are challenging for
any operations conducted in this region and are increasing the probability of having an oil spill in
the Arctic.

This section describes the current knowledge on the various types of oils that could be released
in the Canadian Arctic in case of an incident during oil exploration activities. Using simple
scenarios, we will describe some of the environmental risks from their release and how their
behaviour could impact response options.

49

4.1 Oil Types from Vessel Operations

The type of vessels currently operating in the Arctic and a general list of petroleum hydrocarbon
based fuels and other products including lubricants that they typically carry are summarized in
Table 2. To be prepared for a spill, it is important to understand the types of products that may be

Table 2 Vessel types operating in Arctic waters and oil types carried onboard.
Oil Type Vessel Type
Transported as Fuel Transported as Cargo
Examples of Tank Capacity
Tankers Bunker C, Intermediate
Fuel Oil (IFO)
Crude, Bunker C, IFO,
Diesel, Gasoline
Range: 92 000m
3
(total cargo
tanks) to 12 000m
3
(1 tank)
Supply and
anchor
handling
ships
Marine Gasoil (MGO) Low Sulphur Diesel
(LSD), oil based mud,
glycol, mineral oil
Range: 570 m
3
(total fuel tanks) to
80 m
3
(1 tank)
Tugs/Barges MGO Gasoline, Bunker C Tug range: 300 m
3
(total fuel tank)
to 75 m
3
(1 tank)

Barge range: 1383 m
3
(total cargo
tanks) to 170 m
3
(1 tank)
Cargo/Fuel IFO Diesel Range: 24 309 m
3
(total cargo
tank) to 3000 m
3
(1 tank)
Icebreakers MGO Aviation fuel

Range: 4800 m
3
(total fuel tanks)
to 400 m
3
(1 tank)
Cruise ships IFO, Diesel None Range: 4800 m
3
(total fuel tanks)
to 400 m
3
(1 tank)
Fishing
vessels
MGO

Range: 220 m
3
(total fuel tank) to
35 m
3
(1 tank)
Re-supply
ships
(community)
Diesel IFO, Gasoline 17 207 m
3
(total cargo tank) to
2800 m
3
(1 tank)
Military MGO Aviation fuel

Unknown
Research Diesel Aviation fuel

Range: 4800 m
3
(total fuel tanks)
to 400 m
3
(1 tank)
50


released and potential volumes. The following is based on past marine traffic in the Canadian
Arctic. These numbers will obviously increase as more traffic enters these waters for oil or
mining development and does not include rigs and platforms, which would introduce additional
spill potential. The carriage of crude oil by tanker would also increase the volume of product as
well as the risk.

All of these vessels are also transporting an unknown amount of hydraulic oil, lube oil, and waste
oil onboard as a result of the operation of various pieces of equipment such as cranes, derricks,
thrusters, winches and lifts. However, the quantities of these types of oil are likely to be at a
maximum of 1 m
3
. Ships will also generate oily water during their operation. This must be
retained on board until it can be discharged ashore (bilge water, contaminated ballast water,
waste oil and tank washings). Some oil will be released by being passed through the ships oily
water separator but these releases will not exceed 15 ppm in Arctic waters, the normal limit for
most equipment. Another potential source of discharge from ships are the tail shaft seals, which
may leak when the seals become deformed by the formation and expansion of ice. An accidental
release of these oils would not represent a significant risk because of the small quantities
involved, but the consequent increased possibility of reduced thrust and engine failure, leading to
grounding and collision, increases the risk of a serious oil spill.

4.2 Oil Types from Oil Platform Operations

Oil platforms are complex structures used to either conduct exploratory drilling or extract and
process oil or gas before it is transported to refineries or oil depots. Two potential sources for an
oil spill include crude oil from drilling activities and oil used for the operation of equipment.

51

Crude Oil from Drilling Activities

Crude oils that have been discovered in the Canadian Arctic are listed in (Table 1). Most of these
oils are generally considered medium to light oils and could be spilled during drilling activities,
as a result of a well blowout or during transport by tankers. While the probability of a spill is
low, depending on the characteristics of the formation being exploited, the potential quantities
spilled could be very high as demonstrated during the recent DWH blowout, and the PTTEP
Montara West Atlas spill in the Timor Sea in August 2009.

Petroleum Products used in Operations

Various types of equipment using oil are found on platforms such as generators, cranes, engines,
etc. These all use oil products either as fuel or for lubrication. In most cases, diesel fuel, engine
oil and hydraulic oil are the main oil types necessary for platform operation. Stored quantities of
these, particularly for engine and hydraulic oil, are likely to be limited. Diesel is often used for
power generation and might be found in larger quantities ranging from approximately 100 m
3
to
200 m
3
. Helicopter operations to transport personnel and equipment require the storage of
aviation gas in small amounts on the platform as well, generally less than 10 m
3
.

4.3 Overview of Oil Spill Risks in the Canadian Arctic

Defining Risk

The many interpretations of the term risk has led to different approaches to risk management in
different fields (Hubbard, 2009). Examples include the following.
The ISO 31000 (2009)/ISO Guide 73 definition of risk is the effect of uncertainty on
objectives. In this definition, uncertainties include events which may or may not happen,
and uncertainties caused by a lack of information or ambiguity. This definition also
includes both negative and positive impacts on objectives.
52

Risk can be seen as relating to the probability of uncertain future events. For example,
according to Factor Analysis of Information Risk, risk is the probable frequency and
probable magnitude of future loss (Jones, 2005).
OHSAS (Occupational Health and Safety Advisory Services: OHSAS 18001:2007)
defines risk as the product of the probability of a hazard resulting in an adverse event,
multiplied by the severity of the event.

In statistics, the notion of risk is often modelled as the expected value of some outcome seen as
undesirable. This combines the probabilities of various possible events and some assessment of
the corresponding harms into a single value. In a formula that can be used in the simple case of a
binary possibility (accident or no accident), risk is then:

Risk = (probability of the accident occurring) (expected loss in case of the accident).

Understanding the risks associated with an activity requires (1) the ability to predict the
likelihood or probability of such an event occurring, and (2) the ability to anticipate the potential
adverse consequences of such an event. The magnitude of accidental releases of oil may range
from small spills associated with pipes, tanks or vessels to a catastrophic well blowout. Based on
historical records from Canadian and global databases maintained by regulators and industry,
very large oil spills or blowouts are low probability events, but the consequences may be
disastrous.

The risk of oil spills in the Arctic is anticipated to increase due to
global climate change that may eventually lead to year round marine transport via the
North West passage and the expansion of industrial developments, and
the expansion of exploration drilling outside of the traditional summer open water period
as operations become more technically feasible with advances in marine technology
including the support of novel ice management.
Should an event occur, it is important to note that all Arctic oil spill countermeasures are
hampered by the harsh environment, poor access and a lack of nearby support facilities (Sergy
and Blackall, 1987).

53

Incidents from Vessel Operation

Shipping incident statistics from the Transportation Safety Board of Canada provide a better
understanding of the potential risks of having a spill from vessels operating in the Arctic. In 2009
in Canadian waters, the most common type of vessel to have reported marine accidents were
fishing vessels, which outnumber the other vessel types and make up over half of the total
number of Canadian registered vessels (Figure 8). Fishing vessels are followed by tugs and
barges, and then bulk carrier and other bulk operations in the number of accidents.


Figure 8 Accidents by ship type; OBO = other bulk operations (source: Transportation Safety Board of
Canada).


The most common cause of vessel accidents in Canadian waters since 2004 has been grounding
(Figure 9). This is followed by striking and fire/explosions. All of these categories have the
potential for causing a release of product.

54


Figure 9 Types of accidents (Source: Transportation Safety Board of Canada).


There are less vessel accidents reported in the Arctic region than in all other regions of Canada
(Table 3); however, there is also less traffic (Figure 10). The number of accidents will likely
increase as shipping increases in the Arctic region. Tugs, barges and cargo vessels reported the
most shipping accidents in the Arctic since 2000.

The most common reason for ship source spills in the Arctic is oil transfers by hose, which is
unlikely to cause a very large spill because of preventive systems in place during transfer
operations. Other likely causes would be spills due to major accidents that will cause hull
damage such as groundings because of poor nautical charts or difficult navigation and striking
especially with multi-year ice. For vessels currently operating in the Canadian Arctic, the most
significant threat is from oil tanker operation where potential spill quantities could be significant,
i.e. in the range of 3000 to 4000 m
3
(if one cargo tank was breached) while spills from cruise
ships and icebreakers would be in the range of 1000 m
3
. Supply vessels, which are used
extensively during oil drilling exploration projects, represent a lesser threat because of limited
fuel storage capacity with a range of 300 to 600 m
3
and also because of the less persistent fuel
type, such as marine gasoil (MGO) or marine diesel, that they usually transport.
55





Figure 10 Canadian Arctic shipping routes.

56

Table 3 Reported vessel accidents in the Canadian Arctic.
YEAR 2000 2001 2002 2003 2004 2005 2006 2007 2008 2009
Shipping Accidents 8 4 2 7 6 0 4 3 5 6
On-Board Accidents 0 0 0 5 1 0 3 1 1 0
Type of Vessel Involved
Cargo 3 1 0 3 0 0 0 0 1 0
Bulk Carrier/OBO 2 2 0 0 0 0 0 0 0 0
Tanker 0 0 0 1 1 0 0 1 0 0
Ferry/Passenger 1 0 0 0 0 0 2 0 0 0
Tug/Barge 0 1 2 0 2 0 1 0 2 3
Fishing 1 0 0 0 0 0 0 0 1 0
Other 1 0 0 3 5 0 1 3 1 3



Incidents from Oil Platform Operations

Oil platform operations present a variety of spill risks. Accidental petroleum discharges from
platforms contribute 0.07% of the total petroleum input to the worlds oceans (NRC, 2003). Well
blowout is the most significant spill risk. This could spill large quantities of crude oil as was
demonstrated by the IXTOC I in 1979 and DWH in 2010. However, these incidents are rare
especially in Arctic waters where production and exploration activities are far less prevalent than
in more temperate climates. Operational spills are more common and these typically include
releases of crude, diesel, hydraulic and lube oil. Based on statistics from the Canada-Nova Scotia
Offshore Petroleum Board (CNSOPB) and Canada-Newfoundland and Labrador Offshore
Petroleum Board (CNLOPB), the source of most spills that are in excess of 150 L are flaring,
transfer operations, leaks in fittings and valves or from the riser slip joint. Except for the Uniacke
G-72 gas blowout near Sable Island in 1984, which discharged some 1,500 barrels of condensate
into the marine environment, Canada has never experienced an offshore well blowout involving
discharges of petroleum oil.

57

Statistics from Nova Scotia and Newfoundland-Labrador provide an indication of the frequency
of spills and the quantities potentially involved from platform operations. Few spills are
recorded every year from offshore operations in both provinces as shown by Figure 11 and
Figure 12. The majority of spills that have occurred are less than 150 L. These small spills
outnumber larger spills (>150 L) by 4 to 1 in Newfoundland-Labrador, and 9 to 1 in Nova
Scotia. The most common sources for small spills are hoses, leaks from stern tubes and thrusters,
and drain systems.


Figure 11 Number of spills per year in Nova Scotia from offshore operations.



Figure 12 Number of spills per year in Newfoundland-Labrador from offshore operations.
58


From January, 1994, to August, 2002, five oil spills, including spills of synthetic oil-based
drilling mud, occurred during exploration drilling on the Scotia Shelf that resulted from
exploration and production activities (CNSOPB, 2010). In the Newfoundland and Labrador
offshore petroleum region, in 2004, a 1000 bbl crude oil spill occurred from the Terra Nova
platform due to equipment failure. While there were 296 spill events associated with exploration
and production activities in the Newfoundland and Labrador region from 1997 and 2007, the
total volume of hydrocarbons released was less than 1200 bbl (CNLOPB, 2010). Figure 13 and
Figure 14 provide an indication of spilled quantities from offshore operations. Between 1997 and
2011 the total volume of combined product spilled during offshore operations in Newfoundland-
Labrador was 454,681 L, and 555,239 L in Nova Scotia between 1999 and 2010. Spill volumes
increased during periods of development and production.


Figure 13 Total spill volume in litres per year off Nova Scotia.


Records show that the number of operational and accidental spills and the quantities released on
a yearly basis from offshore activities currently operating on the east coast of Canada are fairly
low and limited.

59


Figure 14 Total spill volume in litres per year off Newfoundland.


The rate of blowouts is very low in the oil exploration and development industry. For example,
in the Gulf of Mexico from 1979 through to 1998 there were 19,821 wells drilled, with 118 wells
resulting in uncontrolled flows or blowouts. This results in a 0.6% occurrence rate of blowouts
for the region during this time. Furthermore, the majority of reported events involved the
diversion of gas. Due to differences in defining a blowout, the American Petroleum Institute
(2009) states that there has only been 17 marine well blowouts in the U.S. since 1964, which
resulted in a total of 249,000 bbl spilled. The largest of these blowouts occurred from the Alpha
Well 21 off Santa Barbara, California, which spilled 100,000 bbl. In general, spill volumes have
been small, with 50% of the well blowouts involving 400 bbl of oil or less. With advances in
technology and improved operational safety measures, data from U.S. operations has indicated a
decrease in combined spill volumes from offshore supply vessels, pipelines, and platforms, from
30,400 bbl y
-1
(from 1969-1977) to 3,900 bbl y
-1
(from 1998-2007), resulting in an 87%
reduction in the average annual spill volume since the 1970s (American Petroleum Institute,
2009). Historical spill and blowout events support the prediction of a low probability of spill and
blowout occurrence for new oil and gas drilling exploration and production activities (Table 4).




60

Table 4 Historical large oil spills in barrels (bbl) from offshore well blowouts (source: Oil Spill Intelligence
Report database).
Offshore Region
Spill Size
(bbl)
Year Type of Activity
Mexico (IXTOC I) 3,000,000 1979 Exploratory Drilling
Dubai 2,000,000 1973 Development Drilling
Mexico 247,000 1986 Work over
Nigeria 200,000 1980 Development Drilling
North Sea/Norway 158,000 1977 Work over
Iran 100,000 1980 Development Drilling
U.S.A., Santa Barbara 77,000 1969 Production
Saudi Arabia 60,000 1980 Exploratory Drilling
Mexico 56,000 1987 Exploratory Drilling
U.S.A., S. Timbalier 26 53,000 1970 Wireline
U.S.A., Main Pass 41 30,000 1970 Production
U.S.A., Timbalier
Bay/Greenhill
11,500 1992 Production
Trinidad Development 10,000 1973 Drilling

The Canada-Nova Scotia Offshore Petroeum Board has 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 times
higher possibility of occurrence (Hurley and Ellis, 2004). Hurley and Ellis (Hurley and Ellis,
2004) provided expected frequencies of exploration well blowouts based on the historical
frequency of blowouts per wells drilled. They calculated that the predicted frequency of an
extremely large blowout (>150,000 bbl) was one every 17,500 years. The frequency of a very
large spill (>10,000 bbl) was one every 5800 years, and of a large spill (>1000 bbl) was one
every 4400 years.

Based on statistical data from the offshore of Nova Scotia, presented by Hurley and Ellis (2004),
the highest frequencies of oil spills are for the smaller, platform-based spills. Spills that are less
than one barrel in size may occur once every two years. Oil spills during exploration, which are
61

larger than one barrel but less than 50 bbl, have an approximately 1-in-10 or 1-in-20 chance of
occurrence per year. In fact, the CNSOPB concluded that their original prediction of 0.5 spills
per year for the Sable Offshore Energy Project was largely accurate when compared with the
observed spill occurrence of 0.57 spills per year.

The stasticical data from all of these offshore oil and gas sites provide an indication of the
likelihood of having spills in the Arctic from exploration and production activities. We can see
that spills are limited in both number and size. While there are advances in technology and
regulations for the safety of operations, recognition must be given to the unique challenges for
offshore oil and gas exploration and production operations in the Arctic environment. Although
a significant spill of crude oil resulting from a well blowout has never been recorded in Canada,
it remains a threat that cannot be overlooked.

4.4 Modelling Spill Scenarios and Oil Behaviour

Predictive models of spill scenarios have been developed to illustrate the expected behaviour of
the most commonly used oil in the Canadian Arctic. From the information obtained it is evident
that spills of crude oil, intermediate fuel oil (IFO), aviation fuel and MGO are the most likely to
be spilled in quantities that would require some sort of response to take place. We recognize that
other oils such as lube oil and hydraulic oil can also be spilled but in quantities that would be
minimal and therefore not requiring any specific response other than monitoring.

The Automated Data Inquiry for Oil Spills (ADIOS2) software package is an oil weathering
model provided by the National Oceanic and Atmospheric Administration
(http://response.restoration.noaa.gov/adios). The ADIOS2 database includes estimates of the
physical properties of oils and products compiled from different sources, including industry,
Environment Canada, and the U.S. Department of Energy. The model uses mathematical
equations and information from the database to predict changes over time in the density,
viscosity, and water content of an oil or product, the rate at which it evaporates from the sea
surface and disperses into the water, and the rate at which an oil-in-water emulsion may form.
Output of the model in graphical and textual format can be used to address questions that
62

typically arise during spill response and cleanup. For example, by predicting change in oils
viscosity (resistance to flow) over time, ADIOS2 can provide information on whether chemical
oil dispersants can be used with success. For comparative purposes, oil weathering scenarios
were prepared using ADIOS2 software in the following subsections of this report with a water
temperature set at 0C and wind speed of 5m/s which correspond to a moderate breeze or a
Beaufort 4 sea state. The ADIOS2 model cannot compute for oil-ice interaction, which is a
significant limitation when evaluating behaviour in Arctic conditions. However, even with this
limitation in mind, this model provides enough information to evaluate the expected behaviour of
oil during specific spill scenarios.

Scenario for Spills of Crude Oil

One of the most plausible scenarios for a crude oil spill in the Canadian Arctic is that associated
with tanker operations. This scenario is more likely to happen than a catastrophic well blowout
as tankers have to transit waters with many navigational hazards such as the presence of ice floes
and constant darkness at certain times of the year. To look at the potential behaviour of crude oil,
we considered a spill of 4000 m
3
of Amauligak crude oil over a one-hour period from a tanker as
a result of a collision with ice or any other structures. This quantity is representative of the loss
of one tank from a typical tanker currently operating in the Canadian Arctic.

From the ADIOS2 weathering model, we can see that a small proportion of this oil would
evaporate and disperse (approximately 20%) within the first 36 hours leaving most of the spilled
oil to be recovered during the response (Figure 15). However, during the same period the
viscosity (Figure 16) and the density (Figure 17) of this oil are also increasing significantly thus
reducing the window of opportunity available for some response techniques such as dispersant
application. With a kinematic viscosity above 10,000 cSt and a density slightly above 1.00 after
36 hours it is possible that this oil would be slightly submerged and non-dispersible. In that case,
only mechanical skimmers and pumps might be available as cleanup options. Given the logistical
challenges of responding in the Arctic, it would be very difficult to have an efficient response for
a spill of this type. As mentioned before, oil-ice interactions are not considered by ADIOS2 in its
calculation. However, it can be assumed that in ice conditions and very cold air temperatures,
63

less oil would be naturally dispersed and viscosity and density would be higher, increasing the
difficulties in mounting an at-sea response. In a scenario such as this one, oil would likely drift
with sea currents and mix with ice floes before reaching the shoreline where recovery efforts
could be conducted in a more efficient manner. It is obvious that a well blowout spilling large
quantities of this oil over a long period of time would cause substantial difficulties. However,
once response equipment was on location, it is possible that response efforts could become
efficient, as fresh oil would be constantly rising to the surface, maintaining the window of
opportunity for successful operations. In spite of this, logistical challenges such as waste
management, and health and safety issues would be very difficult to overcome.




Figure 15 Oil budget for a spill of 4000 m
3
of Amauligak crude oil as calculated by ADIOS2.

64


Figure 16 Predicted change in viscosity for Amauligak crude oil as calculated by ADIOS2.


Figure 17 Predicted change in density for Amauligak crude oil as calculated by ADIOS

65

Scenario for Spills of Intermediate Fuel Oil

Tankers and cargo ships, among others, currently operating in the Arctic use IFO as their
propulsion fuel. IFO is a heavier oil with a pour point of -10C which could be a problem in
Arctic conditions as this oil could become semi solid at temperatures commonly observed in the
North. IFO could be released into the environment through hull damage in a collision. By way of
example in which a severe collision released approximately 1000 m
3
of IFO 180, using the
ADIOS2 weathering model, most of the oil (87%) would remain in the environment and
weathering processes would have little effect on this oil (Figure 18).



Figure 18 Oil budget for a spill of 1000 m
3
of IFO 180 as calculated by ADIOS2.

Natural dispersion and evaporation are marginal weathering processes with this type of oil
especially in cold water environments such as the Arctic. Recovery of this oil would be
necessary using mechanical techniques, as dispersants would not be efficient given the high
kinematic viscosity that is well above 10,000 cSt at the time of the spill, and reaching more than
50,000 cSt after a few days (Figure 19). High viscosity coupled with low Arctic temperatures
66

would make any recovery efforts using pumps and skimmers very difficult due to freeze up of
water intakes and flow rate issues with the equipment. Similar to a spill of crude oil, most of the
IFO would likely drift with ice and sea currents and eventually reach the shoreline where further
recovery could be attempted.


Figure 19 Predicted change in viscosity for IFO 180 as calculated by ADIOS2.

Scenario for Spills of Marine Gasoil

Many ships currently operating in the Canadian Arctic also use marine gasoil (MGO) as
propulsion fuel. This fuel is also used by platforms mainly for power generation. Most vessels
using MGO are smaller vessels such as supply ships and icebreakers. MGO is a light oil with an
API gravity of approximately 31 and a pour point of -12C. Spills of MGO are likely to be
caused by collision between a vessel and ice floes or other structures. However, as opposed to
crude oil and IFO spills, the quantities involved are likely to be smaller since the types of vessels
currently using this fuel have less tonnage than ships using other fuels. A plausible scenario
67

would be damage to the hull resulting in the release of 100 m
3
of MGO from one of the fuel
tanks onboard a vessel. The oil budget calculated with ADIOS2 indicates that the combined
processes of natural dispersion and evaporation would weather most of the spilled MGO (Figure
20). It is also likely that a proportion of this oil would be easily absorbed by snow (if spilled on
ice) as its viscosity remains well below 10,000 cSt (reaching a maximum of approximately
50,000 cSt) even in Arctic conditions. The low viscosity keeps this oil amenable to most
response techniques such as dispersants and mechanical recovery, therefore representing less of
an operational challenge when compared to crude oil or IFO. Recovery of this type of spill
would likely be easier but would still remain particularly difficult given the Arctic setting. This
expected behaviour is one of the reasons why Norway has made it mandatory for all ship traffic
in its northern regions to use MGO as fuel in order to minimize the consequences of an
accidental spill.



Figure 20 Oil budget for a spill of 100 m
3
of MGO as calculated by ADIOS2


68

Scenario for Spills of Aviation Fuel (Jet Fuel)

Jet fuel is very light oil primarily used for air travel and support operations to transport personnel
and equipment throughout the Arctic. It has an API gravity of 43. Stored quantities on
platforms or on icebreakers are generally of a maximum of 10 m
3
(although significantly larger
quantities may be transported via cargo ship as indicated in Table 2). For discussion purposes, if
10 m
3
of jet fuel is spilled from a storage tank into the sea, ADIOS2 modelling shows that this
very light oil would be entirely dispersed and evaporated within the first 12-20 hours after
release into the water (Figure 21). If it was spilled on ice and snow, a similar pattern would be
expected as evaporation would be the primary weathering process. Even if evaporation on ice
would likely be slower than on water, the ultimate fate of that spill would be similar. Because of
this behaviour, jet fuel spills do not represent a significant environmental issue due to the very
low persistence of that oil. In dealing with a response to such a spill it is unlikely that significant
recovery efforts would be required other than possibly using absorbents in the vicinity of specific
environmental sensitivities. Monitoring of natural attenuation (natural recovery) would be the
spill response strategy.


Figure 21 Oil budget for a spill of 10 m
3
of jet fuel as calculated by ADIOS2.
69

Recommendations from Modelling Exercises

These models of various scenarios illustrate to some degree the risk involving an oil spill
because of increasing oil exploration and production activities in the Canadian Arctic. Increased
vessel traffic of multiple ship types and drilling operations would be the likely sources of spills.
With these activities, many different oil types will be transiting the Arctic either as fuel for ships
and equipment or as cargo for tankers. Crude oil is likely to be the most transported oil in this
region using large tankers. The potential quantities from a spill could be quite significant and
pose serious challenges to those in charge of responding to such an incident because of the
logistical issues of working in the Arctic and the presence of ice. A catastrophic event such as a
well blowout would bring supplemental challenges and difficulties because of the potential
quantities involved and the extent of such a spill. However, the most likely event resulting in an
oil spill in the Arctic is a collision between a ship and ice floes, another ship, or any other
structures in the region. A breach in the hull could release fuel and cargo into the environment in
significant quantities. It is also apparent that a spill of IFO would represent a greater threat to the
environment because of its high persistence and high viscosity rendering response very difficult.
Crude oils found in the Canadian Arctic are for the most part medium to light crudes and their
relative persistence in the environment is less than IFO.

As noted previously, oil behaviour in the Arctic is mainly guided by the presence or absence of
ice and by the low temperatures. In the absence of ice, oil will behave as it would at lower
latitudes with well-known processes such as natural dispersion, spreading, evaporation and
emulsification taking place over time. Many studies conducted world-wide have documented
well how these processes occur, changing the behaviour of oil and having a direct impact on
response methods that can be used during a spill. In the case of spills in ice covered waters
similar processes are take place but are directly affected by ice. Usually, oil-weathering
processes are slowed by ice and cold temperatures prolonging the window of opportunity for
some response techniques such as dispersants. Over the last 40 years many research projects
have been conducted to document the behaviour of oil with different ice coverage and the
efficiency of oil spill response techniques in Arctic conditions (See Section 3. Oil Spills in Arctic
Waters). These studies have provided us with a good understanding of how oil will behave and
70

how efficient most response techniques would be in Arctic conditions. In the context of the
Canadian Arctic, we have identified four areas where some research could be undertaken.
1) Oil behaviour in a subsea release
Little information exists on oil behaviour in the Arctic in the case of a
subsea release such as a well blowout. This would assist responders in
better understanding how the oil would behave under such circumstances
and adjust response plans and methods accordingly.
2) Hydrographic data and drifting models (Drozdowski et al., 2011)
Better hydrographic data in the Canadian Arctic should be collected and
integrated with oil spill trajectory models that would take into
consideration the presence and the influence of various ice coverage on
drifting oil.
3) Improving oil weathering models to include Arctic conditions and oil-ice
interaction
Very few oil weathering models take into consideration the oil-ice
interaction and the effect of Arctic conditions on oil weathering. To our
knowledge, only two models exist that consider the influence of ice on
weathering processes. Arctic conditions and oil-ice interactions should be
integrated into mainstream oil modelling software to ensure easy access of
the information to oil spill responders.
4) Improving detection of oil in or under ice
Even after many years of research, detection of oil in or under ice is still a
partially resolved issue. Research into this area should continue in order to
provide more efficient methodologies for responders.

As production and exploration in the Canadian Arctic is potentially increasing in the coming
years, many challenges remain to be able to efficiently respond in northern environments.
Although oil fate and behaviour processes are relatively well known there is still a need for more
detailed assessments in order to provide on-scene commanders and responders with a better
understanding of the behaviour of the oil released and applicable strategies in Arctic conditions.
71

5. Offshore Arctic Oil Spill Response Options

5.1 Development of Policies, Regulations and Capacity

Governments of Arctic coastal states require industry to ensure a high level of environmental
protection. For example, a number of preventive measures related to maritime safety and oil
spill response have been implemented in the Norwegian part of the Barents Sea (Ly and
Bjerkemo, 2011) where measures in surveillance, monitoring, maritime infrastructure and
services have been established along with oil spill response measures. In the United States, from
1993 to 2009, the Bureau of Ocean Energy Management, Regulation, and Enforcement (formerly
the Minerals Management Service) initiated and conducted 31 projects related to improving the
equipment (Mullin, 2011), technologies and methodologies for the prompt identification and
removal of oil in Arctic environments and to provide regulators with the scientific data necessary
to support regulatory decisions pertaining to permitting and approving plans, safety and pollution
inspections, enforcement actions, training requirements, and emergency response. In Canada,
national core funding for environmental research in the Arctic related to the development of
offshore oil and gas activities and spill response countermeasures has been supported by Natural
Resources Canada under the Program of Energy Research and Development and the
Environmental Sciences Research Funds.

The United States Coast Guards current capacity to mount an effective and timely response to a
major oil spill in the Arctic Ocean is limited (Lally, 2011). Currently, the U.S. Coast Guard
surface assets suited to respond to an oil spill are typically at least seven days away. If
conditions require an icebreaker escort, the response time could be 10 days or more. To meet the
many challenges present in the Arctic, all oil spill response technologies need to be considered as
response options including mechanical equipment, ISB, and chemical dispersants.

The Canadian Coast Guard (CCG) has a responsibility to monitor the effectiveness of the private
sector response to ship-source spills in Canadian waters, assume control of the incident if
necessary or deploy CCG resources to respond to spills for which the polluter is unknown,
72

unwilling or unable to respond. With the resources available, the CCG undertakes activities to
help prepare Arctic communities to respond to major pollution incidents. A National Response
Team, comprised of all CCG emergency response personnel and assets, is in place to provide a
coordinated national response to a marine pollution incident. The CCG also provides training in
the Arctic on how to effectively respond to emergencies. Communities are provided with
containment packs to support response capabilities. Basline studies in physical, chemical,
geological and biological oceanography for emerging oil and gas exploration and development
operations in the Canadian Arctic are supported by both industry and academia under the Arctic
Net Centre of Excellence which has a mandate to translate our growing knowledge on the
changing Arctic into impact assessments, national policies and adaptation strategies.

The complexity of Arctic oil spill response operations is compounded by prolonged periods of
darkness, cold temperatures, environmental sensitivities, indigenous peoples and their culture,
distant infrastructure and remoteness, the presence of seasonal dynamic sea ice offshore, and the
generally high cost of doing business (Engelhardt, 1994; Velez et al., 2011). Oil spill response in
ice-free seasons can be comparable to the response in other parts of the world, with the exception
of working in lower temperatures and extended daylight hours. Conventional oil spill
containment and mechanical recovery systems used in temperate waters are often constrained by
the presence of ice. In monitoring and risk assessments, remote sensing capacity is limited, and
the modelling of the fate of oil is complicated by a lack of adequate, detailed knowledge of
under-ice currents. While ice cover can hamper the access of responders to the oil, it can also
restrict the movement of oil on the sea surface and thus be considered an asset to certain
countermeasure applications such as in situ burning (ISB).

The first Norwegian oil field in the subarctic Barents Sea region will start production in 2013,
and crude oil transport along the environmentally vulnerable coast is increasing. A collaborative
program between industry representatives and national oil spill responders has recently been
established in Norway to addresses the publics demand for improved oil spill response
technologies for use in northern waters. The program includes 18 technical challenges divided
into four categories: (1) offshore oil spill recovery, (2) dispersant application technology, (3)
remote sensing technology, and (4) coastal and shoreline response.
73


Shell has undertaken different steps to implement best practices during the preparation for
exploratory drilling in the Alaskan Beaufort Sea and Chukchi Sea (Broje et al., 2011). Spill
prevention measures include continuous 24 hour monitoring of drilling activities through a real
time operation centre, advanced weather monitoring and forecasting and mechanical barriers.
Comprehensive contingency planning incorporates Arctic-capable response equipment, on-site
oil spill response vessels, as well as offshore, nearshore, and on-shore response strategies. A
dedicated Science Program is established to collect environmental baseline data (metocean, ice,
biological, and shoreline) and continued ecosystem-based monitoring and assessment, and
research and development of spill response techniques and advancement. Stakeholders are
engaged with communities along the Alaska North Slope and marine mammal subsistence user
groups.

5.2 Mechanical Containment and Recovery

Responses to spills from vessels, blowouts and pipelines include controlling slicks at source and
removing oil that escapes initial containment. The objective of both operations is to minimize the
spreading of spilled oil and adverse environmental impacts. Containment booms and skimmers
are deployed in either a mobile or stationary mode to intercept, control and recover oil on the sea
surface. When responding to an oil spill in Arctic conditions, the first step is to identify the
physical properties of the oil, particularly, the pour point. If the pour point is 5 to 10 degrees
above the water temperature, there is a strong possibility that the oil has already solidified. Nets
and other collection devices may be required for recovery. If the pour point of the oil is below
the water temperature and if currents and wind conditions allow, then booms and skimmers may
be applicable for use.

Response to spills in broken ice frequently requires strategies to deal with the moving ice.
Factors to consider are dramatic changes in ice concentration, frequent wind shift in direction
and speed, changing relative position of oil and ice, short daylight periods, diminished vessel
mobility and restricted operations due to the very cold temperatures (Solsberg et al., 2002). Ice
management techniques in the freeze-and-thaw seasons should also be part of the oil spill
74

response planning for a particular region. The major challenges are the development of
equipment for better detection and collection of oil spills in darkness and low visibility, icing of
the oil spill response equipment, and the human factor due to the low temperatures and harsh
weather (Juurmaa, 2006).

In Finland, mechanical recovery is preferred due to recommendations of the Baltic Marine
Environment Protection Commission, where chemical agents and ISB can be used only in special
and restricted circumstances (Lampela and Jolma, 2011). The methods and techniques developed
for application on specialized response vessels are based mainly on brush skimmer technology
and separation of oil from ice in the water, thus avoiding the energy needed to lift oily ice blocks
from surface water for cleaning operations. Despite technological advances mchanical recovery
in ice-covered waters operate at lower efficiencies when sea ice is present, due to challenges
such as limited access to the oil, icing, freezing, and jamming of equipment. In ice conditions,
the use of brushdrum, brush adaptors, brush pack skimmers and rope mop concepts comprise
the best choices for dealing with oil spills. Disc, drum and rope mop skimmers can remove light
and medium viscosity oils. Brush and belt skimmers can collect heavy oils. The use of skimmers
in broken ice is considered more appropriate for small batch removals rather than to contain and
remove a large spill. The recovered fluid is expected to contain a significant amount of slurry
and slush ice, and therefore may take additional storage capacity.

While ice can act as an effective oil barrier, snow may be an effective oil sorbent. Equipment and
manual methods can be used to mix and remove large quantities of oiled snow, although this
mixture usually will have low oil content. Barriers made of snow can contain and prevent further
spreading of oil. As diesel or other light oils can penetrate snow, an impermeable boom or snow
barrier can be constructed by spraying accumulations of snow with water to form ice. Collected
oil can be pumped to storage, where another pump or skimmer can be used to remove it. Under
ice, oil tends to accumulate in naturally occurring surface depressions, so that trenches or holes
can be cut to gain access to the oil and pump it to storage. Rope mop skimmers can be deployed
under the ice, or the oil can be dealt with effectively when it surfaces in the spring. If conditions
are favorable, ISB can be attempted.

75

Pumping and vacuum systems have been used more often to remove deposits of non-floating oil;
however, these systems are more suited for thick deposits of liquid and semi-liquid oil close to
shorelines and in water with good visibility. Operation costs are usually higher and the method is
slow compared to conventional cleaning. Cold water temperatures increase oil viscosity and
reduce dive times, which increases cost and cleaning time. The effectiveness of pumps may be
decreased by the presence of sunken debris, medium-sized or larger rocks, rocky bathymetric
irregularities, or silt and fine sediments in the recovery area.

Under ideal conditions, physical recovery of spilled oil can be expected to achieve a maximum
recovery rate of 30% (NRC, 1989), and this could be much lower in cold water and Arctic
environments (Engelhardt, 1994). While improvements are being made on the design of
equipment for the recovery of oil under Arctic conditions, emphasis is now being placed on in
situ methodologies that require fewer response personnel and do not require the storage,
transport and disposal of waste.

5.3 In situ Burning

In situ burning (ISB) seeks to destroy oil through the combustion of ignitable components. This
technique is subject to limitations such as the need for specialized oil booms, concerns over the
generation of smoke and potential toxic by-products, operational constraints linked to sea-state,
and the combustibility of the oil. Based on the limited information published to date, ISB has not
been considered to be a first choice response strategy such as containment and mechanical
recovery, due to potential ecological and human health and safety concerns. However, its
reported success in the removal of approximately 5% of bulk surface oil following the DWH spill
has dramatically increased interest in its potential use (Federal Interagency Solutions Group,
2010). At present, research emphasis for ISB is on attaining higher efficiencies in emulsion
burning, refinement of towing manoeuvres, burn enhancement and extinguishing techniques,
burn equipment evaluations, smoke plume modelling, assessments of potential environmental
effects, training of ISB personnel, and research to extend the window of opportunity through the
use of chemical emulsion breakers.

76

Oil spills on water can be burned if the oil layer is thick enough and an adequate ignition system
is available (Fingas et al., 1995). Burning oil within transition seasons appears to be a practical
response option. In drift ice, oil spills can rapidly spread to become too thin to ignite. Fire booms
are required to collect the slick and keep it thick in open water. Burning oil on solid ice or on
snow is usually feasible within 2 - 5 days following a spill on ice or within several months if the
oil has spilled in or under ice.

Burning oil in a broken ice field is dependent on how well the oil is contained by ice. Brash,
frazil or slush ice will accumulate with the oil against the edges of larger floes and control the
thickness of oil in a given area (Potter and Buist, 2008). Burning efficiency can be about 50% in
calm conditions and 33% with waves (Buist et al., 2003). Cold-water herders can also be used to
thicken a thin oil slick in the presence of brash or slush ice. However, the presence of frazil ice
restricts the spreading of oil and reduces herder effectiveness. Short, choppy waves may cause a
herded slick to break up. Hydrophilic emulsion breakers are most effective when used in a
confined environment.

The use of herding surfactants may also facilitate the containment and burning efficiency of a
slick. Herding agents are usually applied in the periphery of an oil slick to change the balance of
interfacial forces acting on the slick edge and allow the interfacial tensions to contract the oil into
thicker layers. New herding agents are of low toxicity and the main surface-active ingredients of
successful herders are not soluble in water and are not intended to enter the water column, only
to float on the surface (Buist et al., 2008). A series of mid-scale testing was conducted at
different locations to examine the concept of using herding agents to thicken oil slicks among
loose pack ice for the purpose of ISB (Buist et al., 2008; Buist et al., 2006). Field experiments
were performed off Svalbard in May, 2008 to test the ability of a non-proprietary hydrocarbon-
based cold water herding agent (USN) to burn oil in drift ice (Buist et al., 2011). Experiments
were also carried out in a laboratory and large-scale refrigerated test tank to determine if a new
generation of fluorosurfactants or silicon-based surfactants are more effective than USN. In these
experiments, which included field trials, a larger number of tests were done in open water
conditions and the data indicated that the herders also worked efficiently in quiescent open water
(Buist and Nedwed., 2011). These results suggested that potential exists for the use of ISB
77

herding agents as an effective and rapid response option in open Arctic waters and remote
temperate regions.

Research has been directed to the study of potential environmental effects caused by ISB. It has
been demonstrated that the temperature at the slick-water interface is never more than the boiling
point of the water and usually remains close to the ambient temperature (Buist, 2003). There is a
steep temperature gradient across the thickness of the slick. The slick surface is very hot (350
500C) but the oil just beneath it is near ambient temperature. Fingas et al. (2005) analyzed the
soot resulting from the burning of heavy oils and an emulsion of bitumen. The residues are
mainly resins and asphaltenes not burned in the fire and the enriched pyrogenic PAH derived
from the petrogenic PAH of the oil. Resultant PAH concentrations correlated inversely with
burning efficiency. Thus, attaining higher burning efficiencies reduces the amount of pyrogenic
PAH in the soot.

The fate and behaviour of the burn residue has also been investigated (Gunette et al., 1995).
Experiments conducted in a circulating flume to investigate the ignition, combustion and residue
remaining from continuous burning of weathered emulsified Statfjord crude showed that most of
the burn residue remained buoyant and the lost residue was usually found submerged about 30
cm below the water surface. Further experiments with eight different oil types revealed that the
quantity of the burn residue that remained buoyant and the portion of it that sank were dependent
on oil type (Buist et al., 1995). While residues of thicker, heavy oil slicks and weathered crude
may sink in freshwater and saltwater upon cooling to ambient temperature, burn residues of
lighter oils may not sink.

Predicting ignitability of oil spills as a function of oil type and weathering was identified as a
need to make ISB more operational (Brandvik and Faksness, 2009). SINTEF researchers
constructed a small-scale laboratory burning cell to perform controlled experimental burns of
weathered and emulsified oil on a 100 mL scale (Brandvik et al., 2010c). The laboratory cell has
been verified against large-scale field ISB experiments and the results of the limits of ignitability
from both experiments correlated well. The burning cell has been used to generate a
comprehensive dataset with ignitability and burning efficiency as a function of oil type, ice
78

conditions and degree of weathering (Brandvik et al., 2010b). These data are used to develop and
calibrate an algorithm that is implemented in the SINTEF Oil Weathering model, and to predict
ignitability for oil spills. It is expected that this will enable more reliable predictions to be made
within the operational window of opportunity for use of ISB as a function of weathering.

Ignition promoters and emulsion breaking chemicals can also be used to increase the ignitability
of a slick. Studies have been conducted to develop a more complete understanding of the use of
emulsion breakers injected into an oil spill recovery system at both the lab-scale, and the large-
scale such as at the Ohmsett wave tank (SL Ross Environmental Research Ltd., 2002).

5.4 Chemical Dispersion

In cold temperatures, as the oil becomes more viscous, it becomes more difficult to disperse. On
the other hand, dispersed oil droplets will re-coalesce at a slower rate at low temperatures, and
leaching of the surfactant from oil would also be slower. These factors favour enhanced chemical
dispersion. Different oil types have different potential for chemical dispersion in cold waters.
While heavy oils have been considered out of range for effective dispersant use, the latest
generation of improved dispersants appears to be effective with oils having a viscosity
approaching 10,000 cSt (Colcomb et al., 2005; Fiocco et al., 1999).

Dispersion of fresh and weathered Alaska North Slope (ANS) crude oil in nearly freezing water
has been demonstrated (Belore et al., 2009). Corexit 9500 and Corexit 9527 were both effective
in dispersing fresh and weathered Alaskan oils in large-scale wave tank tests completed at the
Ohmsett facility in the winter, with 85-99% dispersant effectiveness (DE) and 50-90% droplet
size corrected DE. There was no apparent difference in DE on fresh and weathered oil within the
relatively narrow range of viscosities tested. These results demonstrate that chemical dispersants
can be effective on oils in near-freezing water and under energetic, breaking wave conditions. In
the laboratory, it has been reported that tests run with three dispersants at 0 and 20C showed
slight differences in volume mean diameter of dispersed LAGO medium crude oil droplets,
depending on dispersant composition and energy input (Byford et al., 1984).

79

The limiting oil viscosity for chemical dispersion of oil spills under simulated sea conditions in
the large outdoor wave tank at Ohmsett, New Jersey, has been determined (Trudel et al., 2010).
DE tests were completed using crude oils with dynamic viscosities ranging from 67 to 40,100 cP
at test temperatures of 15 and 27C. Tests produced an effectiveness-viscosity curve with three
phases when oil was treated with Corexit 9500 at a dispersant to oil ratio of 1:20. The dynamic
oil viscosity that limited chemical dispersion under simulated at-sea conditions was in the range
of 18,690 to 33,400 cP. Visual observations and measurements of oil concentrations and droplet
size distributions in the water under treated and control slicks correlated well with direct
measurements of effectiveness.

Heavy fuel oils, characterized by having a high proportion of polar fractions and diminished
evaporative fractions, are usually more viscous than unrefined crude oils, and tend to be more
difficult to disperse. Therefore, field responses to heavy fuel oil spills are more challenging than
the treatment of light and median crude oil spills.

The influence of temperature and wave conditions on the effectiveness of two commercial
dispersants with a heavy fuel oil were studied using a flow-through experimental wave tank (Li
et al., 2010). This study demonstrated that there was a strong effect of seawater temperature on
the DE of the IFO180 using Corexit 9500 and SPC 1000 within a narrow window from 10 to
17C. The effect of mixing energy from breaking wave conditions agrees with results from field
trials (Colcomb et al., 2005) and large-scale tank testing (Trudel et al., 2005) of DE for heavy
fuel oil. The observed temperature and wave effects were also consistent with laboratory baffled
flask tests of dispersed heavy fuel oil (Srinivasan et al., 2007).

In the presence of ice, an obvious limiting factor is the relatively calm sea state that may not be
able to deliver sufficient mixing energy for the dispersant to take effect in a cold environment. A
possible solution is to provide mixing energy using propeller wash from an ice-breaker. This has
been evaluated by a number of field programs (Nedwed et al., 2007; Spring et al., 2006) to
demonstrate the potential effectiveness of the technique to enhance dispersion efficiency in ice-
infested waters. A more recent field trial has demonstrated the effectiveness of physical
80

dispersion of oil in ice infested waters, by combining the energy input from ice-breaker propeller
wash with the addition of mineral fines (Lee et al., 2011c).

A new silicone-based agent has been engineered to spread viscous oil to an extremely thin sheen
(Nedwed et al., 2011b). A stable water-in-oil emulsion could be treated to spread to a thickness
that was visually undetectable. This new response approach could reduce the impacts of offshore
marine oil spills by promoting rapid spreading of the oil, which increases evaporation rates,
breaks emulsions, and potentially enhances natural dispersion such that the impact and toxicity
of a spill is reduced. It would be useful under Arctic conditions if it proves to operate at
temperatures below 0C.

A new gel dispersant has been tested that has the consistency of warm honey, contains more than
85% active ingredient, and is positively buoyant, cohesive, persistent and oleophilic (Nedwed et
al., 2011a). This new dispersant specifically designed for use in Arctic temperatures, and on
viscous oils, expands the window of opportunity for successful response operations.

From an ecological perspective of the hydrocarbon biodegradation consortium, chemical
dispersants may complement the naturally produced bio-surfactant of primary oil-degrading
microorganisms, with both sharing the same function of dispersing oil into small droplets that
offer a larger surface area for microbial attack. However, uncertainties exist in terms of DE of
different oil types at low temperature and ice-infested waters as well as the persistence and
biological effects of the dispersed oil. Development and performance evaluation of new
dispersants for use in cold waters is required. Understanding the behaviour and environmental
impact of chemically dispersed oil in Arctic waters are essential. Research is also needed to
understand the long-term fate of chemically dispersed oil, including long-term effects of
biodegradation and photo-oxidation. The research requires full-scale field trials that include
tracking and monitoring techniques. The development of operational guidelines requires
investigations with various types of oil and oil dispersant formulations. In addition, a framework
for comparative net environment benefit analysis is needed to assess the different oil spill
response and treatment options.

81

5.5 Oil-Mineral Aggregates

Dispersed oil droplets that are suspended in the water column may interact with suspended
particulate material (SPM), mineral fines in particular, to form oil-mineral aggregates (OMA).
Studies of OMA formation have demonstrated that both mineral fines and organic particles can
stabilize oil droplets within the water column under a variety of environmental conditions (Bragg
and Yang, 1995; Delvigne et al., 1987; Lee, 2002; Lee and Stoffyn-Egli, 2001; Lee et al., 2003;
Lee et al., 1996; Muschenheim and Lee, 2002; Omotoso et al., 2002). Various types of
aggregates can be formed depending on the physico-chemical properties of the particles, the type
of oil, and environmental conditions (Lee et al., 1998; Muschenheim and Lee, 2002; Stoffyn-Egli
and Lee, 2002). OMA formation has been observed at numerous field sites from the rivers of
Bolivia (Lee and Stoffyn-Egli, 2001; Lee et al., 2002) to the shores of Svalbard Island in the high
Arctic (Owens et al., 2003). The occurrence of OMA covers the range of natural variance for
temperature, salinity, oil types and mineral composition. This is a natural process that has been
observed in cold water and Arctic conditions as, for example, clay-oil flocculation in nearshore
waters in Alaska, and the interaction of glacial till with oil in the Beaufort Sea. Controlled
laboratory experiments (Cloutier et al., 2002; Khelifa et al., 2005b; Lee et al., 1997; Omotoso et
al., 2002; Stoffyn-Egli and Lee, 2002) and shoreline field trials (Lee et al., 1997; Lunel et al.,
1997; Owens et al., 1995; Owens and Lee, 2003) have both demonstrated that OMA enhance the
natural dispersion of oil spilled in the environment and reduce its environmental persistence.
Previous research has demonstrated that the formation of OMA has significant potential to
enhance oil dispersion and dilution.

In the natural environment the oil and SPM can interact to form aggregates. The formation of
aggregates between mineral fines, like clays, and oil has been studied in the past, and
enhancement of the process has been used as an operational shoreline cleaning strategy under the
name of surf-washing. When linked with SPM, the dispersed plume is stabilized and the
tendency for oil to coalesce at the surface is reduced because oil behaviour follows SPM
behaviour.

82

Numerical models support the hypothesis that OMA can form rapidly (Hill et al., 2002; Khelifa
et al., 2003a) as long as sufficient mixing-energy is available. Detailed chemical analysis of
samples recovered from coastal waters following surf-washing operations after the Sea Empress
spill in the United Kingdom conclusively demonstrated that OMA formation enhanced the
biodegradation rates of the residual oil (Colcomb et al., 1997; Lee et al., 1997) because the
stabilization of oil droplets by mineral fines increased the oil-water interface where microbial
biodegradation primarily occurs, and operates optimally. Thus, this remediation process not only
dilutes oil spilled into the environment, but it may also effectively eliminate many components
of environmental concern. In terms of protection of fisheries and fish habitat, OMA formation
and dispersion will minimize environmental impacts. For example, field studies in Svalbard,
Norway, demonstrated that OMA within the immediate vicinity of the spill site were dispersed to
levels below regulatory toxicity threshold limits (Owens et al., 2003).

Recent studies have specifically focused on the potential application of mineral fines to facilitate
OMA production as an operational marine oil spill countermeasure. Laboratory experiments
have delineated the effects of factors such as mineral type, physico-chemical properties of oil
and mineral fines, and the co-presence of chemical dispersants on the morphology and vertical
transport of OMA (Lee et al., 2009b; Zhang et al., 2009). Wave tank studies have shown that
chemical dispersants and mineral fines have a synergistic effect on the dispersion of crude oil (Li
et al., 2007). These recent wave tank studies have repeatedly demonstrated the synergistic effects
of mineral fines and chemical dispersants on dispersion of different oil types under both breaking
wave and regular wave conditions (Lee et al., 2009a; Lee et al., 2008; Lee et al., 2009b; Li et al.,
2009a; Li et al., 2009b; Li et al., 2009c).

An oil spill response technique in ice-infested waters based on the application a slurry of mineral
fines mixed by propeller-wash to form OMA has been proposed (Lee et al., 2011c). The process
promotes the physical dispersion of stabilized oil droplets into the water column which support
higher rates of oil degradation by natural bacterial consortia. To validate the operational
effectiveness of this technique a controlled oil spill experiment was conducted from a Canadian
Coast Guard ice-breaker in the St. Lawrence Estuary offshore of Matane, Quebec, Canada.
Time-series changes in oil concentrations were monitored to quantify dispersion effectiveness.
83

Field samples were also recovered for laboratory microcosm studies on the biodegradation of
petroleum hydrocarbons by monitoring CO
2
production and the depletion of specific
hydrocarbon components. Detailed chemical analysis by GC/MS with hopane normalization
showed that more than 60% of the total petroleum hydrocarbons, 75-88% of total alkanes, and
55-65% total polycyclic aromatic hydrocarbons (TPAH), were degraded after 56 days of
incubation at 0.5C. The alkylated PAH was degraded to a greater extent following the addition
of mineral fines. This technique offers several operational advantages as a spill countermeasure
for use under Arctic conditions such as reduced numbers of personnel required for its
application, no need for waste disposal sites, and cost effectiveness.

It has been reported previously by researchers at Cedre, France, that the addition of a dispersant
to water containing SPM enhances OMA formation (Guyomarch et al., 2002). This is attributed
to the fact that OMA formation usually follows the generation of micron-scale oil droplets which
then interact with mineral fines. Furthermore, ongoing preliminary research at Cedre to assess
the impact of chemically and mechanically dispersed oil (Discobiol program) has demonstrated
that the presence of SPM in the form of silt reduces the toxicity of the dispersed oil, since
aggregate formation may bind the oil in a matrix and reduce its bioavailability. In addition,
environmental risk assessment mathematical models have been useful for simulating the
transport of oil as a result of OMA formation in aquatic environments and predicting the
potential risks to the pelagic and benthic ecosystems (Lee et al., 2009c).

A combined application of chemical dispersant and mineral fines to combat oil in low
temperatures and low energy environments in the Arctic has great potential to be an effective
cold water in situ oil spill response. In light of the fact that hydrocarbon biodegradation in the
environment is often limited by nutrients such as nitrogen and phosphorus (Atlas and Bartha,
1972), the use of nutrient-enriched mineral fine particles, such as natural phosphorus-rich
minerals, should also be considered. As an integrated bioremediation process, the effects of
chemical dispersants and nutrient-enriched mineral fines on biodegradation of dispersed oil in
seawater should be evaluated. Although very unlikely to cause significant sinking of oil because
of current flow in the field, the possibility exists for increased sedimentation of oil to marine
sediments by the use of dense mineral fine particles or the interaction of dispersed oil with
84

natural SPM. The effects of chemical dispersants and nutrient-enriched mineral fines on anoxic
biodegradation of hydrocarbons, including alkanes and aromatics, coupled with terminal electron
accepting processes, such as sulphate, nitrate, and iron reduction, would also be of interest.

5.6 Bioremediation

Bioremediation aims to accelerate the natural biodegradation process of petroleum hydrocarbons.
Natural selection of oil degrading microorganisms in polar areas results from the natural seeps of
petroleum (Landes, 1973). Hydrocarbons are ubiquitous in marine systems, originating from
natural seeps of oil and natural gas deposits, marine oil transportation accidents and operational
discharges (Yakimov et al., 2007). Globally hundreds of millions of litres of hydrocarbons enter
the environment from both natural and anthropogenic sources every year.

Marine hydrocarbon-degrading bacteria were identified almost a century ago. Aerobic bacteria
that degrade hydrocarbons are increasingly recognized as key players in the removal of
hydrocarbons from oil-polluted marine environments (Head et al., 2006). Anaerobic
biodegradation of several classes of petroleum hydrocarbons, including alkanes and mono and
polycyclic aromatic compounds have been demonstrated in recent studies (So et al., 2003;
Townsend et al., 2003; Widdel and Rabus, 2001). In the light of these findings, future research
and development in Arctic oil spill bioremediation should also give some consideration towards
in situ methods that employ the biological degradation of hydrocarbon contaminants with
minimal mechanical containment and recovery of oily wastes, their transportation, or disposal.

Numerous laboratory and field studies at subzero temperatures have demonstrated the significant
biodegradation potential of petroleum hydrocarbons in the Arctic (Bragg et al., 1994; Brakstad et
al., 2008; Lee et al., 2009d; Prince et al., 2003a; Sveum and Ladousse, 1989). Biodegradation of
many components of petroleum hydrocarbons by indigenous, cold-adapted microorganisms at
low temperatures in Arctic, alpine, and Antarctic soils has been reported (Whyte et al., 2001).
Much work has been focused on the bioremediation of contaminated soils in the Canadian high
Arctic, where fuel used for transportation, electricity and heating has frequently spilled or leaked
from tanks and pipelines (Mohn and Stewart, 2000; Whyte et al., 2001; Whyte et al., 2002;
85

Yergeau et al., 2009). Because the remoteness and unique character of these sites preclude
conventional physico-chemical technologies for soil treatment, bioremediation is often the only
feasible cleanup option.

Marine waters including those of the Arctic usually have low concentrations of nitrogen and
phosphorus that limit the biodegradation of hydrocarbons (Prince, 1993). Addition of fertilizers
has been reported to stimulate the extent of biodegradation in the field (Bragg et al., 1994;
Prince, 1993; Prince et al., 2003a; Swannell et al., 1996; Venosa et al., 2002). In the case of the
Exxon Valdez oil spill, fertilizer amendments increased the rates of biodegradation by three to
five times (Atlas, 1995).

Different nutrient products have shown variable effectiveness, depending on their nature, the
properties of the oil, and the characteristics of the contaminated environments. Application
protocols for bioremediation (form and type of fertilizer or type and frequency of application)
must be specifically tailored to account for differences in environmental parameters, including
oil characteristics, at each contaminated site (Boufadel et al., 1999; Venosa et al., 1996; Wrenn et
al., 2006; Wrenn et al., 1994; Wrenn et al., 1997a; Wrenn et al., 1997b).

Bioremediation efforts have been carried out mainly on shorelines and in wetlands, rather than
offshore, because effective treatment of an oil spill is likely to take longer than the period for the
slick to reach the coast (NRT Science & Technology Committee, 2000), and because of the
difficulty of maintaining elevated nutrient concentrations in an open system subject to rapid
dilution (Leahy and Colwell, 1990). Delille et al. (1998) investigated the potential use of an
oleophilic nutrient formulation in Antarctic coastal seawater. An order of magnitude increase in
bacterial microflora occurred in seawater after it was contaminated with Arabian light crude oil.
The addition of the oleophilic fertilizer, INIPOL EAP 22, designed to maintain elevated nutrient
levels around the oil, clearly enhanced the growth of both saprophytic and hydrocarbon-utilizing
microflora. Chemical analysis of the residual hydrocarbon fractions confirmed that fertilizer
application increased the rate of oil biodegradation.

86

Significant advances have been made in the development of molecular techniques for detecting
microbial pollutant degrading genes in environmental samples without the need for cultivation of
the bacteria themselves. Catabolic genotypes involved in the degradation of representative
fractions of petroleum hydrocarbons, including n-alkanes, aromatics, and PAHs, are reported to
be widespread in Arctic soils and Alaskan sediments (Margesin et al., 2003). Applying
biotechnological methods in bioremediation studies can determine the prevalence and
composition of specific hydrocarbon degrading populations, and thus aid in assessing the
biotreatability of contaminants, and in monitoring the effects on specific population dynamics
during the remediation process. A recent study on prevalence of alkane mono-oxygenase genes
in Arctic and Antarctic hydrocarbon contaminated and pristine soils indicated that Rhodococcus
spp. may be the predominant alkane-degrading bacteria of the tested genotypes in polar soils.
Pseudomonas populations may become enriched following hydrocarbon contamination, whereas
Acinetobacter spp. are not a predominant member of the polar alkane degradative consortia
(Whyte et al., 2002). The ability to determine the alkane degradative composition of a microbial
community also helps to develop appropriate bioremediation strategies for a particular
contaminated site. For example, the presence of large numbers of alkane degradative
Rhodococcus spp., known to produce at the cell-surface biosurfactants with activity at cold
temperatures, and to directly adhere to solid alkanes at low temperatures (Whyte et al., 1999),
would eliminate the need to incorporate a surfactant treatment as part of an Arctic
bioremediation strategy.

Bioaugmentation consists of the addition of known oil degraders that are not native to the
contaminated sites to supplement the existing microbial population (Bartha, 1986).
Bioaugmentation approaches have been proposed in situations where the indigenous microbial
population may not be capable of degrading the wide range of potential substrates present in
complex mixtures such as petroleum, where they may be in a stressed state as a result of the
recent exposure to the spill, where the indigenous hydrocarbon-degrading population is too small
to maintain a high enough decontamination rate, or where seeding is required to reduce the lag
period in growth prior to the onset of the bioremediation process (Forsyth et al., 1995). However,
bioremediation field trials in open water environments have consistently shown that
87

bioaugmentation provides little or no benefit to treatment of the spilled oil (Lee et al., 2005b;
Nichols and Venosa, 2008).

In contrast to terrestrial hydrocarbon degraders which tend to be metabolically versatile and
utilize a large range of organic substrates, their marine counterparts are highly specialized
obligate hydrocarbon utilizers, who play significant, global roles in the natural cleansing of oil
pollution in marine systems (Yakimov et al., 2007). The marine hydrocarbon-degrading
microorganisms have the capacity to efficiently degrade hydrocarbons (Harayama et al., 2004;
Head et al., 2006). The toxic components of dispersed oil are eventually removed through
biodegradation and other attenuation processes, such as irreversible absorption to suspended
sediments.

Traditional bioremediation studies focused on the microorganisms that directly degrade the
contaminants. However, these microorganisms only form part of an ecological system
consortium (Head et al., 2006), which involves many direct and indirect interactions with other
community members and the environment, and therefore oil degradation processes are
influenced by environmental variables such as nutrient availability or physico-chemical
parameters. Such interactions of oil-degrading microorganisms with the environment include
competition for limiting nutrients, predation by protozoa, lysis by phages and cooperative
interactions to increase contaminant degradation.

The common criterion for decommissioning a contaminated site is based on oil chemistry, and
therefore most previous studies for monitoring bioremediation of hydrocarbon contaminated
matrices (soils, sediments, and waters) focused on chemistry, including target hydrocarbon
components, water quality parameters, and metabolic activities of microbial processes, such as
respiration, enzyme activities, and microbial counts (Margesin et al., 2003; Whyte et al., 2001).
Insights into the ecological structures and functions and understanding of the dynamics of the
microbial communities involved in the oil biodegradation process will facilitate design of more
efficient bioremediation techniques (Whyte et al., 2002; Yergeau et al., 2009). The development
and optimization of in situ bioremediation strategies requires better knowledge of the impact that
hydrocarbon contaminants have on marine microbial communities, which are the most important
88

oil-degrading groups in the open seas, sediments and shorelines of the Arctic region (Greer et al.,
2010). Greater insight into metagenomics the study of microbial genome fragments recovered
from environmental samples in contrast to genomes that are isolated from clone cultures that
sustain the catabolic function, transformation and mineralization of petroleum hydrocarbons is
also necessary (George et al., 2010). Moreover, understanding the interaction of microbial
communities and hydrocarbon pollutants helps to assess the potential recovery of contaminated
areas and to increase the opportunities for successful design and implementation of
bioremediation treatment.

5.7 Natural Attenuation

Monitoring natural recovery of oil relewased into the environment should also be considered an
operational response option under some circumstances. Natural seepage of petroleum
hydrocarbons is known to occur in the Arctic (Grant et al., 1986) and this prior exposure through
chronic or single events influences the number of hydrocarbon-degrading microbes (Delille et
al., 1997; Delille and Delille, 2000; Langworthy et al., 1998). One study of oil in Antarctic ice
and the effects on the bacterial population showed increases in hydrocarbon-degrading bacterial
populations in response to contamination (Delille et al., 1997). More recently, considerable
changes in bacterial populations within Arctic sea ice following exposure to hydrocarbons was
demonstrated, whereby diversity decreased and there were clear indications of hydrocarbon
degradation (Brakstad et al., 2008). Additional studies are required to substantiate these
observations and gain a better understanding of the overall degradation activity and the microbial
populations involved in the degradation of hydrocarbons in ice. Studies are required to address
the ice-water interface, where considerable amounts of hydrocarbon would be expected to pool
and thus be exposed to the high densities of bacteria found near that interface.

Until recently little attention has been given to improve our understanding of oil in ice
interactions and the ecological significance of microbial degradation activity of the residual oil.
While the potential biodegradation of oil in Arctic waters and within ice has been reported to be
extremely slow based on laboratory culture studies, recent respirometry and genomic studies
suggest significant rates of hydrocarbon degradation. The brine channel walls within sea ice
89

provide a large surface area that could be colonized by bacteria and algae and used for protistan
attachment and grazing (Krembs et al., 2000). At 2C, it was estimated that between 6 and 41%
of the brine channel surface area could be covered by microorganisms which is considerably
higher than the total surface area of soil where <1% is covered by microorganisms. Recently the
presence of microbes in the brine pore spaces of sea ice was confirmed (Junge et al., 2001) along
with their interaction and activity on the surfaces (Junge et al., 2004). Microbes have clearly
adapted to this environment, for bacterial strains readily isolated from sea ice have been found to
be cold-adapted, halotolerant species (Brown and Bowman, 2001; Junge et al., 2002). In the
Gulf of Mexico, recent studies have also reported very short half-lives for the dispersed oil
within the plume observed in the low temperature water recovered from a depth of ~1300 m
(Hazen et al., 2010). In terms of Arctic oil spill response, there is a need to understand natural
degradation rates for oil associated with ice.
90

6. Biological Effects of Oil

6.1 Oil Toxicity in the Arctic

Arctic marine ecosystems exhibit a number of unique characteristics in terms of habitat, and
community structure and function including (Camus et al., 2008):
permanently or seasonally ice covered waters;
seasonal and spatial fluctuations in primary production characterized by intense inputs
followed by long periods of limited availability;
multiple sources of primary production (ice algae, phytoplankton, benthic microalgae,
bacterial colonization of organic detritus);
relatively simple food web structure and short food chains;
strong pelagic-benthic coupling;
prevalence of large predators;
long-lived and slow growing organisms;
high lipid content of organisms (energy storage strategy).

Marine oil spills may occur during any phase of oil extraction, storage and transportation (WWF,
2007). As described in Section 3.4 Factors Influencing Oil Fate (Weathering), the lower water
temperatures in the Arctic may reduce oil degradation rates, while extended periods of darkness
and light which influence the breakdown of photosensitive components of oil and the presence of
ice and harsh weather conditions may hamper cleanup operations. All of these factors affect
the concentration and persistence of oil and spill treating agents potentially used in response
operations. The concentration and exposure time of these compounds influences the subsequent
risk to Arctic biota.

Of greatest concern is the biological impact on species considered to be valued ecosystem
components (VECs) - elements of the ecosystem having scientific, social, cultural, economic,
historical, archaeological or aesthetic importance, determined on the basis of cultural ideals or
91

scientific concern (Forbes, 2011). The vulnerability of a VEC is considered in its relationship
with anthropogenic, biotic and environmental drivers (Table 5).

Table 5 Valued ecosystem components (VEC) of various Arctic regions at risk from oil spills (INAC, 2010;
Word and Perkins, 2011).
VEC Reason Barents Sea Beaufort Chukchi
Copepod, Calanus
glacialis
Significant Tier II food web prey of
invertebrates, fish, marine mammals and
seabirds

Polar cod,
Boreogadus saida
Keystone Tier III food web species, predator
of lower trophic levels, and prey of fish,
marine mammals and seabirds

Sculpin,
Myoxocephalus sp
Marine and estuarine sculpin, important
primary carnivore and Tier III food web prey of
fish, marine mammals and seabirds

Harp seal Commercial value, interaction with fish stocks
Ringed seal Important prey for polar bear and Arctic fox
(Vulpes lagopus)

Bearded seal Possible indicator of benthic contamination
(feeds on bivalves and benthic invertebrates

Right whale Endangered
Minke whale Commercial value
Humpback whale
(Megaptera
novaeangleae)
High international profile
Bowhead whale Endangered, cultural value
Beluga whale Near threatened
Polar bear Extreme sensitivity to petroleum intoxication,
high international profile, status: vulnerable

Capelin (Mallotus
villosus)
Commercial value, important prey of cod,
seals, whales

Scallop Commercial value, prey of bearded seal,
possible oil spill indicator species

Recreational seal
hunt
Cultural value
Fast-ice ecosystem A petroleum spill would disperse slowly

92



A number of comprehensive studies have been published describing biological components
within various Arctic systems. For example, Gulliksen and Lnne (1991) recorded sea ice faunal
assemblages with invertebrates and fish both in the Arctic and Antarctic. Ikvalko and Gerdes
(2004) gathered data from 91 different sources for the survey of the diversity of Arctic sea ice
biota. Arndt and Swadling (2006) provide information on distribution, diet and life history
strategies of crustaceans in Arctic and Antarctic sea ice. Iken et al. (2005) investigated the food
web structure of the Canada Basin in the high Arctic. Organisms of all habitats, sympagic (ice-
associated), pelagic and benthic were investigated. Rand and Logerwell (2011) performed a
demersal trawl survey of benthic fish and invertebrates in the Beaufort Sea. Darnis et al. (2008)
investigated sea ice and the onshore-offshore gradient in pre-winter zooplankton assemblages in
the southeastern Beaufort Sea. Hopcroft et al. (2005) sampled zooplankton communities of the
Canada Basin. Bluhm et al. (2005) investigated the macro-benthic and mega-benthic
communities in the Canada Basin. Despite this level of effort for basic knowledge, Pegau (2011)
suggests there is limited knowledge of the biological communities that might be at risk from oil
spills. Little is known about the effects of hydrocarbons on the biology of the Arctic compared
to temperate areas and less is known about the effects of oil in ice. Furthermore, this is
compounded by ongoing changes in the Arctic ecosystem. Along with global climate change
which is modifying ice coverage in the Arctic, the impact of ocean acidification in the Arctic
might alter the sensitivity of organisms such as molluscs, crustaceans, echinoderms, encrusting
algae, and certain types of marine phytoplankton to toxic compounds (NOAA, 2010).

6.2 Routes of Exposure

Little is known about the responses of organisms living between -1.88C and 5C (Camus et al.,
2002a; Camus et al., 2002b). The low temperatures of Arctic regions may alter the behaviour of
oil as oil dissolution is temperature dependent (Payne et al., 2001b), as discussed in Section 3.3
Factors Influencing Oil Behaviour. Marine organisms living at low temperatures, with the
seasonal presence of sea ice, strong seasonal changes and low food availability have developed
specific physiological adaptations such as high lipid content, low production/biomass ratio, anti-
93

freeze mechanisms and low metabolic rates (Weslawski et al., 1988). Four main routes through
which organisms can be exposed to oil or petroleum compounds resulting from a spill are:
direct contact, or when an organism becomes coated;
ingestion by eating or drinking;
inhalation of vapour, mist, or spray; and
absorption directly through its skin or respiratory membranes.

Oil affects aquatic biota by one of three means (Albers, 2002):
physically as by smothering animals, coating them or reducing light transmission;
by altering habitat characteristics such as changing pH, reducing dissolved oxygen or
reducing available food;
by acting as a toxicant.

Generally, PAHs are considered the source of toxic effects of oil (Albers, 2002). In water the
toxicity increases with the molecular weight of the aromatic to a limit of molecular weight 202.
Beyond this limit water solubility decreases as does toxicity (Albers, 2002). The toxic action of
PAHs is related to their ability to interfere with cellular membrane function, and PAH
metabolites may also bind to cellular proteins and DNA (Albers, 2002). Background
concentrations of PAHs in water are very low much below toxic thresholds. However, during
oil spills, concentrations may be higher than threshold values. PAH concentrations in sediment
are much higher than those observed in water but bioavailability is limited, thus reducing the
toxic potential (Albers, 2002).

Biota that can be affected occupy various habitats within an ecosystem (Boyd et al., 2001).
Surface dwellers such as birds, marine mammals and reptiles are subject to risk from exposure to
oil floating on the sea surface. Pelagic inhabitants of the water column, like fish and plankton,
are typically at lower risk although natural, chemical or enhanced dispersion of the oil can
temporarily increase their risk. Benthic creatures living on and in the sediments, such as crabs,
bivalves, and plants, are at risk from oil that sinks or is in contact with the sediment. Intertidal
biota like crabs, bivalves, anemones and algae, are at high risk of exposure if the spill reaches the
shore.
94


For the purpose of natural resource damage assessment, Logerwell and Baker (2011) consider
three habitats that could be impacted by spilled oil in the Arctic: neustonic, pelagic and benthic.
The neuston layer could be directly impacted by the release of surface oil. Oil released from the
seabed is expected to primarily impact both benthic and pelagic habitats. Surface oil that
becomes entrained could impact the pelagic habitat. If the entrained oil becomes associated with
suspended particulate material (Lee et al., 2011c), the residual oil may sink and impact benthic
organisms. Both surface slicks and entrained oil can also impact shorelines as reported during
the Baffin Island Oil Spill (BIOS) project.

Biota associated with the Arctic neuston that could be potentially impacted include ice algae,
phytoplankton, zooplankton and ice invertebrates, ichthyoplankton, pelagic fish and crustaceans,
seabirds, anadromous fish, and polar bears (Logerwell and Baker, 2011). Neustonic, pelagic and
benthic inhabitants that could potentially suffer impact are sea ducks, gray whales, walrus and
seals, while the benthic members that could be impacted are bivalves, polychaetes, macroalgae,
demersal fish, subtidal invertebrates, shorebirds, saltmarsh vegetation and caribou which feed on
grasses and sedges in salt marshes (Logerwell and Baker, 2011).

6.3 Bioaccumulation, Biomonitoring and Toxicity Assessment

Bioaccumulation of hydrocarbons is a concern with shellfish as it can cause tainting, but it is
generally not thought to affect fish or mammals because they are able to metabolize
hydrocarbons (Boyd et al., 2001). Contaminated shellfish can depurate (excrete) hydrocarbons
when moved to clean seawater. Finfish can metabolize oil within several days after exposure.

During a large-scale field experiment in the marginal ice zone of the Barents Sea, 7,000 L of
fresh Troll crude were released between the ice floes (Faksness et al., 2010). The oil was not
contained except by the floes. The water accommodated fraction (WAF) was of particular
interest because most of the toxicity from a slick is attributed to its semi-volatile organic
compounds consisting of naphthalenes, phenanthrenes, dibenzothiophenes and phenols. Water
samplers were deployed approximately 3 m below the ice for risk assessment analysis. The
95

concentration of total extractable hydrocarbons in the water from large volume samplers was 4-
32 ppb, and the concentration of water soluble components from semi-permeable membrane
devices was 0.6 (background) to 4 ppb. A hazard index, an estimate of the hazard of dissolved
oil in the water column from the spill, was calculated from the ratio of the estimated LC
50
of
individual components to the measured concentrations that organisms experience (Faksness et
al., 2010). Results of WAF data obtained from the large water samplers and semi-permeable
membrane devices gave an HI from 0.0004 to 0.11, indicating that the risk of acute toxicity was
very low. Conversely, the HI for the Corexit 9500 chemically enhanced water accommodated
fraction (CEWAF) ranged from about 1.4 to 4.7 indicating acute risk (Faksness et al., 2010).

Biomonitoring tools for use in risk assessment in the Arctic are being developed (Camus et al.,
2010). An environmental impact factor is being used in risk assessment to provide a quantitative
estimate of ecological risk for a region. The environmental impact factor is based on the ratio of
the predicted environmental concentration (PEC) to the predicted no effect concentration
(PNEC). The PNEC is obtained using LC
50
toxicity data (Table 6) from testing the species of
concern with the pollutant of interest, and an assessment factor may be added depending on the
quality of the empirical data. Usually, the most sensitive species (having the lowest LC
50
value)
is used. If PEC/PNEC < 1, no risk reduction is needed; whereas if PEC/PNEC 1, the risk is
unacceptable and measures must be implemented by the operator to reduce it (Camus et al.,
2010).

Table 6 Food web functional group and acute toxicity LC
50
(95% confidence interval) using 2-methyl
naphthalene for co-inhabiting Arctic species (Camus et al., 2010; Carroll et al., 2010).
Species Common Name Functional Group 96 h LC
50
(mg/L)
Acemaea tessulata Limpet grazer 0.35 (0.30-0.42)
Strongylocentrotus
droebachiensis
green sea urchin grazer 0.68 (0.62-0.74)
Boreogadus saida polar cod predator 0.80 (0.53-1.21)
Littorina littorea Periwinkle grazer 1.23 (0.83 1.83)
Gammarus sp. Amphipod scavenger 1.34 (1.08-1.67)
Pandalus borealis Shrimp opportunist 1.56 (1.31-1.86)
Sclerocragnon sp. Shrimp opportunist 1.70 (1.46-1.97)
Anonyx nugax Amphipod scavenger/predator 1.92 (1.48-2.48)
Chlamys islandica Arctic scallop filter feeder 2.75 (1.17-6.47)
Margarites helicina Snail grazer 5.01 (4.14-6.07)
Nymphon gracile sea spider predator/scavenger 5.42 (4.23-6.95)

96


Camus et al. (2010) posed the following questions.
Are available toxicity data for non-Arctic species representative of Arctic species?
Does the use of data for temperate species provide sufficient protection for Arctic
ecosystems?
What are the uptake rates, body burdens, and depuration rates for Arctic species?

Biomarkers are quantifiable responses in an organism that often indicate stress due to one or
more environmental stimuli. Biomarkers that can be used are based on various biological
activities such as gene expression, protein synthesis, antibody production, changes in blood
chemistry, frequency of a particular deformity, or some other morphological or biochemical
change. For example, one of the most widely used biomarkers is ethoxyresorufin-O-deethylase
(EROD) induction. The EROD assay monitors the induction of the cytochrome P4501A enzyme
system (CYP1A) which metabolizes xenobiotics, and is a widely used biomarker for exposure of
wildlife to substances that bind the aryl hydrocarbon receptor (AhR). EROD is induced in fish in
the presence of xenobiotics including PAHs (Hodson et al., 1991). For certain biomarkers, a
change in the biomarker is a toxic response. However, it is important to note that induction of (a
change in) the CYP1A system is not a toxic response, but an indication that the organism is
responding to a xenobiotic compound.

Some of the biomarkers that have been examined for polar cod include CYP1A, EROD,
glutathione S-transferase (GST), bile metabolites and DNA adducts by the comet assay
(Nahrgang et al., 2010a; Nahrgang et al., 2010b; Nahrgang et al., 2009; Nahrgang et al., 2010c;
Nahrgang et al., 2010d). The lysosomal membrane stability test was a suggested addition to the
battery of biomarkers that could be used in risk assessment and monitoring. Baseline levels were
low and did not mask the effect of contaminant induction; however, natural variation in baselines
for an organism can change seasonally according to such things as general health, biological
activity and maturity, and this is an area requiring further research (Camus et al., 2010). In
addition to this, the lipid content of Arctic organisms, and their metabolic kinetics determine the
uptake, accumulation and depuration of lipophilic toxins such as PAHs.

97

Recently, a more ecosystem-relevant approach has been to use a species sensitivity distribution
rather than the toxicity data for the single, most sensitive species (Carroll et al., 2010). The
Arctic and temperate species sensitivity distributions based on LC
50
values were found to be
comparable within the 95% confidence limits. It was concluded that the data obtained for
temperate species can be applied to Arctic species (Carroll et al., 2010). This conclusion is
similar to that for Atlantic and Pacific herring. The toxicity data for Atlantic herring embryo
response to dispersed oil can be used for Pacific herring (Lee et al., 2011b).

6.4 The Arctic Food Web

Figure 22 and Figure 23 are schematics of typical Arctic food webs. While appearing very
complex, Arctic food webs are considered to be simpler than those in temperate climates (Camus
et al, 2008). Camus et al. (2008) reviewed ecotoxicological studies relating to Arctic systems.
They concluded that
adult Arctic organisms are relatively resistant to oil compared to temperate organisms;
Arctic communities may be more vulnerable to oil pollution than temperate ones; and
early life stages of organisms appear to be highly sensitive to oil compounds.


Figure 22 The Arctic Food Web (ACIA, 2004).
98




Figure 23 Energy flow above trophic level 1 from phytoplankton and pelagic detritus in red, or benthic
detritus in blue, and proportional shades in between. Top panel: eastern Bering Sea shelf; bottom panel:
western Bering Sea shelf. Box and text size are proportional to log
10
of biomass for the compartment; area of
each link proportional to volume of flow (Aydin et al., 2002).

99

6.5 Effects on Arctic Sea Ice Communities

Arctic sea ice covers significant portions of the northern hemispheres oceans: approximately 7
10
6
km
2
in during the summer, and about twice as much area in the winter (ArcOD, 2010). Sea
ice occurs in various configurations such as thin or nilas pancake ice, ridged first-year and
multi-year ice, and pack ice. Multi-year ice found in deep basins can be up to 2-3 m thick.

Arctic sea ice forms at a freezing point of about -1.9C, since the salinity is usually about 33
parts per thousand (ppt). As ice crystals grow in the water during the autumn season, small ice
platelets begin to accumulate at the ocean surface, and inter-link to form a porous structure of ice
crystals filled with liquid referred to as brine. The ice-specific ecosystem at the ice-water
interface and brine channels (which range in diameter from microns to a few centimetres while
the temperature remains above -5C) includes bacteria, viruses, unicellular algae, diatom chains,
worms and crustaceans that form the base of the Arctic food web (Bowman and Deming, 2010;
Deming, 2010; Krembs and Deming, 2011).

Temperature controls the physical and chemical aspects of ice. As the temperature drops, the ice
becomes more solid with a corresponding reduction in porosity and smaller brine channels,
accompanied by an increase in salinity of the brine. Winter temperatures can range from -35C at
the surface to -2C at the bottom of the ice floe. Salinity can reach 250 ppt.

Sea ice also provides a critical habitat for photosynthetic algae and a nursery ground for
invertebrates and fish within the Arctic ecosystem. When the ice melts in spring and summer,
excretions of metabolic products and debris from dying cells that accumulated over the winter
are released into the water along with organisms (Michel et al., 2002). The excretion of
metabolic products and debris from dying cells can become incorporated into the cryopelagic
food web, or become deposited on the surface of the sediment to contribute to the benthic food
web. At the peak production of ice algae in spring, the solid ice transforms into pack ice with
individual floes that transport organisms, sediment and anthropogenic pollutants over thousands
of kilometres before they melt and discharge their contents into the water (Krembs and Deming,
2011).

100

Sympagic algae and phytoplankton are considered the most important primary producers in the
Arctic, capturing the energy of sunlight via photosynthesis. In the lower sections of ice, algal
blooms make up about 57% of the Arctic primary productivity (Gosselin et al., 1997). Epontic
algae such as the diatoms, Nitzschia grunowii and N. frigida (Cross, 1982b), grow on the
underside of the ice, beginning each spring as sunlight returns. Many ice-dwelling organisms
accumulate organic molecules including lipids that act as internal antifreeze. Organic compounds
exuded by the microbes have been shown to cause the ice to become sculpted (Krembs et al.,
2011; Raymond and Knight, 2003), facilitating the attachment and growth of algae that utilize
the nutrients locked in the brine channels. The algal diatom, Melosira arctica, can form filaments
several metres long extending into the water column. Zooplankton grazers such as Apherusa
glacialis, Gammarus wilkitzkii and Onisirnus glacialis feed on diatoms and are protected in the
sculpted ice channels. Protozoans and metazoans, especially turbellarians, nematodes,
crustaceans and rotifers, can be abundant in all ice types throughout the year (ArcOD, 2010). On
the underside of ice floes Calanoid copepods and amphipods (e.g. Parathemisto) that graze on
the ice algae are often observed at densities of up to several hundred individuals per square
metre. Polar cod (Boreogadus saida) feed on the amphipods, copepods and zooplankton, and are
in turn eaten by seabirds and marine mammals (Bradstreet and Cross, 1982). The polar cod, B.
saida, is often called Arctic cod, but there is another species, Arctogadus glacialis, which is also
called Arctic cod. A. glacialis is a deepwater gadid lacking barbels, that is found between 85
and 72 latitude north in the western Arctic basin, and along the northeastern and northwestern
coasts of Greenland. For the purpose of this report, B. saida will be referred to as polar cod.

Due to the reluctance of conducting large-scale in situ oil spill experiments in marine areas
where surface ice is prevalent, comprehensive knowledge of the effects that oil has on sympagic
biological communities is largely incomplete (Pegau, 2008), so in many cases, oil spill
researchers must infer oil-in-ice effects based on experiments which were conducted in ice-free
conditions. Pegau (2008) further states that ice environments are very different from typical
marine environments, and that caution must be used when extrapolating effects to the ice
environment.

101

Brackstad et al. (2008) noted that although concentrations of specific oil-degrading prokaryotes
seem to remain static, total counts of heterotrophic bacteria were higher in ice that was
contaminated with oil. Further, bacterial populations in contaminated ice were quickly
dominated by a few microbes, while clean ice communities were more heterogeneous.

The vast majority of the biological activity associated with sea ice is located along the bottom
boundary of the ice (the epontic community) and is primarily composed of microscopic
phototrophic algae, with yearly maximum algal concentrations being found during the spring.
Current evidence clearly suggests that direct contact between algal species and oil can lead to
death, but that certain species of diatoms seem to have a higher degree of resistance to oil
exposure (possibly due to their silica shell or frustule) than other plankton groups and tend to
recover most quickly following exposure (Pegau, 2008). Overall, the effects of oil on the primary
productivity of algal communities associated with ice seems to be negative, but at very low oil
concentrations, primary productivity may actually be increased it is thought that this might be
due to suppressed grazing, or increased supply of micronutrients from the oil (Cross, 1982a;
Cross and Martin, 1987). While large quantities of oil on the ice surface can effectively block
light from reaching the epontic primary producers, moderate amounts of oil tend not to affect the
light field any more than natural variations in snow depth and ice thickness (Cross and Martin,
1983).

Some meiofaunal communities at risk of getting stuck in oil, or ingesting it either directly or
through their prey, have not been shown to actively avoid oil, making them highly susceptible to
its effects. Amphipods have shown different mortality rates in the presence of crude oil
depending on species, life stage, and the nature of the oil to which they were exposed (Pegau,
2008). While copepods and polychaetes show some sensitivity to dispersed oil, little response is
observed when exposed to fresh or solidified oil, and nematode populations are largely
unaffected by the presence of oil in any form (Pegau, 2008).

Laboratory generated WAF from crude oil was used to expose gravid female G. wilkitzkii ice
amphipod gammarids (in embryonic developmental stages 3 to 9) for 30 days in a continuous
flow system (Camus and Olsen, 2008). The TPAH concentration decreased with time, from 55 to
102

8 ppm in the high dose, from 10 to 2 ppm in the medium dose, and from 5 to 1 ppm in the low
dose. The females used in the different treatments were in similar stages, had similar weight and
similar numbers of embryos, which were scored for aberrations. Aberrations were significantly
higher in the 55 ppm TPAH group. Many of these embryos had a wide gap between the
membrane and foetus, which might have been caused by PAH exposure. Aberrations in the 10
and 5 ppm treatments were not significantly different from the control. The study showed that
embryo development can be impaired by exposure to WAF having a TPAH concentration of 55
ppm.

Using the same continuous flow oil exposure system, sea ice amphipods were examined for
cellular energy allocation (Olsen et al., 2008). This biomarker is measured by assessment of the
protein, carbohydrate and lipid content of individuals, as well as their electron transport system,
to calculate an energy budget. Protein levels were significantly higher (p = 0.02) in the animals
subjected to the medium dose. This might indicate sublethal stress causing a shift in protein
metabolism; however, negative cellular energy allocation values indicated that more energy was
being consumed than stored, suggesting that the experimental animals (including the controls)
were under stress during the manipulation.

6.6 Effects of Oil on Arctic Biota

Bacteria

Studies on the effects of chemically dispersed oil and hydrocarbon biodegradation have been
conducted with natural phytoplankton and bacteria in mesocosms under cold temperature
conditions. Atlas et al. (1976) showed that bacterial populations increased under oil slicks
experimentally floated in Prudhoe Bay, Alaska, in large part due to oil degrading Pseudomonas
spp., but species diversity appeared to be unaffected. Siron et al. (1993) noted a marked increase
of both density and proportion of oil-degrading bacteria observed two days after the oil addition
that suggested the potential capability of the indigenous bacterial community to adapt to an oil
103

spill event. However, under such extremely icy conditions, the biodegradation of the dispersed
oil was reduced and only the aliphatic hydrocarbon fraction was slightly degraded.

Peterson and Dahllff (2007) studied the phototoxicity of pyrene to bacteria in Arctic sediment.
Their results indicated that shallow Arctic marine areas might be affected by phototoxicity if
concentrations of oil components in the sediments increase. The impact of crude oil on changes
in bacterial communities of Arctic sea ice has also been studied by Gerdes et al. (2005). They
reported a strong shift in community composition to a population dominated by -proteobacteria
in sea ice and melt pool samples incubated with crude oil. Similar research by Brakstad et al.
(2008) looking at in-ice bacterial community response to North Sea paraffinic oil showed that
over the 112 day experiment, oil-contaminated ice cores showed stimulated bacterial growth, but
decreased bacterial diversity which was evident from fewer bands in the denaturing gradient gel
electrophoresis, DGGE. In these contaminated cores, the microbial community became
predominated by Gammaproteobacteria related to the genera Colwellia, Marinomonas, and
Glaciecola, while populations in clean ice cores showed more heterogeneous populations.
Through additional chemical analysis used to assess n-C17/Pristane and
naphthalene/phenanthrene ratios in the ice cores, Brakstad et al. (2008) were also able to show
signs of slow oil biodegradation, primarily in the deeper parts of the ice where low hydrocarbon
concentrations were found.

Phytoplankton and Macroalgae

Pelagic organisms in the water column include unicellular phytoplankton species of diatoms,
dinoflagellates, and protists. When the water column does not support phytoplankton growth, the
ice algae and associated invertebrates become the main forage for polar cod, which are a primary
food source for seabirds and marine mammals.

Field and laboratory tests have shown that primary production (photosynthesis) is reduced in the
immediate area of an oil spill, but recruitment of phytoplankton from nearby, unaffected areas
happens quickly (Percy et al., 1985). In the presence of low concentrations of oil, some
phytoplankton species of the Beaufort sea (especially green algae) showed increased growth, and
104

at the Balaena Bay experimental oil spill site, diatom growth appeared to be enhanced (Percy et
al., 1985). The dispersant, Corexit, affected the plants in the same way as the oil did. Macroalgae
in the intertidal zone usually survive a single, short-term spill event, but die under chronic
exposure, while salt marsh plants are more fragile (Percy et al., 1985).

Reports of benthic macroalgae (seaweeds) in the sublittoral zone of the Arctic include three
species of Chlorophyta, 14 species of Phaeophyta, and 11 species of Rhodophyta (Hamel and
Mercier, 2005). Research by Stepaniyan (2008) shows that species of benthic macroalgae can
exhibit varying responses to the presence of oil. In these studies it was found that brown algae
such as Lamanaria saccharina and Fucus vesiculosus were largely unaffected by crude oil, while
other species such as Ascophyllum nodosum and red algae Porphyra umbilicalis and Palmaria
palmate exhibited a decrease in growth rate at oil concentrations of 5-10 mg/L. Photosynthetic
activity decreased with increases in the concentration of oil in all species of macroalgae studied
other than Ascophyllum nodosum, which experienced an increase in photosynthetic activity when
exposed to oil. It is believed that these differences in response could be related to structural
differences between these species of macroalgae, and in all cases, observed effects diminished
accordingly when oil exposure was reduced or removed.

Whereas the PAH, pyrene, was shown to increase nutrient uptake in bacteria, the findings of
Olsen et al. (2007) and Stepaniyan (2008), supported by Petersen and Dahllf (2007), found that
the presence of pyrene reduced the photosynthetic rate and nutrient uptake of the benthic algal
community.

Hsiao (1978) examined the growth of marine Arctic phytoplankton exposed to crude oils. Four
unialgal cultures (diatoms Chaetoceros septentrionalis, Navicula bahusiensis, Nitzschia
delicatissima, and green flagellate Chlamydomonas pulsatilla) were randomly selected from a
collection of southern Beaufort Sea isolates and maintained on incubator shakers at 200 rpm and
0, 5 and 10C under continuous light, in artificial ASP 2 seawater amended with 1.26 g/mL
Na
2
MoO
4
2H
2
O and 192 g/mL NaHCO
3
. Three northern crude oils (Atkinson Point, Norman
Wells and Pembina) and one southern crude (Venezuela) were added in volumes of 2, 20, 200
and 2000 L directly to 200 mL of 4.8 10
4
cells/mL of each algal suspension to make test
105

concentrations of 10, 100, 1000 and 10,000 ppm. No chemical dispersant was used and controls
contained no oil. Algal growth was assessed every two days by cell enumeration on a Neubauer
haemocytometer. Exponential growth rate and generation time were determined. Survival was
determined on the basis of pigmentation and motility after ten days of exposure to 10 ppm of the
four crude oils at 10C.

Survival of the green flagellate, C. pulsatilla, within the first two days of exposure was 44-68%,
which increased with increased exposure time. Growth rate and generation time were not
significantly different among the various oils at 10 ppm, although growth was slightly stimulated
by Norman Wells and Pembina crudes at 0C. Survival of C. septentrionalis after two days
ranged from about 50-90% and decreased continuously thereafter. Survival also decreased
sharply with increasing concentration of all oils and growth inhibition increased with increasing
exposure time. Survival of N. bahusiensis and N. delicatissima was significantly reduced with
increased exposure time, and growth inhibition increased with increased time, temperature and
oil concentration.

In general, lethality of the crude oils to the phytoplankton varied with species, oil, temperature
and length of exposure. The diatoms (C. septentrionalis, N. bahusiensis, N. delicatissima) were
inhibited by a 10 ppm concentration of oil after ten days of exposure at 10C. Longer exposure at
100 ppm, 5 and 10C, caused greater inhibition than at 0C, the diatoms suffering the most. N.
delicatissima was the most sensitive, followed by N. bahusiensis and C. septentrionalis. C.
pulsatilla (green alga) was the most tolerant. It was able to recover and grow after exposure
(unlike the diatoms), with no mortality observed under any of the test conditions. Arctic waters
are dominated by diatoms and flagellates, and an oil spill in a habitat occupied by these algal
species could result in an overabundance of the green alga (Hsiao, 1978).

The same four crude oils were used to test the effect of the water accommodated fraction (WAF)
and chemically enhanced water accommodated fraction (CEWAF, using Corexit dispersant) on
natural populations of phytoplankton from the Beaufort Sea and Eskimo Lakes, and on the
macrophytic sporophyte seaweeds, Laminaria saccharina and Phyllophora truncata (Hsiao et
al., 1978). Results with showed that pimary production by phytoplankton was variously
106

stimulated or inhibited for Beaufort Sea samples at 10 ppm crude. Atkinson Point crude caused
about 10% inhibition. The other three WAFs resulted in mean stimulation: Norman Wells, 2%;
Pembina, 18%; Venezuela, 8%. For the WAF/CEWAF comparison using phytoplankton
assemblages from the Eskimo Lakes, CEWAF caused more inhibition and less stimulation than
the WAF, except for Pembina, which showed a mean stimulation of about 8% as CEWAF
compared with 10% inhibition as WAF. The authors suggested that the general increase in
toxicity of the CEWAF might have been due to the dispersant making the lighter ends and
aromatic compounds more available for uptake. Corexit alone caused an average stimulation in
photosynthesis of about 18%.

Photosynthesis was consistently inhibited in both macrophytes by all the oils, especially
Venezuela crude. Laminaria was most sensitive. The authors note that these results were
consistent with those of Hood et al. (1973) who found that 7 ppm of Prudhoe Bay crude inhibited
photosynthesis in L. saccharina, Cladophora stimpsonii and Ulva fenestrata. It is suggested that
inhibition might be a result of impaired cellular permeability, chlorophyll destruction,
chloroplast membrane disruption, accumulation of end products, and blocking of gas exchange
(Hsiao et al., 1978).

Salt Marsh Vegetation

Salt marsh plant species of the Arctic that could be impacted by oil spills include the sedges
Carex glareosa, C. subspathacea and Eleocharis acicularis; grasses Festuca rubra,
Calamagrostis deschampsioides, Puccinella phryganoides; dicotyledons Primula stricta,
Stellaria longipes, Parnassia palustris, Chrysanthemum arcticum, Ranunculus cymbalaria;
submerged macrophytes Myriophyllum exalbescens and Potamogeton filiformis; mares tail
(Hippuris tetraphylla); and low willow shrubs (Logerwell and Baker, 2011).

At Balaena Bay, the experimental oil spill significantly reduced the photosynthetic activity of
salt marsh macrophytes and algae by up to 90%, and researchers felt that coastal salt marsh
areas, which are important feeding grounds for birds, and nurseries to fish and invertebrates,
107

would be more likely to suffer permanent damage from a spill than a rocky coast (Percy et al.,
1985).

Slow recovery of vegetation in colder climates has been attributed to the low temperature (which
also reduces the length of the growing season), the high organic peaty soil, sheltered location,
and the type of oil spilled which is often fuel oils (NRC, 2003).

Zooplankton

In the Chukchi Sea, between Siberia and Alaska, copepods comprise the most abundant group of
Arctic zooplankton along with krill (euphausids) and the meroplanktonic larvae of barnacles and
bivalves (ArcOD, 2010; Logerwell and Baker, 2011). There are also the more fragile gelatinous
forms which are often underestimated by present sampling techniques, and whose trophic
importance is not well known (ArcOD, 2010). Other types of zooplankton include jellyfish,
larval finfish (ichthyoplankton comprising about 23 taxa in the Chukchi, dominated by polar cod,
Bering flounder and yellowfin sole), and the larvae of many bottom-dwelling invertebrates
(Logerwell and Baker, 2011). The zooplankters of the Beaufort Sea are similar to those of the
Chukchi Sea, with larvaceans, chaetognaths (tube-like carnivores with a transparent cuticle about
2 to 120 mm), copepods and barnacle larvae in abundance, and polar cod dominating the
ichthyoplankton (Logerwell and Baker, 2011).

Relatively little effort has been directed towards studying effects of acute oil spills on
zooplankton. A reduction in zooplankton concentrations in a critical area and time period could
have significant effects on a large, economically important fish stock. The planktonic life stages
(eggs, larval and post-larvae) of cod cohorts are considered the most susceptible to oil spill
effects as they have low mobility, restricted horizontal distribution and shallow vertical
distribution (Stige et al., 2011). The effects of oil on Calanus finmarchicus, a temperate-boreal
copepod found in sub-Arctic waters, and Calanus glacialis, a true cold water Arctic species,
were compared. The two species are adapted to 10C and 2C, respectively. They were
compared on the basis of acute lethality (LC
50
), lipid content and the WAF-mediated induction
of the gene encoding glutathione S-transferase (GST). LC
50
values

differed between the two
108

species and the Arctic copepod appeared less sensitive than the temperate-boreal species. Lipid-
rich copepods survived longer than lipid-poor copepods at the same exposure concentrations.
Both species showed trends in GST expression that were dependent on concentration and
exposure time. The Arctic copepod appeared to respond more slowly and with a lower intensity.
The differences were attributed to the colder temperature rather than inherent differences
between the Arctic and temperate species (Hansen et al., 2011). Hjorth and Nielsen (2011) also
performed a comparative assessment of faecal pellet production, egg production, and hatching
success in Greenland Calanus finmarchicus and Calanus glacialis populations when exposed to
oil. During this study the two copepod species were exposed to daily nominal pyrene
concentrations of 0, 0.01, 0.1, 1, 10 and 100 nM at water temperatures of 0.5, 5, and 8C for 9
(C. finmarchicus) and 7 days (C. glacialis). While hatching success in both species was
unaffected by pyrene exposure, results showed that in C. finmarchicus, faecal pellet production
decreased with increased temperature and pyrene concentrations, while C. glacialis showed no
negative effect to pyrene over 7 days. Egg production in C. finmarchicus was also negatively
affected by pyrene at all temperatures while C. glacialis only showed a dose-dependant decrease
in egg production to pyrene at a temperature of 0.5C.

Although the majority of zooplankton toxicology work has been conducted on species found in
temperate climates, results are often used to infer the response of Arctic species under similar
conditions (Hansen et al., 2011). Research by Jensen and Carroll (2010) shows that in many
cases, direct links cannot, and should not be drawn between temperate zooplankton species and
their more northern cousins. As they point out, Arctic zooplankton species typically have longer
life spans which can lead to increased duration of oil exposure, and contain much higher lipid
concentrations allowing them to survive for long periods of time without food. These and other
differences between temperate and Arctic zooplankton species should be a clear indication that
research focused on Arctic species must be conducted to investigate the effects of oil spills on
zooplankton populations.

The lethal toxicity of Norman Wells crude to the jellyfish, Halitholus cirratus, exposed to 2
mL/L was 100% mortality, while mortality for the copepod, Calanus hyperboreas, was only
about 10% (Percy et al., 1985). This concentration was used as a worst case scenario. Exposure
109

to Atkinson Point crude caused 30% mortality to H. cirratus, but did not kill C. hyperboreas,
whereas exposure to Pembina crude caused 100% mortality in H. cirratus but only 35% to C.
hyperboreas. The jellyfish lost its ability to swim due to narcosis induced by the oils. This
narcosis was reversible by transferring the animal to clean seawater, as long as the oil dosage
was light to medium. The results highlighted the variation in response to different crude oils that
is dependent upon oil chemistry.

Results of studies looking at the effects of crude oil WAF on the copepods Calanus finmarchicus
(temperate) and Calanus glacialis (Arctic) show that both species are significantly affected by
the presence of oil, although in different ways (Hansen et al., 2011; Jensen and Carroll, 2010).
Experiments conducted with Calanus finmarchicus show that adult feeding success decreased in
the presence of WAF. Studies of Calanus glacialis show that while faecal pellet and egg
production in females was initially unaffected by exposure to WAF due in part to their more
extensive energy stores, egg hatching success was significantly reduced.

Toxicity tests with the Arctic copepod, Calanus glacialis were conducted by (McFarlin et al.,
2011c; 2011d) using WAF and CEWAF prepared from fresh Alaskan North Slope (ANS) crude
oil using a modification of a spiked exposure/continuous flow protocol (Aurand and Coelho,
2005), designed to simulate site-specific conditions of Arctic open waters. The Corexit 9500
dispersant was also tested for its toxicity. Biodegradation of chemically and physically dispersed
fresh and weathered ANS petroleum was measured by CO
2
production in a respirometer and by
GC/MS analysis. Results indicated that the WAF generated under a low energy regime did not
have a sufficiently high concentration of TPH to be toxic to the copepod; however, under high
mixing energy, the WAF LC
50
as TPH for Calanus was >3.7 mg/L. The CEWAF LC
50
values as
TPH for early and late season copepods were 22 mg/L and 62 mg/L respectively (McFarlin et al.,
2011c; 2011d).

According to a simulation model for northeast Arcto-Norwegian cod (Gadus morhua) using
empirical data, recruitment was highly sensitive to reduction of zooplankton biomass in zones of
cod larvae (Stige et al., 2011). Zooplankton, particularly the dominant copepod Calanus
110

finmarchicus, are major prey for the cod larvae and the biomass could be reduced by oil spills,
but this is an area requiring study.

Although cod eggs and larvae are affected by 2.5 g/L TPAH (Brude et al., 2010; Stige et al.,
2011), it is uncertain how this concentration would affect zooplankton biomass, since C.
finmarchicus may be less susceptible to oil components than cod larvae (Stige et al., 2011), and
the later stages of Calanus can survive exposure to PAH WAF at 3.5-10 g/L, although at 10
g/L hatching success of C. glacialis was lower, and at 7 g/L C. finmarchicus exhibited reduced
feeding (Jensen and Carroll, 2010). C. finmarchicus showed inhibition of feeding and egg
production on exposure to 20 but not 2 g/L pyrene, a relatively low toxicity PAH (Jensen et al.,
2008). Reduction in copepod nauplii, the main food items for larvae of Arcto-Norwegian cod, G.
morhua (Ellertsen et al., 1989), and polar cod, B. saida (Drolet et al., 1991), during their first
months of life, could significantly impact cod recruitment; however, we have not found any
published studies on the direct effects of PAH exposure to nauplii.

The synergistic effect of ultraviolet (UV) sunlight and 2 g/L TPAH from ANS crude caused
photoenhanced toxicity and mortality in Calanus marshallae and Metridia okhotensis, North
Pacific copepods (Duesterloh et al., 2002). The interaction of the PAHs and UV radiation was
highly significant (p < 0.005). Other effects were impaired swimming, and lipid sac
discoloration. There was 80-100% mortality and morbidity of C. marshallae exposed to UV and
oil compared to less than 10% effect in oil-only or UV only. For M. okhotensis, 100% mortality
occurred in the UV and oil treatment, 43% mortality and 27% morbidity in the UV-only
treatment, and less than 5% effect in the oil-only treatment. Bioaccumulation factors were about
8000 for C. marshallae and 2000 for M. okhotensis. Whether bioaccumulation occurs in fish
feeding on affected copepods, or how the fish might be affected, is not known.

A laboratory study of ISB of a slick of Norwegian Troll crude showed 33% immobility of C.
finmarchicus kept for 96 h in water recovered from under the burn, when the slick was burned
immediately (Faksness et al., 2011). When the slick was left for two days and then burned, there
was only 19% immobilization. Mortality in either treatment was less than 50%.

111

Fish

Arctic seas and adjacent waters contain over 400 species of finfish including marine,
diadromous, anadromous and freshwater fish that frequent brackish waters (ArcOD, 2010;
Logerwell and Baker, 2011). Of these, approximately 240 species are marine. Common to the
Chukchi and Beaufort Seas are capelin, Pacific herring and rainbow smelt. The dominant Arctic
families are cods (Arctic, Pacific and saffron), eelpouts, snailfish, sculpins (shorthorn and
staghorn), flatfish (Bering flounder, starry flounder, Alaska plaice, yellowfin sole and Greenland
turbot) and salmonids. Commercial fisheries do not exist in the high Arctic because traditional
methods of collecting fish by trawls do not work well in ice-covered waters. The lack of high-
Arctic fisheries catch and by-catch data is a void of basic knowledge (ArcOD, 2010).

Certain species live near the underside of the ice and are noted for having antifreeze compounds
in their blood. Most live on or near the bottom. Small fish feed on ice algae and zooplankton.
Polar cod (Boreogadus saida) use the underside of the ice as a nursery (Krembs and Deming,
2011). The amphipods dwelling on the underside of the ice as well as planktonic copepods are
the major prey for B. saida, which are themselves a key prey of seals, birds and whales (ArcOD,
2010) in the Arctic food web (Figure 23). The polar cod (P. cod in Figure 23), is the most
northerly distributed gadid, occurring roughly between 60N and the North Pole, nearshore as
well as offshore.

Factors affecting the toxicity of oil spills to fish populations include the physical state of the
ocean (especially mesoscale circulation), distribution of spawning sites, length of the spawning
season, time of the spill, trophic effects, and chronic sublethal effects of persistent residues
(Hjermann et al., 2007). Fish are able to metabolize hydrocarbons fairly quickly therefore
bioaccumulation is not thought to affect them (Boyd et al., 2001).

Existing studies on the effects of oil on fish have been largely focused on commercial species
such as Atlantic cod, salmon and herring. Stige et al. (2011) noted that juvenile and adult cod are
highly mobile and widely distributed, therefore it is unlikely that they would be impacted at the
population level by spilled oil. Hjermann et al. (2007) stated that although there is little evidence
112

to date that oil spills have a significant impact on fish stocks, this does not mean that fish stocks
cannot be significantly affected by oil spills. The authors noted that the vast majority of the data
on large spills comes from temperate and subtropical environments where biological activity is
high, ecosystems are complex, and fish tend to spawn for longer period of time compared to
species found in the Arctic. Although adult fish are largely able to detect and avoid oil, fish eggs
and larvae are not, which means that temporal and spatial spawning characteristics of Arctic fish
species are of primary importance. Short and intense spawning seasons and relatively limited
(although highly concentrated) distribution of eggs and larvae in the Arctic result in increased
susceptibility to the detrimental effects of oil spills. There is also evidence that certain juvenile
fish species found in the Arctic such as the polar cod are highly susceptible to decreased
zooplankton production as a result of an oil spill (Stige et al., 2011).

Biochemical biomarkers have been frequently used to monitor stress response in fish exposed to
petroleum hydrocarbons. Eggs, larvae and juveniles (the most susceptible life stages of fish due
to their low mobility and restricted distribution) of Atlantic cod (G. morhua) were exposed to the
WAF of North Sea Statfjord B crude oil (40-300 g/L) for 1-6 weeks (Stige et al., 2011) to
examine the CYP1A response using an antibody indirect ELISA technique (Goksyr et al.,
1991). Monoaromatics (benzenes, toluene and xylenes) dominated the WAF. Exposures that
began with the eggs and continued through hatching to larvae showed that induction of CYP1A
did not occur during the first days following hatch. After about five days, larvae showed a dose-
response induction of CYP1A. Even the low 40 g/L level, which is reasonable for actual spill
concentrations, elicited induction. Although petroleum hydrocarbons can accumulate in fish eggs
and be transferred to the resulting larvae (Solbakken et al., 1984) the CYP1A system was not
immediately induced in the hatchlings (Goksyr et al., 1991). Wong et al. (2011) showed that
production water from Canadian east coast platforms can affect fertilization and hatching success
in cod eggs but only at high concentrations.

Lyons et al. (2011) measured hepatic ethoxyresorufin O-deethylase (EROD) activity in the
Atlantic cod (G. morhua) and showed that water temperature affected the temporal response as
well as the absolute level of EROD activity in livers of fish exposed to CEWAF. Fish exposed to
CEWAF at 2C showed lower EROD induction than fish exposed at 7C or 10C. The EROD
113

response was also detectable more quickly at the higher temperatures than at 2C. Short-term
exposure of juvenile Atlantic cod to WAF and to CEWAF prepared with chemical oil dispersant
formulations significantly increased EROD activity (Lee et al., 2011b). Maximum EROD
induction occurred at 24 hours post-exposure and remained elevated over the next 48 hours
compared to controls. As expected, oil dispersants which elevated the concentrations of oil in the
water induced higher levels of EROD activity.

Two studies have examined EROD inducibility in the anadromous species, Arctic charr
(Salvelinus alpinus). Jrgensen and Wolkers (1999) measured hepatic EROD activity in winter-
and summer-acclimated charr at two temperatures, 1 and 10C. Strong EROD induction was
observed 2, 6, 12 and 23 days following one oral dose of 60 mg benzo[a]pyrene (BaP) per kg
body weight. Induction was delayed and prolonged in the lower temperature group but showed
comparable levels of induction. The magnitude of induction was seven times greater in the
summer experiments than the winter ones. The authors concluded that Arctic charr are a suitable
species for oil and gas monitoring in the Arctic. This conclusion was supported by Jonsson et al.
(2003) who also reported strong EROD induction in gill filaments of Arctic charr, but not spotted
wolfish (Anarhichas minor), following a 48-hour exposure to the model PAH, -naphthoflavone.
The results of these studies suggest that Arctic charr may indeed be a good indicator species, and
EROD induction a good biomarker for monitoring Arctic petroleum impacts.

Three consecutive year classes (1990-1992) of pink salmon fry (Oncorhynchus gorbuscha) that
had been exposed for ten days to seawater containing either 2554 g/L or 178349 g/L of
WAF North Slope crude oil were released into the Pacific Ocean where they completed their life
cycle (Birtwell et al., 1999). The WAF was composed primarily of monoaromatics. Preliminary
tests revealed that concentrations higher than the exposure solutions were acutely lethal to the fry
(96-h LC
50
= 1 to 2.8 mg/L). There had been no effect from the exposures on the survival and
growth of the fry before their release. The fish were tagged, and adults that returned to their natal
river in British Columbia did not show any detectable deleterious effect on growth or survival
compared to unexposed controls.

114

Pink salmon embryos are sensitive to 1-20 ppb PAH (Carls et al., 2001b; Heintz et al., 1999)
showing abnormalities comparable to Pacific herring embryos (edema, damage to skeleton, fins
and chromosomes). The effects apparently are cumulative across the life stages of an individual
resulting in a 50% mortality before being able to reproduce (Carls et al., 2001b). Pink salmon
incubated in oiled gravel that amounted to TPAH exposure of 5 and 19 mg/L (low and high
dose) were not found to stray from their natal stream any more than controls which had been
incubated in unoiled gravel (Wertheimer et al., 2000). The tagged fish in the exposed groups
were found straying with frequencies of 0.030 and 0.025 for the low and high dose treatments
respectively. The control fish strayed with a frequency of 0.023. Differences among treatments
were not statistically significant. Their results do not support the hypothesis that oil exposure of
embryos in intertidal spawning grounds was responsible for the high rates of straying of wild
stock pink salmon that were observed in PWS after the Exxon Valdez oil spill (Wertheimer et
al., 2000). Nevertheless, there is a burden of evidence that points to oil from the spill causing
elevated embryo mortalities and elevated CYP1A levels in larvae from streams oiled by the spill
(Carls et al., 2001b).

PAHs are known as the most toxic constituents of oil and petroleum related-products (Carls et
al., 2001b; Heintz et al., 1999). Fish embryos exposed to PAHs have shown mortality, edema,
haemorrhaging, cardiac disruption and deformities all of which seem to be due to high
bioaccumulation and limited biotransformation. Weathered oil retains the heavier 3 and 4-ring
PAHs. In low 0.4 part per billion (ppb) TPAH concentrations, these have been shown to cause
increased mortality, abnormalities, poor swimming, decreased incubation time and reduced
length-at-hatch in Pacific herring eggs exposed for four or 16 days (Carls et al., 1999). Edema
and defects in chromosomes, skeleton and finfolds were observed.

Barron et al. (2004) tested mechanistic models for the chronic toxicity of complex mixtures of
PAHs in Pacific herring (Clupea pallasi) and pink salmon (O. gorbuscha) embryos from
southeast Alaska. The four models tested were: narcosis, AhR agonism, alkyl phenanthrene
toxicity, combined toxicity. Of these models the alkyl phenanthrenes were the primary cause of
early life stage sublethal toxicity in both species tested. This model predicted toxicity with 67-
80% accuracy. Naphthalenes were primarily associated with narcosis which contributed most to
115

embryo mortality. Because of very low concentrations of potent AhR agonists, these PAHs did
not contribute significantly to early life toxicity. It was concluded that the alkyl phenanthrene
(C2 to C4) component of complex PAH mixtures can be used to predict significant toxic effects
to early life stages of pink salmon and Pacific herring, but the model predictions did not account
for exposure time due to limitations in the data (Barron et al., 2004). Although the alkyl
phenanthrene model was a good predictor of toxicity, the authors recommend the use of TPAHs
to quantify exposure until these mechanistic models can be tested further.

Toxic effects on embryos have been found to be associated with the concentrations of alkyl PAH
(Hodson et al., 2007) which comprise more than 90% of the TPAH in petroleum. TPAH has
been recommended for continued use as a metric of exposure, because although it is simple and
does not depend on the differences in PAH composition or differences in potency, TPAH values
were found to correlate well with toxic effects observed in pink salmon and Pacific herring
(Barron et al., 2004).

Osmoregulatory function of the gills and cellular volume regulation were measured in European
flounder (Platichthys flesus) from Oslofjord, Norway, exposed to lethal or sublethal
concentrations of dispersant alone (Corexit 9527) or a dispersant-to-oil ratio (DOR) of 1:1,
mixed with North Sea crude (Baklien et al., 1986). The concentrations tested in the 14 day
sublethal experiment were 20 ppm of Corexit alone or a DOR of 10 ppm:10 ppm crude. For the
96 hour lethality study, 80 ppm Corexit alone was used, or a 40 ppm:40 ppm crude oil mixture.
In addition to survival, the following blood parameters were measured: osmolality, haematocrit,
plasma Na, plasma K, plasma Ca, plasma Mg, plasma Cl, plasma protein and plasma total free
ninhydrin positive substances. The lethality study produced approximately 50% mortality. Blood
parameters in fish surviving the oil-dispersant mixture were all significantly elevated as were
most parameters in the dispersant exposure alone (except for haematocrit, plasma K or plasma
Cl). However, exposure to sublethal concentrations of the dispersant or the oil-dispersant mixture
did not significantly elevate these blood parameters indicating that these stress factors only
become elevated prior to death and are consequently not good biomarkers for sublethal testing
(Baklien et al., 1986). It should also be noted that a DOR of 1:1 is not the currently
recommended dosage for field applications, but rather 1:20.
116


Oil related pathologies were also assessed by Lukin et al. (2011) in two species of freshwater
fish in sub-Arctic Russia. These authors studied the health of these fish in the Pechora River
which received a multitude of pollutants including chronic oil spills such as from the largest oil
spill in Russian modern history (ca. 100 000 tons) in 1994. The objective of this study was to
evaluate the use of histopathological abnormalities in fish as an indicator of environmental
quality. These abnormalities were measured in the gills, liver and kidneys of ide (Leuciscus idus)
and whitefish (Coregonus lavuretus). Results indicated that these lesions were caused by direct
toxicant effects and secondary stress effects (parasitism). The authors concluded that the impacts
observed in these fish were caused by both acute (short-term) and chronic (long-term) exposures
to pollutants in the Pechora River. Of the two species tested, whitefish appeared to be the more
suitable indicator. The authors noted that while these histopathological lesions have been
observed in other studies, and could be caused by a variety of contaminants, some were directly
linked to oil contamination in the river. It has been suggested that the migration of petroleum
hydrocarbons from a spill into brackish waters of Arctic estuaries would increase the risk of
impacts since oils seem to be more toxic in freshwater than saltwater (ESL Environmental
Sciences Limited, 1982).

Pacific sand lance (Ammodytes hexapterus) are an important benthic fish that form a link
between the zooplankton and upper trophic levels of the Arctic food web (Moles and Wade,
2001). The larvae serve as prey for salmon, while groundfishes feed on both the larvae and
adults. The sand lance is also an important prey of Arctic birds and mammals. Adults collected
from southeast Alaska were exposed to weathered ANS crude oil contaminated sediments for
three months at sediment concentrations of 10 and 61 g TPH/g (Moles and Wade, 2001).
Superoxide production, phagocytic function and gill parasites (Gyrodactylus sp.) were quantified
at the end of the exposure period. Fish produce superoxides to catabolize compounds that result
from the phagocytic action of leucocytes. Moles and Wade (2001) found that sand lance exposed
to the lower concentration (10 g TPH/g) had increased superoxide production and phagocytic
activity, with a corresponding decrease in the presence of the parasitic trematode, Gyrodactylus
sp. In the other treatment using 61 g TPH/g, the sand lance exhibited decreases in both
superoxide production and phagocytosis (decreased immune response), with an increase in
117

occurrence and abundance of Gyrodactylus. The differences were statistically significant (p <
0.05). These authors suggested that low levels of oil may stimulate the immune system which
results in lower numbers of gill parasites, but higher concentrations may overwhelm the immune
system resulting in higher levels of infection. These low concentrations of TPH are similar to
what can be expected in subtidal regions following an oil spill. Increases in Gyrodactylus were
also found in Atlantic cod exposed for four months to crude oil WAF (Khan and Kiceniuk,
1988).

Phototoxicity is emerging as an issue of concern as the toxicity of PAH and weathered oil has
been reported to increase with a factor of 2 to over 1000 in the presence of sunlight (Barron et
al., 2003). There are extended periods of light and darkness in the Arctic. The consequences of
these extreme shifts in photoperiod on biological effects have not been studied to any extent.
Many species at the lower end of the food chain are most prevalent during periods of extended
daylight hours. It is possible a spill of oil during the Arctic summer would not only affect a large
number of species but the toxicity of the oil could be increased by the production of toxic
products of photodegradation.

Photo-enhanced toxicity of WAF and Corexit 9527 CEWAF of weathered ANS crude oil was
studied in eggs and larvae of Pacific herring (C. pallasi) from Sitka Sound, Alaska (Barron et al.,
2003). At test concentrations below 50 mg TPAH/L, median values for LC
50
and for the
effective concentration affecting 50% of test organisms decreased with time after initial oil
exposure, and brief exposure to sunlight for 2.5 hours on each of two days significantly increased
toxicity 1.5 to 48 fold. Ultraviolet radiation-A (400 nm - 315 nm) treatments combined with
sunlight induced the greatest phototoxicity. The toxicity of chemically dispersed oil which
accelerated PAH dissolution into the aqueous phase was also significantly more toxic in sunlight.
For chemical dispersion exposures, the 96-hour no-observed-effect concentration in the presence
of UV-A and sunlight was 0.2 mg TPAH /L. Hence, weathered ANS is phototoxic and UV can
be a significant and causative factor in the mortality of early life stages of herring exposed to oil
and chemically dispersed oil. Barron et al. (2003) noted that most studies do not consider the
effects of phototoxicity and therefore may underestimate potential impacts of spilled oil.
118

Phototoxicity can be particularly important for pelagic species such as herring which inhabit the
photic zone and have translucent eggs and larvae, which allow UV penetration.

Weathered crude oil can cause immunosuppression and viral haemorrhagic septicemia in Pacific
herring (Hjermann et al., 2007). It is suspected that this was a factor in the collapse of the Pacific
herring industry following the Exxon Valdez spill, when the 1989 year class was exposed to oil at
the egg stage, and subsequently displayed a high incidence of viral haemorrhagic septicemia as
well as extremely low survival to spawning age.

In Pacific herring embryos from central Puget Sound, Washington, the primary target of crude
oil toxicity is the heart (Incardona et al., 2009). Pacific herring exposed to effluent from oiled
gravel columns showed defects in the heart rate and rhythm as soon as regular heart rhythm
could be observed (five days post-fertilization). This cardiac dysfunction was produced at low
PAH tissue burdens of only 0.8 mol/kg (total tricyclics). Consequently, for oil related impacts,
the authors recommended biomarkers which related to cardiac performance such as CYP1A or
levels of fluorescent aromatic compounds in bile.

Exposure to 2 mL/L of Norman Wells crude resulted in 100% mortality of the juvenile stage of
the fourhorn sculpin, Myoxocephalus sp. (Percy et al., 1985). In addition to the immediate
mortality effects associated with contact between fish and oil, perhaps a more worrying issue is
potentially longer lasting sublethal effects. Studies show that ingestion and absorption of oil by
B. saida, which is highly possible since they feed on epontic biota that could accumulate oil from
pooled oil under the ice, imposes a significant metabolic burden on exposed individuals which is
readily detectable through the analysis of various types of biomarker responses (Christiansen et
al., 2010; Jonsson et al., 2010; Nahrgang et al., 2010b; Nahrgang et al., 2009; Nahrgang et al.,
2010d).

Much of the research involving potential impacts of oil and gas related activities in the Arctic
region has focused on polar cod (B. saida). Baseline knowledge and toxicological studies on this
species provide useful information on its potential use in future biomonitoring programs for the
Arctic oil and gas industry for these reasons: the Arctic food web is relatively simple and
119

seasonal; these cod have a circumpolar distribution; and baseline levels of bile PAH-metabolites
and liver EROD were found to be very low but highly responsive to petrogenic exposure. Polar
cod were recommended as an indicator species by the Arctic Monitoring and Assessment
Programme (AMAP, 2008; Stange and Klungsr, 1997) because of its low trophic level diet
(zooplankton) and its short lifespan, which provides a better indication of immediate conditions.
Furthermore, water column monitoring near oil installations in the Norwegian Sea and the Grand
Banks have utilized its temperate counterpart, the Atlantic cod (G. morhua), to quantify
environmental effects of their effluent, produced water.

It is important to determine differences in sensitivities within the Gadidae family because fish
inhabiting sub-Arctic and Arctic regions may be more sensitive to chemical insult as a
consequence of their adaptations to survival in freezing waters (Christiansen et al., 2010).
Because of these adaptations, Arctic fish may be particularly at risk from additional stressors
such as petroleum activities. Hjermann et al. (2007) also noted that, because of low water
temperatures, the physical and biological processes involved in oil degradation may be slower in
Arctic waters, hence increasing the duration of exposure to oil. These authors also point out that
the simplicity of polar ecosystems renders them more vulnerable to collapse if there are changes
to some of the keystone species. It is hypothesized that decreased recruitment of polar cod is
likely to be the major impact on the cod stocks of an oil spill in the Arctic (Logerwell and Baker,
2011). Because it is a keystone species, the effects of cod mortality, reduced recruitment and
reduced production are expected to cause changes in community structure, trophic alterations
due to reduced prey for dependent species and less food for humans.

Polar cod (B. saida) and larval sculpin (Myoxocephalus sp.) were exposed to WAF and CEWAF
of fresh ANS using a spiked exposure regime to simulate an oil spill. Oil was added at a given
concentration followed by dilution with a continuous flow of clean seawater (McFarlin et al.,
2011c; 2011d). The exposures with oil and the chemical dispersant Corexit 9500 were
conducted in open systems to permit the evaporation of volatile organic compounds that would
normally occur. Biodegradation of chemically and physically dispersed fresh and weathered
ANS was measured by CO
2
production in a respirometer and by GC/MS analysis. WAF
generated under a low energy regime was toxic to the cod and larval sculpin (LC
50
= 1.6 and 2.2
120

mg TPH/L respectively). Under high mixing energy, the WAF LC
50
for polar cod was 3.3 mg
TPH/L. The CEWAF LC
50
for cod was 55 mg TPH/L, and 27 mg TPH/L for the larval sculpin.

Effects of produced water on Atlantic cod living in the Norwegian Sea have been monitored
using PAH biomarkers (Hylland, 2006). Nahrgang et al. (2010b; 2009; 2010c; 2010d) examined
the applicability of these biomarkers to polar cod in a series of experiments. Polar cod were
exposed to BaP by intraperitoneal injection (Nahrgang et al., 2009). PAH biomarkers tested for
gene transcription analyses included -Actin, CYP1A-1, AhR-2, pi-class homolog of glutathione
S-transferase (GST), Cu/Zn-superoxide dismutase, Mn-superoxide dismutase, catalase and
glutathione peroxidase. Protein analyses included microsomal CYP1A, cytosolic GST and
catalase proteins. EROD was also quantified. All biomarkers were successfully measured in
polar cod but enzyme activities were the least reliable.

To further determine the applicability of these biomarkers to complex mixtures such as oil, and
through more realistic exposure conditions, Nahrgang et al. (2010b) exposed polar cod to
environmentally relevant concentrations of WAF of North Sea crude oil using oiled rock
columns. Molecular and biochemical biomarkers tested were similar to those in the Nahrgang et
al. (2009) study. The most responsive biomarkers were CYP1A-1 mRNA expression and EROD
activity. The authors recommended that GST-mRNA expression and GST enzyme activity be
used as complementary biomarkers, while the antioxidant defense biomarkers were deemed
inappropriate for monitoring purposes. EROD and GST induction were delayed and were still
elevated after 2 weeks of depuration.

Nahrgang et al. (Nahrgang et al., 2010c; Nahrgang et al., 2010d) also tested the response of the
suite of molecular biomarkers following a dietary exposure to North Sea crude oil. Polar cod
from Rijpfjorden, Svalbard, were fed for four weeks a diet of pelagic amphipod (Themisto
libellula) blended with gelatine powder and North Sea crude oil for final concentrations of 0.5
and 2 mg crude/g. This was followed by two weeks of depuration, being fed the control diet of T.
libellula blended with gelatine alone. The diet having concentrations of 0.1, 18.0 and 43.3 g
TPAH/g in the control, low dose and high dose respectively, was readily received by the cod.
Expression of the mRNA for CYP1A-1 in the exposed fish was strongly induced, up to 60 that
121

of the control for the high dose from two to four weeks, and returned to normal after the two-
week depuration. Expression of mRNA for the cytosolic pi-class homologue of GST followed a
similar pattern only for the high treatment (3 the control and returning to normal). Expression
of antioxidant genes showed no difference from controls over the course of the experiment. Liver
EROD increased by a factor of 10-12 in low and high treatments (about 30-50 pmol/min/mg
protein), and falling during the two weeks of depuration to normal (about 2-4 pmol/min/mg
protein) for the low treatment, but remaining significantly above normal for the high oil
treatment (about 8 pmol/min/mg protein). Synchronous flourescence scan spectrometry of the
bile in exposed fish indicated the presence of PAH metabolites of naphthalene, pyrene and BaP
in a dose-dependent manner, with maxima at week two, decreasing by week four, and returning
to control levels after two weeks of depuration. Responses of the molecular biomarkers were
similar to those observed following the WAF exposure leading to similar conclusions on the
usefulness of the various biomarkers tested. The similarity in the results of these studies
demonstrates the consistency of the response of these molecular biomarkers and validates their
use for biomonitoring studies.

High levels of BaP in water-exposed B. saida have been found partitioned into gills, olfactory
organ, kidney, skin, intestinal mucosa and liver (Ingebrigtsen et al., 2000). The highest
concentration was found in the bile. Partitioning occurred in epithelial cells in which CYP1A
activity takes place (Jnsson et al., 2003). Uptake of radioactive
3
H-BaP (mixed with unlabelled
BaP to give a concentration of 1 L/L) was predominantly from the water through the gills,
metabolism principally occurred in the liver, and excretion of metabolite was by way of bile
(Ingebrigtsen et al., 2000).

The cod accumulated BaP at -1C, which is not unusual since these fish have a very high
assimilation efficiency of about 80% (Hop et al., 1997) and long gastric half times ranging from
36 hours to 12 or 13 days at -1.4 to -0.5C (Hop and Tonn, 1998). The gastric evacuation rates
for polar cod are much longer than for Atlantic cod (about 13 hours). Evacuation rates for
starved polar cod are even longer (Hop and Tonn, 1998). In spite of the slow evacuation rates,
food contaminated with oil (and PAH) was evacuated in less than two weeks, with complete
excretion of PAH metabolites from the bile (Nahrgang et al., 2010c).
122


Digestive processes are slower at sub-zero temperatures and this might limit daily food intake
when food is abundant (Hop and Tonn, 1998). Polar cod have slow growth rates (0.15-0.25%
body weight per day, although this is faster than snailfishes), so the high annual production of
these fish is mainly because of their high growth efficiency and large standing crop (Hop et al.,
1997). These cod grow faster on Calanus copepods than on Themisto amphipods, and their high
growth and assimilation efficiencies at low temperatures makes them efficient converters of
energy between the zooplankton trophic level and the higher trophic levels of marine birds and
mammals (Hop and Tonn, 1998; Hop et al., 1997).

To further validate the use of polar cod as an indicator species, Nahrgang et al (2010d) measured
EROD activity in gills and liver following dietary and waterborne exposures. This was the first
study to measure EROD activity in gills of polar cod, one of the primary tissues to uptake
pollutants (Ingebrigtsen et al., 2000). EROD levels in both the gills and the liver were
significantly increased by both exposure routes, were dose-dependant, and were higher in the
liver than the gills. Time of exposure was indicated better by the gills than liver but the response
threshold was higher with gill tissue which could prove problematic for field monitoring of low
PAH levels and particularly in the presence of potential environmental confounding factors
(Nahrgang et al., 2010d).

The biochemical biomarker CYP1A has also been examined in polar cod (George et al., 1995).
Because CYP1A induction has been shown to be reduced or absent at low temperatures the
authors questioned the validity of this biomarker for cold Arctic regions. However, CYP1A can
be induced at 0-6C in male and female polar cod during sexual maturation, by -naphthoflavone
or crude Oseberg C mineral oil (George et al., 1995). Polar cod (B. saida) from the Barents Sea
and off Svalbard were kept at 3-6C and fed a diet contaminated with 0.2 mg/g crude oil for 52
days. At sexual maturity, the females showed diminished CYP1A activity, which has been
demonstrated in many other fish species. Following a three-week depuration, the fish were
sacrificed for EROD analysis. Controls at the start of the experiment ranged from 28-31
pmol/min/mg protein. Sexually mature females at 72 days had an EROD activity of 8 2
pmol/min/mg protein. EROD activity for exposed males at 72 days was 132 14 compared to
123

females which measured 42 6 pmol/min/mg protein. The study showed that sexual maturation
in female polar cod does not suppress the CYP1A system. In spite of the fish having ceased
feeding three weeks prior to the end of the experiment (the depuration phase) because they
naturally do this in preparation for spawning, they continued to assimilate the ingested food (up
to two weeks before clearance at low ambient temperatures), and therefore the inductive effect of
the oil in the diet was prolonged (George et al., 1995). This study clearly demonstrated that the
CYP1A enzyme system was induced by common inducers in both males and females during
their sexual maturation at a temperature of 0C. Christiansen and George (1995) also showed
that crude oil dietary exposure did not impair the overall feed intake of exposed polar cod,
despite growth reduction in exposed fish.

Whole body metabolism, inferred from oxygen consumption rates (respirometry), of polar cod
exposed acutely for about 60 minutes and chronically for 4 weeks to the WAF of North Sea
petroleum was tested by Christiansen et al. (2010). This biomarker reflects the stress and
bioenergetic costs of petroleum exposure. The authors expected oxygen consumption to increase
in exposed cod but both chronic and acute exposure actually resulted in reduced routine
metabolism. A review included in this paper demonstrated the variability (increase, decrease and
non-response) of this biomarker among fish species, indicating the necessity of species-specific
calibration.

Another potential biomarker, DNA adducts, are pieces of DNA covalently bonded to
carcinogenic chemicals. Because DNA adducts can be caused by PAH exposure, Aas et al.
(2003) established baseline levels of DNA adducts in the livers of 11 Arctic and sub-Arctic fish
species, including polar cod, from pristine areas. The technique employed in this study (the
nuclease P1 version of
32
P-postlabelling) was selected because it can detect DNA adducts caused
by PAHs. In all species, DNA adduct levels were undetectable. Laboratory exposure to BaP and
to crude oil caused the induction of DNA adducts in polar cod, which validated the use of this
biomarker and the nuclease P1 version of the
32
P-postlabelling technique as a tool in oil and gas
monitoring studies.

124

Polar cod (B. saida) from Svalbard were exposed for two weeks to 1 mg/L dispersed, crude
Statfjord A, in a continuous flow system (Jonsson et al., 2010). The WAF was analyzed for 26
PAH species (EPA 610 protocol) and was primarily composed of 2 and 3-ring PAHs. The sum of
these was 19.9 g/L, while the concentration of the 2-ring PAHs ranged from 1.28 to 6.79 g/L,
and for 3-ring was 0.005 to 0.408 g/L. Examination of exposed fish revealed BaP-type
metabolites in bile, indicating metabolism of PAHs. The 5-ring PAHs had been bioconcentrated
from the water into the fish. Control fish were found to have very low levels of bile PAH-
metabolites, in the range of 1899-1940 ng/g for the sum of the means of naphthalenes,
phenanthrenes and pyrene, compared to exposed fish that had a mean sum of about 202,000 ng/g.
The highest EROD liver activity in control fish was 5.7 pmol/min/mg protein, while in the
exposed group, the values ranged from 9.5-80.7 pmol/min/mg. Hepatic DNA-adducts (DNA
bound to a carcinogen) were low in control cod (mean of about 0.4 nmol adducts/mol normal
nucleotides) compared with the exposure group (12.2 3.5 nmol adducts/mol normal
nucleotides), indicating that exposed fish could potentially suffer genotoxic effects and therefore,
cancer. EROD and DNA-adduct values were highly correlated, which is grounds for using the
EROD assay alone instead of the more complex DNA-adduct assay. PAH-metabolites in the bile
and CYP1A induction (EROD) in the liver of polar cod were suggested as useful biomarkers
(Jonsson et al., 2010).

Xenobiotic excretion patterns of polar cod were described by Christiansen et al. (1995). In fish,
foreign chemicals are primarily excreted via urine and bile. Xenobiotics and their metabolites
excreted via urine are filtered in the glomerulus and secreted by tubular transport within the
kidney. Unlike most fish species, including Atlantic cod, polar cod display a complete absence of
the glomerular apparatus. Consequently, Atlantic cod excreted an inert, model polysaccharide
solely in the urine while polar cod excreted it in the bile.

To our knowledge only one study has reported acute toxicity concentrations of oil for polar cod
(Carls and Korn, 1985). The reported LC
50
was 1.6 ppm total aromatic hydrocarbons in WAF of
Cook Inlet crude oil, reaching a stable LC
50
after five days exposure. Polar cod responded most
quickly and proved most sensitive among six Arctic species tested which included three species
of amphipod, a mysid, and a sculpin. Polar cod were also most sensitive to the aromatic
125

reference toxicant, and constituent of Cook Inlet crude oil, naphthalene with an 8 day LC
50
of
1.35 ppm.

Also lacking is information on effects of oil (or PAH) exposure on the embryonic and larval
stages of polar cod. Because these ichthyoplanktonic life stages have limited mobility, they are
most vulnerable to localized environmental insult, such as oil spills or an influx of produced
water. Early life stages, being periods of rapid development and morphological re-organization,
are particularly sensitive to chemical exposure (Weis and Weis, 1989). Adults spawn under the
ice in nearshore waters from November to March and the buoyant eggs remain at the surface
until hatch, after which the young-of-the-year larvae continue to reside near the surface to feed
for several months (Sekerak, 1982). Exposure routes would be adsorption (direct contact),
absorption through skin and gill membranes, and ingestion. There are no field data for egg and
larval exposure, and few laboratory studies have been done (Logerwell and Baker, 2011). This
knowledge gap is significant, since polar cod can be considered a keystone species (Figure 23).

As noted by Hjermann et al. (2007), some species of fish in the Arctic such as Atlantic cod and
herring, are close to their climatic boundary and environmental conditions in these habitats may
limit their spawning season to short, intensive and localized spawning events. Consequently,
these populations may be at greater risk if a spill were to occur during their spawning season. In
addition to the morphological and genetic abnormalities that PAHs are known to cause, they
might also be responsible for disruption of complex behaviour such as predator avoidance, and
reproductive and social behaviour in adult fish which have been exposed to PAHs during their
early life stages (Hjermann et al., 2007).

Benthic Invertebrates

Dead ice algae falls off the under surface of ice and sinks, along with other dead organisms and
fecal pellets, providing food for a complex benthic ecosystem on the aphotic sea floor that is
home to over 90% of the approximately 5000 species of Arctic marine invertebrates (ArcOD,
2010). On the continental shelves, crustaceans (amphipods, sand fleas, crabs), polychaetes
(bristle worms) and bivalves (clams and mussels) dominate the sediment infauna (ArcOD, 2010;
126

Blanchard et al., 2010). The larger epibenthic animals are usually dominated by brittle stars with
up to several hundred individuals per square metre. Other epifauna include sea cucumbers,
gastropods, ascidians, sponges, bryozoans, hydrozoans, shrimp, starfish, sea anemones, sea
urchins and hermit crabs.

Food supply rather than water temperature is the limiting factor for the Arctic benthic
community (ArcOD, 2010). In certain of the shelf areas such as the Chukchi and Bering Seas,
the benthos consists of an abundant and diverse community (Logerwell and Baker, 2011)
including large worms, bivalves and amphipods which are prey for gray whales, bearded seals
and walruses (Blanchard et al., 2010). These macrofauna are supplied mainly by pelagic and ice-
edge sources (Figure 22) such as senescent phytoplankton (Aydin et al., 2002). Food availability
and benthic biomass in deep Arctic basins are much lower, although the major animal groups,
comprising about 350-400 species, are similar to those of shelf regions.

The BIOS Project in the Canadian Arctic provided an opportunity to collect nearshore benthos
data systematically from the same area over a four-year period (Snow et al., 1987). Chemically
dispersed and undispersed crude oil was spilled in shallow bays. Bivalve molluscs, Mya truncata
and Macoma calcarea, were examined for histopathological and biochemical responses. Mya
truncata (blunt gaper edible saltwater clam) were collected for biochemical analysis immediately
before, immediately after and two weeks after the simulated spills. Concentrations in the clam
tissues of glucose, glycogen, trehalose, total lipid and free amino acids were measured. The
results of the biochemical analyses indicate that Mya from the bays were not severely stressed by
either dispersed oil or oil alone. It was concluded that the acute effects of dispersed oil are
greater than those of undispersed oil, but effects of undispersed oil on infaunal molluscs develop
more slowly and persist longer than those from dispersed oil (Neff et al., 1987). Humphrey et al.
(1987) examined the fate of chemically dispersed and untreated crude oil in Arctic benthic
bivalves and urchins. Physiological parameters measured in bivalves exposed to oil included
elements of scope for growth, activity of aspartate aminotransferase and glucose-6-phosphate
dehydrogenase. Dose-response relationships between physiological rates and hydrocarbon body
burden were apparent. Depuration of the stored hydrocarbons occurred during the experimental
127

recovery period. Furthermore, in vivo biodegradation of hydrocarbons was indicated in the
bivalves.

Presently, there have been no surveys of the Arctic salt marsh benthos. Benthic communities of
the intertidal zone are at risk of direct oiling, which in addition to the toxicity of the WAF
containing organic compounds, can simply smother the biota (NRC, 2003). The subtidal benthos
consists of oligochaetes, midge larvae, amphipods and isopods (Logerwell and Baker, 2011). In
both the intertidal and subtidal zones, the most sensitive organisms tend to be the crustaceans
(Dean and Jewett, 2001; NRC, 2003). Residual Exxon Valdez oil in the intertidal zone of
Latouche and Evans Islands is suspected to be putting the intertidal invertebrates and other
species at risk, but effects have not been examined; however, oil under the soft sediments of
mussel beds that have resulted in PAH-contaminated mussels indicates the bioavailability of the
residual oil, and the potential for contamination of higher trophic level species that feed upon
mussels (Carls et al., 2001a).

Two polychaetes species, Capitella capitata and Cirratulus cirratus, which inhabit sediments of
the Beaufort Sea are resistant to pollution in general, more specifically, are resistant to effects
from fuel oil (ESL Environmental Sciences Limited, 1982; George, 1971). C. capitata has been
shown to accumulate and metabolize benz[a]anthracene, and this metabolic pathway can be
induced by exposure to oiled sediments (Forbes et al., 2001).

The gammarid amphipod, Gammarus setosus, common to the intertidal zone of Svalbard,
exhibited about 60% less stability of lysosomal membranes when exposed to chemically
dispersed, naphthenic rich Norwegian crude oil (Troll) for 12 d (Faksness et al., 2011). This is an
indicator of stress. Lipids in the gammarids suffered peroxidation by day 12 in WAF and
CEWAF treatments, but the animals recovered after another 13 days in the flow-through system.
The mean body burdens of semi-volatile organic compounds from WAF and chemically
dispersed treatments were about 1.8 and 5.6 g/g respectively on day 12, decreasing to about
0.25 and 0.6 g/g by day 25. The authors state that differences in response between temperate
and Arctic benthic communities is likely due to different community structure, sensitivity of
individual taxa, and different contamination history of the two areas.
128


Typical biomarker responses developed for temperate species may be affected at cold
temperatures. Olsen et al. (2007) looked at benthic community response to petroleum-associated
components in Arctic versus temperate marine sediments. In addition to a control, two oil
concentrations were used for the study. After 21 days of incubation, significantly higher
sediment oxygen demand was measured in the Arctic high oil concentration (1.6 mL oil/kg)
sediment cores, and sediment cores injected with drill cuttings, compared to Arctic control cores.
In the temperate cores no significant differences were seen. The variance in response was likely
related to differences in the benthic community structures, sensitivity of individual taxa to oil-
related compounds and contamination history of the two areas.

Alterations in the energy budget of a bivalve, Liocyma fluctuosa, and two Arctic amphipods,
Gammarus setosus and Onisimus litoralis, exposed to oil-related compounds were investigated
(Olsen et al., 2007). The cellular energy allocation biomarker measures the energy budget of
organisms by biochemically assessing changes in carbohydrates, protein and lipid content as well
as the electron transport system activity. Significantly lower cellular energy allocation values
and higher electron transport system activity as the cellular aspect of respiration were observed
in Gammarus setosus subjected to WAF of crude oil and drill cuttings compared to controls.
Higher cellular energy allocation value and lower cellular respiration were observed in Onisimus
litoralis exposed to drill cuttings compared to controls. No difference in the energy budget of
Liocyma fluctuosa was observed between treatments. Different responses to oil-related
compounds between the three test species were likely the result of differences in feeding and
burrowing behaviour and species-specific sensitivity to petroleum-related compounds (Olsen et
al., 2007).

Studies on the effects of oil pollution on the sea ice amphipod G. wilkitzkii have taken place
because of increased petroleum related activity in the Barents Sea of the European Arctic. Hatlen
et al. (2009) exposed the amphipods to a low and high dose of the WAF of oil for 36 days and
for 113 days, and saw a dose-related significant increase in respiration rate. Sublethal effects
were seen for the low dose group using the biomarkers malondialdehyde and total oxyradical
scavenging capacity (TOSC). Cellular energy allocation was measured in sea ice amphipods
129

after a one month exposure to a high, medium or low dose of WAF of oil. Significantly higher
protein content was observed in the medium dose compared to controls. The total energy budget
was not affected in G. wilkitzkii (Olsen et al., 2008). Camus and Olsen (2008) exposed female G.
wilkitzkii to a high, medium and low dose of WAF of crude oil for 30 days and their embryos
were scored for abnormalities. The frequency of embryo aberrations was significantly higher in
the high-dose compared to controls.

The amphipod Boeckosimus affinis, widespread in the North American Arctic, was exposed to
the WAF of Prudhoe Bay crude oil and survival, movement and food search success was
monitored. Mortality was correlated to the strength of the solution and chronic exposures were
more detrimental than one-time exposures. Movement and time spent moving decreased as the
concentration of WAF increased in the chronic exposure experiments. Animals that were
exposed once moved less distance but the time spent moving was not decreased. Feeding success
of the amphipods was also reduced after exposure to the WAF of crude oil but animals exposed
just once showed recovery in feeding success after removal of the pollutant (Busdosh, 1981).
The Arctic amphipods Gammarus zaddachi and Boeckosimus affinis, the dominant amphipod
species in coastal pond and marine lagoon ecosystems were exposed to Arctic diesel and the
WAF of Prudhoe Bay crude oil in bioassays. Mortality and respiration, sensitive indicators of an
animals physiological and metabolic state were measured. It was determined that spills of crude
oil or refined oils in the Arctic are likely to cause large-scale mortality in the region of the spill,
resulting in serious ecological changes in decomposition processes and food-web relationships
(Busdosh and Atlas, 1977).

Gammarus oceanicus, sympagic amphipods, were collected from the tidal zone in Kongsfjorden,
Svalbard, and exposed in the laboratory to WAF, CEWAF and emulsion (a constantly mixed
test tank of CEWAF) of Statfjord A+B crude oil, and the dispersants Finasol OSR-5 and OSR-12
(Aunaas et al., 1991). The DOR was 1:1 using Finasol OSR-12 and 1:10 using Finasol OSR-5
according to the manufacturers recommendations. Test concentrations in the sublethal exposure
experiments were based on preliminary acute toxicity experiments to establish LC
50
values, and
using concentrations that fell within the LC
50
range of 0-10%. The concentrations of WAF were
20% for crude oil, 10% for OSR-5 and the crude oil + OSR-5, and 100% for OSR-12 and crude
130

+ OSR-12. The concentrations used in sublethal exposures to water emulsions were 30 mg/L
crude oil, 20 mg/L Finasol OSR-5, 300 mg/L Finasol OSR-12, 200 mg crude oil + 20 mg OSR-5
per litre seawater, and 100 mg crude oil + 100 mg OSR-12 per litre seawater.

Finasol OSR-5 WAF induced the highest mortality, followed by CEWAF with this dispersant.
Crude oil WAF alone was less toxic. Finasol OSR-12 plus crude (CEWAF), and WAF of Finasol
OSR-12 alone, caused only moderate mortality. The OSR-12 dispersant was less toxic than
OSR-5. Adding Finasol OSR-12 to the crude oil caused a reduction in the mortality of the
amphipods compared with amphipods exposed to the WAF and the water emulsion of crude oil
alone. This reduction in toxicity with the dispersant was attributed to a corresponding reduction
in the mole fraction of toxic oil components in the mixture.

Exposure to sublethal concentrations of WAF increased the respiratory rates of the amphipods in
most cases (Aunaas et al., 1991). WAF slightly increased the concentrations of sodium in the
haemolymph and whole organism. Some WAF exposures resulted in a significant increase in the
water content of the amphipods, perhaps because there was an increase in membrane
permeability to water and ions. The increased respiration probably occurred as part of the
corresponding osmoregulation process.

Exposure to sublethal concentrations of water emulsions led to reduced respiration (probably
from oil droplet adherence to gill membranes, thereby reducing oxygen diffusion), and increased
concentrations of sodium in the haemolymph and the organism as a whole (Aunaas et al., 1991).
The decrease in total free amino acid concentration in these amphipods was ascribed to volume
regulation of swollen cells, and reduced co-transport of sodium and free amino acids from the
haemolymph and into the cells.

Exposure to 2 mL/L of Norman Wells crude resulted in a range of mortalities to amphipods
(Onisimus affinis, 85%; Corophium clarencense, 58%; Atylus carinatus, 7%) but isopods
(Mesidotea entomon, M. sibirica, M. sabini) experienced less than 1% mortality (Percy et al.,
1985). In fact isopods were unaffected by direct contact with the oil. The metabolism of
amphipods was inhibited in low oil concentrations, but stimulated by high concentrations.
131

Contaminated sediments were acutely toxic to the amphipod O. affinis, but another species was
unaffected. The barnacle, Balanus crenatus, suffered approximately 74% mortality under the
same exposure regime. From the data, it was concluded that a severe spill would eliminate
sensitive species and allow other resistant ones to survive and dominate in the short-term due to
reduced competition. Epifauna that are unable to escape would suffer immediate effects from
WAF, and the worst affected in the long-term would be the infauna if the oil enters the
sediments.

A temperate species, the Arctic spider crab (Hyas araneus), was exposed to PAHs via sediment
and injection. After two weeks of exposure, heart rate showed a significant increase compared to
controls. Respiration was not affected by either oil treatment. The basal oxygen consumption of
control Hyas araneus was lower than reported in those living in temperate water. Results
indicated that although decreased uptake and metabolism of oil compounds into reactive oxygen
species (ROS) was mainly due to the low temperature reducing the bioavailability of PAHs, the
relatively low metabolic rate of Arctic Hyas araneus was also implicated (Camus et al., 2002a).
Two cellular endpoints, haemocyte counts and protein levels, were significantly elevated
following the dispersed oil exposure. Cell membrane stability and phagocytosis demonstrated a
significant reduction. These results indicate alterations in the immune endpoints measured but
they appear reversible upon removal of the stressor.

Arctic scallops could be used as sentinel species in the Arctic in addition to the common blue
mussel, Mytilus edulis. The common blue mussel extended its distribution recently to the
Svalbard Archipelago, where it was discovered at Sagaskjret, Isfjorden (Berge et al., 2005).
The population seems to be young, having become established during the summer or autumn of
2002 from larvae transported to Spitsbergen by the West Spitsbergen Current. Horse mussels
(Mytilus trossulus) were collected at Barrow, Alaska, during the 1990s (Feder et al., 2003).

Camus et al. (2002b) injected Chlamys islandica adductor muscle with BaP every 24 hours for
four days. TOSC in the digestive gland was significantly reduced in exposed groups indicating a
depletion in oxyradical molecular scavengers. The antioxidant defences appeared to be
overwhelmed by the reactive oxygen species as the plasma membranes of haemocytes were
132

destabilized. The data indicate that ROS were produced by Arctic scallops via the metabolization
of BaP at 2C.

Biomarker responses in the blunt gaper clam, Mya truncata, exposed to sediments contaminated
with PAHs have been studied. After two weeks of exposure to sediments contaminated with
crude oil TOSC showed no change. The high TOSC value of Mya truncata (control group) is
thought to efficiently protect biomolecules with a low turnover rate in a low food availability
environment. Respiration rate of control and PAH-exposed Mya truncata was similar and
relatively low as is typical in polar bivalves, reflecting a strategy to minimize energy expenditure
to cope with nine months of starvation. In the exposed bivalves, the haemocyte cellular
membranes were significantly destabilized compared to controls. Bioaccumulation of PAH by
Mya truncata was low with mainly uptake of low molecular weight compounds (two and three
ring molecules). A combination of low metabolic rate and reduced solubility of the oil compound
at low temperature probably were the reasons for the low uptake (Camus et al., 2003).
Supporting these observations, a separate study by Hannam et al. (2009) has shown that the
presence of dispersed oil triggers a significant immune response, ultimately reducing immune
function in the Arctic scallop, Chlamys islandica.

C. islandica collected by divers from Porsanger, Norway, and exposed to sublethal dispersed
Ekofisk crude oil (low of 0.06 and high of 0.25 mg/L; 4 0.5C) in continuous flow for either
seven or 15 days, followed by seven days recovery in clean seawater had significantly altered
blood and cell biochemistry (Hannam et al., 2009). The sum of the PAHs during exposure were
about 0.509 and 2.55 g/L for low and high treatment tanks, respectively, consisting mostly of
naphthalene and its alkyl homologues. Body burden of summed PAHs was 3.36 and 5.70 g/g
for low and high treatments respectively. Bioaccumulated priority PAHs in the tissues included
fluoranthene, pyrene, benzo[a]anthracene, chrysene, benzo[b,j]fluoranthene and
benzo[b,j,k]fluoranthene. The concentrations of these in the experimental seawater were below
the detection limit, so bioconcentration factors could not be calculated. The bioconcentration
factors that could be calculated for individual priority PAHs ranged from 169 for naphthalene to
5430 for phenanthrene. Scallops clearly could bioaccumulate toxins from residual oil buried in
133

sediments since they tend to be slower at depuration than other bivalves (Hannam et al., 2009;
Wohlgeschaffen et al., 1992).

In the high treatment, haemocyte counts (16.9 10
6
/mL compared with 11.1 10
6
/mL for
controls) were significantly (p < 0.01) elevated and did not return to control levels over the
seven-day recovery period. Protein levels were significantly (p < 0.01) elevated to a maximum of
2.39 mg/mL in the low treatment after seven days. This level returned to control values (1.36-
1.46 mg/mL) during recovery. Hannam et al. (2009) suggest that the increase in haemocytes
might reflect a response to immune stress, but might also be an increase in blood cell production
triggered by the PAH-mediated lysis of lysosome enriched haemocytes. The latter mechanism
would explain the increase in plasma protein that was observed, although this protein increase
might also be a result of stress-induced blood enzyme synthesis.

There were significant reductions in cell membrane stability (p < 0.001) and phagocytosis (p <
0.01). Membrane stability did not show a drop until after 15 days of exposure, when stability in
the low treatment fell to 66% of controls and in the high treatment, 40%. Exposure time and
treatment were statistically significant factors. The scallops recovered during the depuration
period. Membrane stability is important for proper cellular function and is known to be disrupted
by PAHs. Consequently, the activity of the lysosomes in detoxifying xenobiotics can be
compromised. Membrane damage could also occur from ROS generated during metabolism of
xenobiotics. The ROS cause direct damage by lipid peroxidation.

Phagocytic activity was significantly altered by exposure time and treatment. After 15 days,
phagocytosis in the high treatment was 16.2 10
6
particles/mg protein compared to about 30
10
8
particles/mg protein for controls. This response also returned to normal after the seven-day
depuration period. Tests in vitro of the effects of PAHs on haemocyte and haemolymphatic
parameters in the Pacific oyster (Crassostrea gigas) using field concentrations (10
7
and
10
9
mg/mL) observed in the Marennes-Oleron Basin, France, and high concentrations (10
3
and
10
5
mg/mL) seen during oil spills have shown that fluorene, pyrene and heavy fuel oil
significantly decreased phagocytosis (Bado-Nilles et al., 2008). The percentage of lysosomes and
the percentage of phenoloxidase activity increased from dibenz[a,h]anthracene and
134

benzo[b]fluoranthene exposure, respectively. Modulation of immune parameters in the Pacific
oyster by PAHs suggested that PAH pollution may be related to enhanced susceptibility to
diseases.

Mean alkaline phosphatase activity ranged from 3.4-5.8 U/mg protein for all the animals and was
not significantly affected by exposure or treatment throughout the experiment (Hannam et al.,
2009). This enzyme works to breakdown foreign material. It was uncertain why these levels were
unaffected by dispersed oil.

Although all immunological changes in C. islandica normalized during recovery, the impact of
long-term continuous exposure remains unknown, but could result in increased susceptibility to
disease, and ultimately, increased mortality and decreased survival (Hannam et al., 2009).
Recovery from such damage is likely to be much slower in the Arctic due to the low growth
rates, higher generation turnover times and increased age at maturation that are typical of Arctic
organisms. The test concentrations had been chosen based on the OSPAR produced water
discharge limit of 30 mg/L oil, and represent 120-500 dilution of this limit.

C. islandica were exposed in the laboratory to weathered North Sea crude oil WAF for 21 days
at 8 1C in a simulated oil spill (Hannam et al., 2010). In this acute exposure, the maximum
sum of the 16 EPA priority PAHs and the alkyl homologues of naphthalene, chrysene,
dibenzothiophene, phenanthrene and anthracene was 163 g/L which decreased over the 21 days
to about 8% of the maximum. By day 21, 50 of the initial 80 exposed scallops had survived, but
there were no mortalities in the control group. The condition index, calculated as the percentage
of dry weight of tissue (g) divided by shell length (mm), was significantly depressed from days
2-14, then returned to normal. Immune function and oxidative stress were examined. Haemocyte
count significantly increased upon initial exposure, then fell below controls by day four, and by
day seven rose again to become indistinguishable from control values. The rebound was likely
due to the proliferation of these cells after PAH concentrations continued to decline. It is clear
that haemocyte count varies with exposure time and concentration (Hannam et al., 2010;
Hannam et al., 2009).

135

Phagocytosis initially increased, then became inhibited by day two and afterward to 41-57% of
control values. The inhibition may have been due to the large amount of energy required for this
activity, which cannot be continuously supplied for an extended period, as well as PAHs causing
the membrane to become less flexible and therefore less efficient at phagocytosis (Hannam et al.,
2010). There was a decrease in cytotoxic ability, measured as the lytic ability of haemocytes, but
it was not statistically significant. Cell membrane stability was significantly impaired. Under
oxidative stress, there was a significant reduction in glutathione levels, probably due to an
overwhelming metabolic production of ROS. The drop in glutathione was accompanied by lipid
peroxidation and consequently membrane instability. Following the exposure, immune function
did not recover, showing that these scallops can suffer sublethal effects from oil exposure even
though acute effects might be brief.

Hutcheson and Harris (1982) studied the blunt gaper (M. truncata) and the northern astarte clam
(Astarte borealis). M. truncata were exposed to 10, 25 and 50% WAF of Venezuelan Lago
Medio crude oil (2.1 mg/L total hydrocarbons, 96 h). A. borealis were exposed to WAF of the
weathered crude at 25% (0.95 mg TPH/L) and 50% (1.8 mg TPH/L) concentrations for 96 hours.
The energy that could be allocated to growth and reproduction once maintenance metabolic
needs are met (scope for growth) was determined for the clams after 96 hours and again 14
days later after depuration in clean seawater. Scope for growth has been used as a response
indicator and also as a monitoring tool after oil spills in temperate regions.

Scope for growth in M. truncata decreased (deficit in growth energy) in all treatments relative to
controls after exposure. The negative energy balance was a result of respiration exceeding
assimilation. The 10% WAF group had the greatest deficit in growth energy, indicating that the
decrease in scope for growth was independent of dose. Increased respiratory demands accounted
for the decreased scope for growth, and the clams did not recover to normal metabolic levels
even after 14 days of depuration.

For A. borealis, after exposure to the weathered crude WAF, there was no marked change in
scope for growth at the 25% concentration; however, at 50% (1.8 mg/L), the scope for growth
decreased by 220%. CEWAF of the weathered oil was also prepared using Corexit 9527 at a
136

1:10 DOR. CEWAF did not cause a significant change in the scope for growth at a nominal oil
concentration of 0.5 mg/L, but weakened energy balance was noted at 1.6 mg/L, and there was a
57% reduction in scope for growth in 3 mg/L. WAF (1.8 mg/L) was more toxic than CEWAF (3
mg/L). Some recovery of energy balance in the group exposed to 3 mg/L CEWAF occurred after
the 14-day depuration period, but it was insufficient for a positive scope for growth.

It was concluded that the reductions in scope for growth in this experiment were the result of
slightly depressed assimilation rates coupled with increased respiration rates, and WAF was
more toxic to A. borealis than CEWAF (Hutcheson and Harris, 1982). Arctic marine
invertebrates common in the nearshore benthic ecosystem have developed various adaptations
and traits to cope in the unique Arctic environment. Bivalves have adapted reproductive
strategies that lead to maximum survival of offspring, slow growth rates and long life spans, plus
metabolic adaptations that will encourage energy production under low thermal regimes. These
important traits help Arctic species survive the challenges of a cold environment.

Brandvik et al. (2006b) report that water taken from beneath an in situ burn was found to be non-
toxic to bivalve larvae and juvenile fin fish. In addition, the WAF from weathered Alberta sweet
mixed blend, and the WAF from its ISB residue during the Newfoundland Offshore Burn
Experiment was found to be non-toxic to three-spine stickleback (Gasterosteus aculeatus) and
the gametes of the white sea urchin Lytechinus pictus (Blenkinsopp et al., 1996). The fish tests
were for 96 hours with lethality as the end-point. The urchin test was of 20 minute duration and
the end-point was inhibition of fertilization. GC/MS analysis showed low levels of volatile
hydrocarbons with a maximum concentration of 1.1 g/mL.


Mammals

Several species of seals and whales are found in the Arctic Ocean including harp seals (Phoca
groenlandica) that consume fish living just under the ice; bowhead whales (Balaena mysticetus)
that filter small zooplankton through their baleen; ringed seals (Phoca hispida) that are ice
obligate and eat amphipods, mysids, euphausids, shrimp, polar cod, and small fish like smelt,
herring and capelin; narwhals (Monodon monoceros), beluga whales (Delphinapterus leucas),
137

ribbon seals (Phoca fasciata) and hooded seals (Cystophora cristata) that all hunt larger fish,
including polar cod, and squid; spotted seals (Phoca largha) and bearded seals (Erignathus
barbatus ice obligate) that feed on benthic fish and invertebrates; and walrus (Odobenus
rosmarus) and gray whales (Eschrichtius robustus) that feed extensively on benthic invertebrates
(infaunal and epifaunal). The effects of oil on these mammals can occur through direct contact,
ingestion and inhalation.

Although grey seals (Halichoerus grypus) are not native to the Arctic, the effects of an oil spill
on nursing mothers and newborn pups might provide some insight into effects of oiling in
general upon seals. Oil stranded onshore of Pembrokeshire, Wales, in 1974 occurred at the time
of grey seal pupping, and 25 of 62 pups were oiled (Davis and Anderson, 1976). Adults were
also oiled but the number could not be accurately counted. Oiled mothers transferred the oil from
their fur to the young during nursing. The percentage of mortality of both oiled and unoiled pups
from the contaminated beach was high in comparison to other unoiled beaches. Two of the oiled
pups were unable to swim, and drowned when washed off the beach. Behaviour of the seals
appeared unaffected by oiling (Davis and Anderson, 1976); however, the peak weight of oiled
pups was significantly lower than unoiled (t = 3.07, p < 0.01). Normal pups ingested little or no
oil during the 14 to 17 day lactation period.

Ingestion of oil has been the cause of death of grey and harbour (or common, Phoca vitulina)
seals (Griffiths et al., 1987). The northern fur seal (Callorhinus ursinus) is an eared seal found
along the north Pacific Ocean, the Bering Sea and the Sea of Okhotsk, and oiling of this animal
is expected to have serious consequences (Griffiths et al., 1987) since even a small amount of
oiling of the fur of sub-adults caused a significant increase in conductance from 26 to as much as
53 W/m
2
/C (Kooyman et al., 1977). Increased conductance translates to decreased thermal
resistance (insulation). Oiling of a pups fur caused conductance to increase from 40 to 54
W/m
2
/C (Kooyman et al., 1977). Live animals tend to groom, and this behaviour in a live, oiled
animal might cause the increase in conductance to be less.

Griffiths et al. (1987) described possible effects of spilled oil to various Arctic pinnipeds.
Harbour seal (P. vitulina) pups swim within a few days of birth but nurse for almost a month and
138

could become fouled, putting them at risk of physical impediment and drowning. Lactating
mothers could become fouled during their daily excursions for food and likely transfer oil to the
fur of pups during suckling (as was noted for grey seals). Depletion of crustacea and
invertebrates could threaten young first-year seals. Oil spill countermeasure activities could
cause the seals to abandon haul-out sites. Harp seals risk becoming fouled and younger seals that
feed in the upper water column could suffer from reduced prey. Hooded seals are migratory and
will probably not be affected much by a spill. Their pups do not have lanugo (fine hair) at birth,
and gain blubber rapidly so the reduced thermal resistance caused by oiling would likely not be a
factor. Bearded seal pups are born with adult fur, not lanugo, so oil fouling would be a problem,
but they do gain blubber rapidly so that thermal resistance would probably not be compromised.
There is little likelihood that adult bearded seals would become severely oiled as they frequent
open pack ice; however, if high concentrations of dispersed oil are encountered down to the 50 m
diving limit of this species it could pose a threat to their survival.

Ringed seals are the smallest and most numerous of the Arctic phocids, and are important to
Arctic peoples as a source for food as well as income. They inhabit fast-ice (ice that is fastened
to the coastline) during the late spring breeding season in areas where there is little ice
movement. This makes their risk of oil fouling small, so long as the oil does not migrate
appreciably under the fast-ice. If a spill were to occur within the fast-ice, there is a high risk of
the adults transporting oil on their fur to the lair and transferring it to the pups. The risk would be
most severe during the pupping season while the pups have their lanugo which usually lasts for
4-6 weeks after birth (Kelly et al., 2010). Some breed on permanent pack ice which adults
regularly occupy during late summer through winter to early spring. Oil in drifting pack ice
would cause fouling of adults.

Immersion of six ringed seals in a 1 cm thick slick of Norman Wells crude for 24 h at about 8C
followed by removal to clean seawater caused eye problems and minor kidney and possibly liver
lesions, but no permanent damage was observed (Engelhardt et al., 1977). Ringed seals suffered
eye irritation, lesions, severe conjunctivitis, swollen nictitating membranes and evidence of
corneal erosions and ulcers (Geraci and Smith, 1976). Eye damage in seals might be a common
result of oil exposure. Analysis for hydrocarbons after two and six or seven days of exposure
139

showed 1-11 g/g in the blood, 1-4 g/g in blubber, 1-2 g/g in brain tissue, 2-5 g/g in the
kidney, 1-8 g/g in the liver, 1 g/g in lung samples and 1-14 g/g in skeletal muscle. Urine and
bile had the highest concentrations (39 and 58 g/g respectively) two days after oiling, and these
were the major excretory pathways for oil metabolites. Oil in the brain correlated with
neurological symptoms early in the exposure.

Three ringed seals died within 7l minutes that were transported to the University of Guelph and
had oil added to their pool (Geraci and Smith, 1976). In this case, haematologic and blood
chemistry analyses indicated that the cause of death was the stress of captivity exacerbated by oil
exposure.

In an ingestion study, five ringed seals were fed over five days with 5mL/d of Norman Wells
crude oil as 1 mL capsules inserted into their fish food (Engelhardt et al., 1977; Geraci and
Smith, 1976). Blood samples for chemistry and haematology were taken prior to the ingestion of
oil as well as four weeks after. Three of the four animals had appreciably higher levels of
creatine phosphokinase after ingestion. This was the only measured blood parameter that
changed. Another ingestion study using radiolabelled
3
H-benzene in the oil capsules fed at the
same rate was used to trace metabolism of the ingested oil in blood and plasma up to 23 days
after feeding (Engelhardt et al., 1977). A peak of about 22-39 disintegrations per minute (dpm)
activity in the blood occurred at day 2 which rapidly declined to about 6-12 dpm by day 7.
Plasma activity peaked at about 15-32 dpm on day 4, declining to 4-10 dpm by day 7. All
blubber, liver and muscle tissues showed elevated activity (0.65-9.77 dpm) on day 2, declining
by day 28.

A high level ingestion dose was carried out on fasting two to three-week old harp seal pups in
two groups of seven (Geraci and Smith, 1976). Each group consisted of one control and six
experimental animals. Blood was drawn for chemistry and hematology prior to feeding. One
group was fed a single dose of 75 mL Norman Wells crude oil, and the other a single dose of 25
mL. Blood was drawn from the animals every two days over the ten-day period prior to
sacrificing. Oil was noted on the anus and hind flippers 1.5 hours after feeding. The only
difference noted in the oiled animals was a period of six to eight hours of restlessness
140

immediately after being given the dosage and one animal that had received a 75 mL dose was at
first unresponsive to manipulation. Specimens in the high-dose administration showed a sorbital
dehydrogenase peak of high activity 48 hours after feeding, reflecting transient liver damage.
The study demonstrated that a single dose of up to 75 mL of Norman Wells crude oil (which is
similar to Beaufort Sea oil being light, highly volatile, and of low viscosity) does not cause
irreversible damage to harp seal pups. Oil effects on the liver and kidney of seals could be
significant in the field where animals would be exposed for several weeks (Griffiths et al., 1987).

Walrus are gregarious, so groups of them would become fouled by a spill on pack ice or coastal
haul-outs. Any spill affecting benthic invertebrates will affect local walrus as these constitute
their prey. Walrus skin is very thick and is a good insulator with a conductance value of only 15
W/m
2
/C (Kooyman et al., 1977), so it would likely not be affected by oiling.

Griffiths et al. (1987) cite a publication entitled, Report of the Task Force - Operation Oil
(Clean-up of the Arrow oil spill in Chedabucto Bay) to the Minister of Transport, Volume II
that documented seals appearing to show considerable pain and suffering from contacting the
oil. The primary mode of death for these seals seemed to be suffocation. Griffiths et al. (1987)
note that if the animals were indeed displaying a pain reaction to the oil, it would likely be from
oiling of the eyes by the volatile components of the fresh crude. Other health risks to seals
include skin irritation, disorientation, lethargy, conjunctivitis, corneal ulcers, and liver legions
(Kelly et al., 2010).

Engelhardt (1983) reviewed the toxicity responses to petroleum shown by whales, seals, sea
otters (Enhydra lutris) and polar bears (Ursus maritimus). Ringed seals birth and nurse in
subnivean lairs which might pose a particularly high risk of oil exposure, in addition to their use
of ice pockets for breathing on long excursions beneath ice. Both situations increase the risk of
oil inhalation and coating. Seals and sea otters do not generally avoid oil, and so become coated.
Grey and harbour seal deaths have been associated with oil coating from the Kurdistan spill
(Parsons et al., 1980). In an accidental spill of Bunker C and No. 2 fuel oil from Regal Sword,
humpback whales (Megaptera novaeangliae), fin whales (Balaenoptera physalus) and white-
141

sided dolphins (Lagenorhynchus acutus) were observed feeding at and below the surface,
apparently unperturbed by the slick (Goodale et al., 1981).

A brief review of petroleum hydrocarbon effects on blood chemistry in mammals is provided by
ritsland et al. (1981). By way of example, they report that naphthalene inhibits cation transport
and thereby can effect a change in electrolytic balance. Increases in bilirubin (free or
conjugated), hypercholesterolemia, hyperuremia and elevated serum creatinine can result from
hydrocarbon toxicity. Changes in blood serum enzymes can be indicative of tissue lesions,
especially of the liver which is often the organ of focus in toxicity studies. Benzene can disrupt
leucocyte formation, as evidenced by upset of leucocytic enzymes including alkaline
phosphatase, and is known to cause haematopoietic problems. It generally causes the release of
immature erythrocytes destined for haemolysis either by inhibition of haeme synthesis,
influencing the production of fetal haemoglobin, and Heinz bodies. Naphthalene poisoning is
also evident by the formation of Heinz bodies and haemolysis. The resultant peripheral
haemolysis can drive the haematopoietic system to release more immature or less viable
erythrocytes which again will be destroyed and released as bilirubin.

General effects of an oil spill on cetaceans and points to consider given by Griffiths et al. (1987)
are as follows. Right whales migrate through ice by way of leads. They habitually skim the water
surface for food which makes them susceptible to oil ingestion. Sei whales (Balaenoptera
borealis) are a warm water migratory species that stay clear of pack ice, but feed similarly to
right whales and therefore run a similar risk. Humpbacks, on the other hand, feed below the
surface, and do not feed during migration, reducing the risk of oil ingestion. When migrating
through the Barents Sea, these whales follow the coastline which would expose them to external
oil contamination if a spill were to occur there. Blue whales (B. musculus) are found in the
Barents Sea to Svalbard. They feed primarily on krill, so any contamination of krill has potential
to contaminate feeding blue whales. Fin (B. physalus) and minke (B. acutorostrata) whales are at
low risk of suffering effects from an oil spill either to themselves or to their prey, as they feed
subsurface on both krill and fish, unless, as stated previously, the krill themselves are
contaminated. Sperm whales (Physeter catodon) feed at great depths, and those found in the
Arctic are usually solitary males. They are at low risk of effects from oil spills. Other species at
142

low risk from spills are the killer (Orcinus orca), northern bottlenose (Hyperoodon ampullatus)
and beluga (Delphinapterus leucas) whales, the common porpoise (Phocaena phocaena), and the
Pacific white sided (or white beaked) dolphin (Lagenorhynchus acutus).

Gray whales forage in the Chukchi and western Beaufort Seas, with occasional records in the
Canadian Beaufort Sea (Logerwell and Baker, 2011). It is not known whether gray whales are
able to detect surface oil from a spill, and in fact, during the Exxon Valdez spill, they were
observed to swim through the slick (Moore and Clarke, 2002). There was also one observation of
a whale lying in the surface oil, for unknown reasons. Gray whales probably experience irritation
to their eyes and tactile hair follicles from oil contact, but there is no evidence to show that
effects on skin tissue are long lasting (Geraci, 1990). It is more likely that oil might aggravate
broken areas of skin, or damage areas with barnacle growth, leading to bacterial infection
(Moore and Clarke, 2002); however, experimental skin exposure of a stranded sperm whale
(Physeter catodon) to gasoline (Geraci and Aubin., 1982) for 17 hours caused erosion and severe
damage likely because it removed natural oils. Experimental lesions purposely contaminated
with crude oil and petrol did not delay normal healing. Sperm whale skin biopsies that were
treated with 0600 M of -naphthoflavone, a known CYP1A inducer, were tested for levels of
CYP1A induction in the endothelium, in smooth muscle, and in fibroblasts (Godard et al., 2004).
Significant (p = 0.05) concentration-dependent increases in CYP1A staining in response to -
naphthoflavone treatment were found in cells of the endothelium and smooth muscle. This
inducibility of the CYP1A system in cetaceans suggests that they may be able to transfer,
metabolize and detoxify hydrocarbons.

The epidermis of cetaceans acts as a barrier against molecular antigens, and lacks Langerhans
cells which might otherwise cause oil to elicit an immune response (Geraci and St. Aubin, 1985).
This might explain why gray whales were observed to swim through oil slicks apparently
unperturbed. It is unlikely that cetaceans would be exposed to fresh, unweathered, crude oil for
extended periods. The number of mortalities due to a spill would likely be low, while the number
of individuals suffering significant irritation might be high (Griffiths et al., 1987).

143

Although oil can adsorb to the baleen, it can be removed with the continual flushing of seawater
(Geraci and St. Aubin, 1985). Heavy oils like Bunker C were found to clog the baleen of fin
(Balaenoptera physalus), sei (Balaenoptera borealis) and humpback (Megaptera novaeangliae)
but not gray whales, and reduce water flow, which could conceivably interfere with normal
feeding. At 15C, heavier oils could persist on the fibers of fin and gray whales for 15 to 20
hours, while lighter oils were cleared from the baleen plate within one hour. Geraci and St Aubin
(1985) concluded that the level of fouling tested would not normally impair flow rate or feeding.

Whereas the risk of oil contamination through ingestion of benthic prey contaminated with
dispersed oil has been suggested (Neff, 1990), there has been no evidence to support this (Geraci,
1990). However, Geraci and St. Aubin (1985) found low levels (<7 g/g wet weight) of
naphthalene in mysticetes (fin, sei and minke) and pinnipeds (Phoca vitulina, H. grypus, E.
barbatus, 0. rosmarus). The highest levels were found in beluga (6.3-21.1 g/g) and narwhal
blubber (10.5-17.9 g/g). Bowhead whales and bearded seals that are harvested for subsistence
on the Alaskan North Slope have not been found to have PAHs (OHara and OShea, 2005). This
lack of detectable PAHs underscores that it is necessary to establish biochemical and histological
indicators in marine mammals, since they do not bioaccumulate these chemicals, unlike
polychlorinated biphenyls and other organochlorines.

In an unpublished study (Fraker, 1984), gray whales seemingly respired more rapidly on
surfacing, and surfaced less frequently while passing through a region off the California coast
that has natural subsea oil seeps, but the differences in behaviour were slight and not statistically
evaluated. Whales could inhale toxic vapour from a fresh spill, which has more potential to cause
harm to respiratory membranes than any vapour from oil that has weathered for two to four
hours. However, the impact would likely only affect individual animals that have been weakened
by parasites or adrenal dysfunction (Fraker, 1984).

Bowhead whales could be at high risk of oil ingestion if they were to feed on contaminated
zooplankton. It has been found that food organisms of this whale can accumulate hydrocarbons if
they reside in an oil contaminated environment (Malins, 1977). There appears to be little or no
144

possibility of a whale aspirating ingested oil that becomes regurgitated because of an anatomical
adaptation of the larynx peculiar to cetaceans that prevents aspiration (Geraci and Aubin., 1982).

It is unknown what the effects are of prolonged inhalation of crude oil vapours by marine
mammals. This is a knowledge gap that might be significant considering the migration of
cetaceans to the Arctic during the spring and summer months. In terms of exposure, the
reduction in sea ice might lead to increased range of migrant cetaceans such as fin, humpback,
minke (Balaenoptera acutorostrata), gray and killer (Orcinus orca) whales which now can be
found in the Barents and Bering seas (Moore and Huntington, 2008).

In addition to effects mentioned above, exposure of marine mammals to oil can result in oiling of
eyes and skin, and although they are considered to be less affected from fouling by oil as
compared with other animals (e.g. seabirds), a spill in a location where mammals serendipitously
occur, whether for migration, feeding, breeding, or otherwise, has the potential to affect a
population (stock) or isolated group (Huntington, 2009).

Residual oil from the Exxon Valdez has been implicated as the cause for high mortality rates
among sea otters (Enhydra lutris kenyoni) in Prince William Sound; the otters showing elevated
levels of CYP1A that suggest continued low-level oil exposure (Monson et al., 2000). The
European sea otter subspecies (Enhydra lutris lutris) is found on the Kamchatka Peninsula and
Commander Islands of Russia, adjacent to the Barents Sea, and as with the kenyoni subspecies,
they are on the International Union for Conservation of Nature red list as endangered (IUCN
Otter Specialist Group, 2010). A small population of E. l. lutris was reported in Finmark,
Norway (Griffiths et al., 1987). Not much is known of the biology of this otter which feeds on
fish, clams, crabs and urchins along the beaches and up to about 100 m offshore (Griffiths et al.,
1987). The otter does not have blubber but relies on its thick fur for insulation (IUCN Otter
Specialist Group, 2010), making it vulnerable to fouling with oil which is a leading cause of
death for oiled individuals (Griffiths et al., 1987). There is also the danger of harmful effects
through ingestion of contaminated food. The sea otter can be used as a sentinel species for
habitat health monitoring (IUCN Otter Specialist Group, 2010).

145

Harris et al. (2011) analyzed blood samples from 29 live-captured sea otters (E. lutris) in British
Columbia, as well as their invertebrate prey, for alkanes and PAHs. Hydrocarbon concentrations
in sea otters were similar between areas and among age and sex classes. Biomagnification factors
(ratio of otter to prey) suggested that metabolism plays a significant role in bioaccumulation.
Most PAHs did not biomagnify, but some higher alkylated 3-ring and 4-ring PAHs did, even
when concentrations of PAHs in dietary prey were low, as a consequence of the otters extreme
consumption rates (up to 25% of their body weight per day). The estimated body burdens for the
sum of PAHs was 5120 3600 g (alkyl PAHs = 89 7%) for males, and 3710 2560 g (alkyl
PAHs = 84 10%) for females.

Oiling with Prudhoe Bay crude of the pelts of adult sea otter (E. lutris) caused only a slight
increase in conductance from 22-26 W/m
2
/C to 26-29 W/m
2
/C; however, for a pups pelt,
conductance increased from 7 to 15 W/m
2
/C (Kooyman et al., 1977). It should be noted that if
an oiled animal were to groom its pelt, the increase in conductance might be less severe. Also,
Prudhoe Bay is a light crude oil that did not cause the fur to clump. Costa and Kooyman (1982)
found that an average of 17% oiling of the fur of live E. lutris with Prudhoe Bay crude increased
metabolic oxygen consumption by 41% from an average of 16.0 mL O
2
/kg/min at 15C prior to
oiling to 22.2 mL O
2
/kg/min. Curiously, after washing, metabolism increased 106% to 32.6 mL
O
2
/kg/min. Two of the experimental animals experienced shivering, one of which subsequently
died of pneumonia, which suggests that should these animals be cleaned of oil following a spill
(very expensive considering the animals would have to be housed at 20C for about a week to
groom), they risk dying of complications including pneumonia and hypothermia (Griffiths et al.,
1987). The authors feel that grooming is necessary to maintain the furs water repellence and
insulation. Oiling caused the fur to clump resulting in the loss of the air layer that is usually
trapped between fur and skin and prevents water penetration. It is not known if the oiling and
washing procedure had irritated the skin to cause increased peripheral circulation and heat loss.

Kooyman (1977) found that conductance of a bearded seal (E. barbatus) pelt was unaffected by
oiling with Prudhoe Bay crude. Mammals that depend on their fur for insulation rather than
blubber tend to increase their metabolism in response to oiling. The oil causes an increase in
conductance of the pelt and corresponding cooling of the skin. Although the peripheral
146

temperature decreases, there is a compensatory increase in metabolism that allows the animal to
maintain its core temperature (Engelhardt, 1983).

Polar bear (Ursus maritimus) are considered an ice-obligate species, spending most of their time
on the ice where they hunt, rest, and bear young (Moore and Huntington, 2008). They are a top
predator of the Arctic Ocean food web consuming mostly ringed and bearded seals, but they will
also eat belugas, narwhals, walrus, harbour seals, reindeer (Rangifer tarandus), and birds (Laidre
et al., 2008). For the Svalbard polar bear population it was estimated in 1983 that a take of two
males per female at a rate of 54 animals per year, would not endanger the population (Ritsland
and Schweinsburg, 1983). This could apply to a reduction of the population due to an oil spill,
but the calculation must be reiterated using current population statistics.

Oil readily adheres to the fur of polar bears (Ritsland et al., 1981). An experimental exposure
(by enticing bears with seal blubber to enter a seawater pool) of three polar bears for 15-50 min
to a 1 cm thick slick of Midale crude oil (Ritsland et al., 1981) led to the appearance of
petroleum in most tissues due mainly due to ingestion while grooming, followed by skin
absorption and inhalation. The animals avoided swimming in the oiled pool. The two bears that
had been exposed for 30 and 50 minutes died one month later. Liquid samples of whole blood,
plasma, bile, urine and vomitus were obtained, in addition to tissue samples of liver, kidney,
skeletal and heart muscle, brain, spleen, subcutaneous fat, mesenteric fat, marrow and lung. The
fur of one bear was sampled. The amount of oil calculated to have adsorbed to the fur was 9-27
kg, and a bear crossing an oiled lead would be expected to have its fur rapidly and heavily
contaminated. Physical removal of the oil by the bears was mostly limited to grooming. All bears
began shivering within eight hours of oil exposure, and all vomited more than once within 12
hours of oil ingestion. Separate pelts that were tested showed an increase by factors of two to
four in conductance (decreased insulative capacity) following oiling. Slight increases in skin
temperature were noted and could not be specifically attributed to the direct oil coating or to the
ingested oil, but were likely a compensatory response to the loss of insulative capacity. Solar
heating of the skin due to oiling increased by 28%. The first 12 hours during an oil spill would be
critical to dealing with contaminated polar bears.

147

Immediately after exposure, oil was present in the vomitus at a concentration of 520-605 mL/kg
and in feces at 46-90% by volume or weight. The main route for hydrocarbon excretion was by
way of feces. Urine samples had concentrations as high as 714 L/L, and 27-42 L/L three to
four weeks post exposure. Urine was another major excretory route of oil. Of the tissue samples,
the highest hydrocarbon levels were found in the bone marrow (471 L/kg), kidney (200 L/kg),
and brain (325 L/kg). Liver, lung, skeletal and heart muscle had decreasing levels in that order.
Oil was undetectable in the fat samples.

Haematocrit values declined from 45-56% to 14-30% at the end of the experiment. The maxima
in erythrocyte count decline was 6-8 10
6
/L. Values of other parameters including mean cell
volume, mean cell haemoglobin, and mean cell haemoglobin concentration also declined.
Leukocyte counts initially increased, but thereafter, changes were inconsistent. Serum sodium
decreased in two of the three bears, while calcium increased in all exposed bears. Levels of
lactate dehydrogenase increased after exposure. Total bilirubin increased from control levels of
0-0.8 mg/dL, rising as high as 4-19 mg/dL. Serum protein also rose, and cholesterol was high in
all three bears from the second to fifth week post-oiling. The normal level of blood urea nitrogen
was 4-33 mg/dL, which initially showed a slight decline and then rose to about 150 mg/dL near
the end of the experiment. Creatinine also increased from control values of 0.7-2.5 mg/dL to
about 13 mg/dL. In summary, the bears showed evidence of severe anemia from suppressed
haematopoiesis (from the myelotoxicity of benzene) and from haemolysis, due to the oiling,
although the white blood cell counts were not conclusively altered by the oil. Sodium and
phosphate showed definite imbalances attributable to oil exposure. All three bears showed
biochemical and pathological evidence (lesions of the gastrointestinal tract, nephrosis) of kidney
failure from oiling. Hepatotoxicity was minimal. ritsland et al. (1981) concluded that the polar
bear is a potentially greatly impacted species when exposed to oil spills. The surviving polar
bear was treated and nursed back to health over a period of several months. This would be a
great expense if many individuals were severely fouled by a major oil spill in the Arctic.
Cleaning oiled marine mammals could require weeks of humane confinement and the use of
antibiotics, tranquilizers and other drug therapy (Griffiths et al., 1987).

148

Any model predicting the consequences of oil fouling must account for the season, the
productivity of the area, and the variable health status of the target population (Geraci and Smith,
1976). The difficulty in assessing toxicosis in marine mammals lies in the fact that body burdens
of contaminants and declines in populations have rarely been linked directly to a source, and
especially not for petroleum hydrocarbons. The main reasons for this lack of evidence are
(OHara and OShea, 2005):
lack of understanding of marine mammal health and physiology;
reliance on extrapolation of historical, unrelated species data;
reliance on tissue residues;
ethical aspects of studying mammals;
experimental designs that lack statistical power;
correlations and associations of chemicals with effects rather than clear mechanisms;
lack of effective monitoring and funding.

The assessment of potential toxicity must be multidisciplinary, begin with measurements of
contaminants in the environment and organisms, and scale up to populations, species and
ecosystem levels (OHara and OShea, 2005). The measurements of chemicals should include
congener or isomer concentrations, oxidation states and metabolites. These will help to examine
questions of bioaccumulation, biomagnification, toxin kinetics and transformations at the
molecular, cellular and tissue levels in sickly individuals compared with healthy specimens
obtained through stranding, by-catch, hunted, or catch and release. This would, in turn, help to
evaluate the impact on, and epidemiology of individuals, populations, species and so forth.
Studies need to be long-term, and designed for rigorous statistical analysis. In this way, exposure
to chemicals could be linked to effects observed in mammals. Among the species suggested for
monitoring by the 1998 Marine Mammal Commission were these Arctic ones: harbor seals, grey
seals, beluga, bowhead and minke whales, sea otters, walruses, polar bears. Mitigation of
contaminant effects in the Arctic will require international collaboration and global management
in order to provide adequate protection of circumpolar biota (OHara and OShea, 2005).
Recommendations for marine mammal scientific research, policy, and supportive infrastructure
are given by OHara and OShea (2005).

149

7. Arctic Oil Spill Field Trials

Much of what we have learned about the behaviour of oil in Arctic and ice-infested waters has
been linked to a number of limited field experiments involving controlled releases of fresh and
emulsified crude oil for research purposes. The Canadian governments approval of exploratory
drilling operations in the Canadian Beaufort Sea in the 1970s motivated a number of programs
to study oil spill countermeasures and study the fate and effects of oil and spill response
technologies in offshore cold water and Arctic environments.

7.1 Balaena Bay experiment 1974 - 1975 (Norcor)

One of the first large-scale oil spill experiments occurred in 1974-75 in the Canadian Arctic in
Balaena Bay, a small bay 20 km to the southwest of Cape Parry (70 02 N, 124 53 W), NWT.
This study was conducted with the principal objective to assess the impact of an offshore oil well
blowout on the thermal regime of the Beaufort Sea, and to develop potential countermeasure
techniques. The Balaena Bay experiment was designed to generate fundamental data on the
interaction of crude oil with Arctic sea ice (NORCOR Engineering & Research Ltd, 1975).
These experiments were the first to study all aspects of oil in ice behaviour, including spreading
under ice, encapsulation, progressive vertical migration as the ice warmed, spreading on surface
melt pools in the spring, and weathering (Dickins, 2011). Oil spill countermeasures, namely in
situ burning (ISB) and mechanical cleanup (booms, skimmers, pumps, sorbents and beehive
burners) were conducted near the end of the experiment once the ice had melted. A smaller
component of the project was to measure the biological impact of the spills and possible
relevance to other areas. A follow-up chemical/biological study was conducted in 1981 to assess
the degree of environmental impact as a result of the experimental oil spills (Dickins and
Hellebust, 1981).

During the conduct of the study, two crude oils (Norman Wells and Swan Hills) were injected in
nine separate discharges under the ice inside large Fabrene containment skirts at various stages
of ice growth and depletion from October 1974 to May 1975. In addition to the contained spills,
two additional spills were carried out 30 km off shore where oil was allowed to spread freely in
150

the presence of a 10 cm/s current and movement documented by divers and underwater camera
footage. A total of 54.4 m
3
were spilled in eleven separate discharges (NORCOR Engineering &
Research Ltd, 1975). The test areas were continually monitored and oil, water and ice samples
were recovered for laboratory analysis.

In May of 1975, oil started to appear on the ice surface after migrating naturally through brine
channels within the sea ice. Most of the oil rose to the surface and lay in large pools and these
pools were successfully burned. Oil residue was shovelled off the ice until the ice began to
disintegrate. As the ice cover was broken, it grounded on the shore where the remaining residue,
mainly in the form of tarry clumps and balls was swept up (5 m
3
from the ice and 0.6 m
3
from
the shore). The bulk of the spilled oil (> 85%) was evaporated (10 m
3
) or burned (27 m
3
).

Oil that was experimentally released rose to the surface and upon striking the underside of the
ice, radiated outward, progressively filling depressions in the sheet. Within a matter of hours of
oil contacting the ice, it became encapsulated and once entrapped was stabilized with little
evidence of degradation or weathering through the winter months. With the onset of spring and
upon exposure to increased solar radiation and air temperatures, the oil began to migrate upwards
and at the surface saturated the snow cover, reducing albedo, hence accelerating the process.
Oiled areas were free of ice one to three weeks earlier than unoiled areas. This study gave
sufficient information on the behaviour of oil in stationary first-year ice, and on spill
countermeasures to be employed, but further work was needed on the entrainment and migration
of oil in second and multi-year ice, the effects on an oil slick on sea ice formation (particularly
under dynamic conditions), and long-term fate and impact of unrecovered oil. ISB was found to
be most efficient and effective with a minimum film thickness of 0.5 cm required to sustain
combustion. Manual recovery of the oil was costly and labour intensive. Detailed physical and
biological studies were conducted prior to the first discharge of oil. Salinity and temperature
profiling and dissolved hydrocarbon content were measured throughout the year.

Controlled spills of 40 50 L (20 L/m
2
) of fresh and weathered Norman Wells crude oil were
conducted on uncontaminated salt marsh areas on the northwest side of the Bay by Dr. J.
Hellebust (Department of Botany, University of Toronto). Observations of the sites six weeks
151

after the spill showed that vegetation was extensively killed by both fresh and weathered crude,
with the exception of scurvy weed (Cochlearia officinalis) which appeared surprisingly resistant
to the very high levels of crude oil applied. Only marsh plants above the water line at the time of
oiling were affected by the spill. Toxicity experiments were conducted with fresh and weathered
crude collected from the under-ice oil experiments with freshly collected phytoplankton and
zooplankton samples from Balaena Bay. Young stages of the jelly fish (Cyanea capillata) were
very sensitive to both fresh and weathered crude oil.

Results from the Balaena Bay study provided evidence of limited environmental damage from
and oil spill during the winter as the oil was effectively trapped in or under the ice. In most
cases, little effect was noted on phytoplankton and no effect on zooplankton and benthic
organisms. This was attributed to the very effective cleanup of the weathered crude oil before
the breakup of the ice. There were no significant impacts on zooplankton or phytoplankton
populations. The overall conclusion for the study was that there was no evidence of oil in the
water column or deleterious effects on the marine ecosystem with the exception of very light
oiling of the shingles of the tidal zones over about 900 m of shoreline. It was estimated that after
mechanical removal, burning, and physical degradation, only 1 m
3
of residual oil remained on
the site at the end of 1975.

Balaena Bay Revisited: 1981

The Balaena Bay site was revisited in 1981 to verify if the no environmental effects still
applied after six years (Dickins and Hellebust, 1981). The original biological and chemical data
collected were insufficient in quantity and definition to provide an adequate baseline, so an
independent study strategy was developed, involving comparison of benthic histopathology and
population between stations both within and outside the impacted areas in 1974. Field sampling
and analysis programs were designed to evaluate any long-term effects on both benthic and
seashore organisms and to provide data on any oil residues still present. Benthic, beach and salt
marsh sediments were sampled and analyzed for non-polar and polyaromatic hydrocarbons.
Histopathological studies were conducted on the clam Macoma calcerea and the polychaete
152

Pectinaria hyperborean, and the benthic community was assessed as well as vegetation in the
contaminated area.

After six years, weathered oil residues were found coating rocks in a narrow strip of intertidal
zone. Petrogenic hydrocarbons were still detectable in the beach sediments and at the salt marsh
spill site but not in the benthic or salt marsh control station sediments. The total volume
remaining was estimated at 0.2 m
3
with coverage ranging from 0.1 to 5% in the worst areas. The
heavily oiled salt marsh experimental sites showed grass recovery rates of less than 13% with
about 5% of the original concentration remaining in the upper soil layers. The oiled plots showed
a definite recovery of the higher dominant marsh grass, Puccinellia phryganoides, about 13% in
weathered crude oil plots, and 3.5% in fresh crude oil plots, while little recovery was found for
the subdominant Wilhelmsia physodes. A very active blue-green micro-algal flora developed in
both the oiled plots. No measurable effects were found on the moss and algal marsh zones which
were not directly covered, but closely adjacent within 30 cm of the applied oil. Approximately
5% of the applied crude oil remained in the salt marsh soil at the spill sites. Areas covered with 1
to 10 mm of crude oil residues showed little vegetation, however some algae, lichens and the
macrophyte, Cochlearia officianalis, were observed growing in close proximity to the oil
residues. Amphipods were also seen in close proximity with crude oil residues in salt marsh
depressions containing seawater. In conclusion, there appeared to be little or no toxic effects to
algal photosynthesis or marine invertebrate behaviour by the oil residues, which covered a
relatively small area of the shoreline near the spill sites. However, the residual oil appeared to be
unsuitable substrate for algal or higher plant growth with the exception of blue-green algae.

The only statistically significant difference in the health and condition of polychaetes and
bivalves from the oil spill control sites was the minimal reproductive development observed in
animals from the oil spill site, with the possible cause of this attributed to some natural and/or
anthropogenic stress in the environment. Differences were seen in the benthic communities
found at the control stations and the oil spill. Final conclusions were that overall impact of the
spill was small (Dickins and Hellebust, 1981).

153

7.2 Baffin Island Oil Spill Experiment (BIOS)

The BIOS conducted multidisciplinary experimental oil spill field studies over a four year period
(May 1980 and August 1983) in the Canadian Arctic at a field site located near Cape Hatt on a
small peninsula of northern Baffin Island, Northwest Territories (now Nunavut), Canada, 65 km
southwest of the community of Pond Inlet. The experimental design and overall results have
been summarized (Sergy and Blackall, 1987), and details of various components of the study can
be found as individual papers in a dedicated volume of the journal, Arctic (Volume 40,
Supplement 1). Information obtained from the study was intended for planning, approval and
operational phases of hydrocarbon development in the Arctic.

The primary objectives of the BIOS project were to determine if the use of chemical oil
dispersants in the Arctic nearshore 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 the
Arctic nearshore and shoreline areas (Sergy and Blackall, 1987).

There were two separate but complementary studies, each involving a controlled release of crude
oil. The nearshore study compared the consequences of dispersing an oil slick close to shore with
the option of allowing the oil to beach and leaving it to natural attenuation (self-cleaning). The
study compared the fate and effects of a short-term high concentration of dispersed oil with the
fate and effects of a long-term low level release of oil stranded on the shoreline. The shoreline
study evaluated the technological cleanup of beached oil. Findings on oil fate and persistence
were used to compare the relative effectiveness of promising Arctic shoreline cleanup techniques
against a benchmark of natural self-cleaning processes (Sergy and Blackall, 1987).

Three bays of similar coastal morphology and sedimentology were selected as test areas
(Semples, 1987). Studies on climate, ice conditions, geomorphology and oceanography were
conducted and document the physical features of the area and discuss them in relation to regional
conditions (Buckley et al., 1987; Dickins et al., 1987; Meeres, 1987; Semples, 1987).

154

In the first year of the study (1980), physical, biological and chemical baseline data were
collected and analyzed. Two experimental oil spills were conducted the second year (1981),
where one chemically dispersed oil spill would impact one bay and a second untreated oil spill
would impact a similar bay. Sampling was repeated in the third and fourth years to document the
long-term fate and effects of the spill. Baseline studies of petroleum residues in the sediments
(Cretney et al., 1987b) and waters (Cretney et al., 1987c) were measured and found to be in the
low to sub-microgram per gram and microgram per litre concentration range as might be found
anywhere on Earth. Prior to the spill of oil, Cretney et al. (1987a) also measured tissue
hydrocarbons in a variety of Arctic marine species and were found to be low (ng/g) with origins
from combustion products.

Nearshore Study

Two experimental discharges of 15 m
3
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 an oil/dispersant ratio of 10:1. The oil and dispersant
were selected based on a reasonable data base existing for both. The surface slick release began
at high tide and under influence of an onshore wind. Oil was carried up on the beach by wind and
wave action as the tide receded. The test area was boomed to prevent contamination at the
beginning of the experiment and remained in place for several weeks to capture any oil sheen
redistributed by the tides. Oil remaining on the surface was collected with skimmers after a full
tidal cycle. Oil was dispersed using a sub-sea diffuser pipe through which the
oil/seawater/dispersant was released. The chemically dispersed oil caused a variety of conditions.
The sub-sea discharge caused benthic organisms to be exposed to abnormally high
concentrations of toxic aromatic hydrocarbons. Oil was also introduced at greater water depths
than would likely be the case of a surface oil slick dispersed at the surface in the Arctic. In many
coastal Arctic areas, the often present and well-pronounced pycnocline would suppress the
downward mixing of oil to the productive benthos.

The beach could not retain all of the oil that reached the shoreline and as a result, one third of the
spilled oil was recovered in cleanup activities on the water, one third was evaporated and the
155

remaining was stranded in the intertidal zone. The stranded oil was subject to natural cleaning
processes during approximately six months of open water periods from 1981 to 1983. Over this
time, the surface oil cover was reduced by approximately half, whereas estimates indicated that
80% of the oil initially stranded (5.3 m
3
) was removed. It was concluded that oil was removed in
substantial quantities from the intertidal zone even in such a low-energy Arctic environment
(Owens et al., 1987a).

There was no major large-scale mortality of benthic infauna to either oil release. Multivariate
analyses showed no significant change in infaunal community structure, and effects attributable
were found in three of 72 univariate analyses of density, biomass or size data for individual taxa
(Cross et al., 1987a; Cross and Thomson, 1987). A progressive decrease in the condition of the
filter-feeding bivalve Serripes groenlandicus in the reference bay several kilometres away from
the dispersed oil release was the result of exposure to dilute dispersed oil for several days. A
similar effect on condition in the surface deposit-feeding bivalve Macoma calcaerea was
apparently caused by relatively low oil concentrations in the sediments of the dispersed and
surface oil release bays. Decreases in density of the polychaete Spio spp. indicated that oil in the
sediments of the surface oil release and dispersed oil release bays affected reproductive
processes. Bivalve and polychaete effects were still evident two years after the spill (Cross and
Martin, 1983).

The sub-surface release of the oil resulted in short-term relatively high oil concentrations in
waters of the two adjacent bays and oil in the water could be detected below 1 m in the untreated
oil release. Immediately after the spill, divers observed narcosis in urchins and starfish. Overall,
there were no major effects of oil release on densities of epibenthic crustaceans and amphipods.
Minor effects attributable to the untreated oil included depth distribution of Anonyx juveniles,
and dispersed oil in the water column apparently had a delayed adverse effect on reproduction in
the amphipod family, Stenothoidae (Cross et al., 1987a).

Biomass, number of species, and reproductive condition of the dominant understory macroalgae
at 3 m depth did not seem to be adversely affected by either oil in sub-tidal sediments or by
chemically dispersed oil in the water column. The lack of major effects on macroalgae may have
156

been partly attributable to the lack of effects on herbivores and the vegetative mode of
reproduction in the dominant macroalgal species (Cross, 1987).

Bacterial numbers and microheterotrophic activity (uptake of glutamic acid) were monitored in
the water column and sediments of selected bays between 1981 and 1983 prior to and after the
spill (Bunch, 1987; Bunch, 1984; Bunch et al., 1983). Dissolved organic carbon and inorganic
nutrients were measured in the water column. Total organic carbon was measured in the
sediments. At the concentrations of oil released into the water by the surface and dispersed oil
releases, the effect on measured bacterial activity was transient and minimal. No changes in
bacterial activity in the sediments could be ascribed to either oil release. Bacterial numbers were
not affected by the oil releases. The Corexit 9527 significantly altered kinetic parameters in
laboratory experiments, but was not seen in field conditions where the concentration of
dispersant was three orders of magnitude lower. Oil-degrading capacity was determined in the
water and sediments of all bays and did not appear to be altered by the oil releases. The effects of
the surface slick or the dispersion into the water column would be inconsequential or marginally
deleterious to bacterial numbers or the microheterotrophic uptake of glutamic acid (Bunch,
1987).

The natural biological degradation of oil was compared with microbial action enhanced by the
artificial application of nutrients in three oiled plots in the Arctic backshore sediments
(Eimhjellen and Josefson, 1984). The effects of the addition of artificial fertilizer were positive
but varied depending on the geomorphological conditions of the beach. In fine sediments there
was a 1-100 times increased level of oil degrading bacteria relative to unfertilized controls over a
two-year period. Biodegradation rates were enhanced three to five times for the alkane fraction
of the oil. The effects of fertilizer addition on coarse sediment were marginal. Mechanical
mixing of oil and fertilizer into the sediment gave further improvement in biological degradation
in spite of lower levels of oil-degrading bacteria. Sergy and Blackall (1987) noted that the
biological impacts were relatively minor. All short-term effects were temporary and apparently
without serious consequence. There were no measured biological effects in the intertidal zone
due to the absence of Arctic intertidal life. The intertidal zones in more temperate regions are
more vulnerable to the effects of an oil spill.
157


The BIOS study was successful in achieving conditions to determine oil fate and effects that
would be relevant to future Arctic oil spills. The countermeasure options, chemical dispersion
and natural shoreline restoration were assessed and compared in an intertidal and subtidal habitat
common in the eastern and high Arctic. Biological impacts from both spills were relatively
minor and effects were short-term. After two years of post-spill monitoring, there was no
evidence of large-scale mortality of subtidal benthic biota attributable to either the chemically
dispersed oil or the oil contaminated beach. Infauna, epifauna and microalgae populations and
community structure showed few changes (Cross et al., 1987a; Cross and Thomson, 1987; Cross
et al., 1987b).

Shoreline Study

The shoreline study evaluated options for enhancing the cleanup of oiled shorelines by
comparing the methods to natural self-cleaning. The shoreline study required the use of several
sections of coastline with different beach materials and exposure to wave and tide action. A large
embayment on the east coast of Cape Hatt (Z lagoon) was selected as the site for the shoreline
countermeasure experiments. The location was sufficiently distant from the other BIOS
experimental oil releases to preclude the possibility of cross contamination. The study sites
provided a representative range of high, intermediate and low-energy wave environments, and
energy levels at the shoreline are directly related to local fetch characteristics (Owens and
Robson, 1987). Small plots (20 - 40 m
2
) were used in the intertidal and backshore zones of the
test beaches. Two control plots each 40 m
2
were laid down in 1980 in the upper intertidal zone.
Consistency was maintained in the type, quantity and method of application of the oil.
Emulsified and non-emulsified Lagomedio crude oil of the same stock used in the nearshore
study were applied in paired plots. After application of oil to the beach, the plots were left for 24
hours prior to the initiation of cleanup tests to simulate minimal response time.

The intertidal plots were oiled to their maximum holding capacity, as evidenced by the migration
of surplus oil down the beach face. Wave energy was the dominant factor in removal of oil from
exposed beaches (99% in 48 hours). On a partially exposed beach, the majority of oil was
158

removed within the first open water season (40 days) and by the end of the second season, less
than 0.05% remained. On the sheltered beaches, the rising tides removed large quantities of oil
during the first two days after which relatively stable conditions prevailed. Oil was less persistent
on the fine-grained sediments than on the pebble-cobble beaches. The back shore plots were not
affected by marine processes and climate and soil conditions played a major role in the fate of
the oil.

Weathering and biodegradation altered the composition of the oil. The beach cleaning methods
evaluated are summarized in Owens et al. (1987b). Techniques evaluated were based on the
operational realities of the eastern Canadian Arctic. 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 requirement. The methods chosen were in situ combustion using
an incendiary device, 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.

The natural fate and persistence of oil was examined to determine the rates and physical factors
influencing the self cleaning process (Owens and Robson, 1987). Neither the incendiary device
nor the low-pressure flushing techniques proved to be effective, whereas over the short period of
testing (5-6 weeks), mixing and chemical dispersion demonstrated a potential to mitigate the
effects of beach contamination or accelerate the removal of the stranded oil (Owens et al.,
1987a). Both produced an immediate reduction in the quantity of oil on the beach surface.
Chemical solidification was effective in stabilizing the oil but very labour intensive. The BIOS
project results provided no major ecological reasons to prohibit the use of chemical dispersants
on oil slicks in nearshore areas. Despite unusually severe exposure to chemically dispersed oil
the impact on a typical shallow-water benthic habitat was not of major ecological consequence.

Chemical dispersants may be the only alternative in situations where the immediate protection of
shoreline and nearshore habitats is of primary importance or where a shoreline cleanup operation
is environmentally less desirable. There was no strong ecological reason for the cleanup of oil
159

stranded on Arctic shorelines, except where wildlife is present or their critical habitat is
threatened, or areas of human use. Oil residues do persist for long periods on low-energy beaches
and backshore areas and in near-by sea-bed sediments. Arctic beaches can be cleaned by oil by
natural processes, especially in low priority shorelines, such as Cape Hatt. Response can be
directed towards areas of greater importance and sensitivity, such as shores adjacent to
communities, wildlife breeding and staging areas, and traditional hunting and fishing camps.
Dispersant washing and mechanical mixing may be desirable countermeasures in areas where
wildlife frequent the shoreline to reduce contact with the oil. Dispersant washing may also
prevent the formation of oil-sediment consolidation (asphalt pavement).

BIOS Revisited

Following the experimental application of oil in August of 1981, the BIOS site was revisited and
field studies were conducted in subsequent years (1982, 1983, 1985, 1987, 1989, 1993, 2001) in
experimental beach, Bay 11, which was left to weather naturally. Systematic documentation of
changes in the distribution, quantity and character of the oil through time has been done (Owens,
1994; Owens et al., 2002; Prince et al., 2002). In 1985, sediment samples and benthic biota were
collected in two of the BIOS study bays, one used for surface oil release and the other the
reference site (Cross and Humphrey, 1987). Hydrocarbons were still present in subtidal sediment
samples four years after the oil releases but the oil content in sediments appeared to have
stabilized between 1983 and 1985. Hydrocarbons remaining were predominantly biogenic in all
samples from the reference bay and in most samples from the oiled bay. Weathering indices
indicated that any oil present in both bays was highly weathered. Low levels of oil remained in 3
benthic species with concentrations higher in the oiled bay than the reference bay. Oil
concentration increased between 1983 and 1985 in tissues of the bivalve Serripes groenlandicus
from the surface oil release bay and in tissues of the bivalve Macoma calcarea and sea urchin
Strongylocentrotus droebachiensis from the reference bay. Both biogenic and highly weather
petrogenic hydrocarbons were detected using gas chromatographic analyses. Benthic organisms
appeared to be taking up more polar weathered residues present in the water column or
remaining in the sediments. There was no evidence of oil-related effects on mean size data or
size frequency distributions in three bivalve species.
160


Owens (1994), reported in 1989 an approximately 80% decrease in the total oiled area. Prince et
al. (2002), revisited the site of the BIOS in August of 2001, twenty years after the original
experiment was conducted to report on the chemical composition of the oil. The vast majority of
the original oil was gone, but small patches remained with an estimated <0.5% of the original
coverage area. This is a low-energy level beach, and gross physical loss of oil by heavy wave
action was an unlikely mechanism to explain the gradual loss of oil. Oil-mineral fines interaction
could explain the loss (Owens, 1994).

Samples collected by Prince et al. (2002) remained essentially unaltered despite 20 years of
exposure to the elements while others showed that photo-oxidation and biodegradation played an
important role in removing the majority of the components of the oil. Prince et al. (2002) used
the conserved biomarker 17(H)21(H)hopane (Prince et al., 1994) to show in a sample
collected from the intertidal zone that more than 87% of the initial hydrocarbons present were
biodegraded, including a significant proportion of initial chrysene and alkylated congeners.
Other samples from the surface of the intertidal zone were extensively depleted of chrysenes
suggesting photo-oxidative losses. While some samples were extensively biodegraded, others
remained essentially unaltered. Samples above high tide were the least biodegraded. This was
attributed to the lack of nutrients for microbial growth as their position on the beach had them
rarely inundated with water.

Prince et al. (2002) concluded that biodegradation can be a major fate of spilled crude oil on an
Arctic beach, even though biological processes are probably restricted to an average of only 63
ice-free days per year (Dickins, 1987). Biodegradation can still occur at low temperatures by
indigenous microorganisms (Eriksson et al., 2001) and a more important limitation may instead
be the availability of the necessary trace elements for microbial growth. This in turn suggests that
a bioremediation strategy that would stimulate oil biodegradation, as has been shown in
Spitsbergen (Prince et al., 1999). The fate of the residual oil that does not leave the shorelines by
physical transport will eventually be biodegraded, but this will continue to be severely limited by
the availability of trace nutrients required for microbial growth.

161

7.3 Field Trials in Svalbard, Norway

Early Field Studies on Oil Bioremediation

In 1976, one of the first shore-line field trials to assess the feasibility of bioremediation as an oil
spill countermeasure was conducted in Spitsbergen (Svalbard), Norway using inorganic
fertilizers as the bioremediation agent (Sendstad, 1980; Sendstad et al., 1984). Unweathered
Forcados crude (10 L/m
2
) was spread on two test sites and the oil was treated with a
commercially available fertilizer at an application rate of 1.2 % (wt/vol). The fertilizer
stimulated oil biodegradation with a 3 fold increase in microbial respiration rate until 1979. By
the end of 1983, the oil on the unfertilized control plot biodegraded to a degree similar to the
fertilized plot.

In Situ Treatment of Oiled Sediment Shorelines (ITOSS) Program

The ITOSS program was initiated in Svalbard, Norway in 1997 at a site located located on the
Van Mijenfjord, approximately 40 km from the open ocean at approximately 77 56 N and 16
43 E (Sergy, 1999; Sergy et al., 1998). While the purpose of this study and the BIOS program
was to better understand in situ shoreline cleanup options, many of the results are applicable to
the natural fates of oil on shorelines in general. A series of experimental oil spill studies were
conducted on a number of shoreline locations to to quantify the effectiveness of selected in situ
shoreline treatment methods to accelerate the natural processes that remove oil stranded on
mixed sediment beaches and to investigate the processes that naturally remove stranded oil, such
as wave-induced mechanical abrasion and the interaction between oil and fine mineral particles
(Owens et al., 2003).

The experimental design of the field trials is described by Gunette et al. (2003). At each of three
distinct shoreline sites (exposed to different energy levels), treatment test plots and control plots
were established within a continuous stretch of oiled shoreline ranging from 40 to 143 m long. A
total of 5500 L of intermediate fuel oil (IFO 30) was deposited along a 3 m wide swath in the
upper intertidal zone of each site, the location where oil would be expected to strand in the event
162

of a spill. Approximately one week after oiling, a different treatment technique was applied to
each plot. The treatment techniques included sediment relocation (surf washing), mixing (tilling),
bioremediation (fertilizer application) and bioremediation combined with mixing. One plot at
each site was monitored for natural attenuation. The quantity of oil removed from the plots was
measured six times up to 60 days post-treatment and then again one year later. Changes in the
physical character of the beach, oil penetration movement of oil to the subtidal environment,
toxicity and biodegradation were monitored over the 400 day period.

Sediment relocation of the oiled shoreline significantly accelerated the rate of oil removal
(Owens et al., 2003) and reduced oil persistence by at least one year in the sheltered wave energy
environment where oil was stranded on the beach in the upper intertidal and supratidal zones,
above the level of normal wave activity. Sediment relocation accelerated the short-term rate
(weeks) of oil loss from the intertidal sediments on the relatively low wave-energy shoreline,
where the stranded oil was located in the zone of wave action.

Oil-mineral aggregation (OMA) formation played a significant role on the effectiveness of the in
situ shoreline treatment options of natural attenuation (natural recovery) and sediment relocation
(surf washing). At both sand and pebble beach sites, the amount of oil remaining in the
experimental plots was dramatically reduced within five days after sediment relocation
treatments. Time-series microscopy and image analysis of breaker zone water samples
demonstrated that OMA formation occurred naturally on the oiled beaches at both sites and was
accelerated by the sediment relocation procedure. Oil in OMA does not adhere to surfaces and is
thus more easily removed from mixed sediments. OMA have considerably more buoyancy than
mineral particles and can therefore be transported much further before settling to the bottom. The
increased surface/volume ratio facilitates weathering processes, in particular biodegradation (Lee
et al., 2003). Procedures developed in this experiment to support the use of sediment relocation
as an operational spill response option and a guidance document were developed to identify the
extent of OMA formation (Stoffyn-Egli and Lee, 2002). No toxic effects were detectable in
adjacent coastal waters and sediments at either experimental site as a result of these inherent
dispersion processes and of documented oil biodegradation (Lee et al., 2003).

163

Mixing and tilling did not clearly contribute to removal of oil within the intertidal sediments
(Owens et al., 2003). The changes in total oil loading caused by treatment were small and
unlikely to have accelerated natural recovery. Results suggested that mixing/tilling made the
shoreline sediment more permeable to seawater and air for at least ten days after tilling, which
could lead to enhance microbial activity (Prince et al., 1999).

Oil biodegradation was observed in the intertidal sediments (Prince et al., 2003a) and increased
by the application of soluble slow-release fertilizers. Fertilizer application to the surface of the
beach delivered nutrients to the oiled sediment beneath the beach surface. Fertilizer application
was followed by increased oxygen consumption, increased carbon dioxide evolution from the
beach, increased microbial biomass, and significantly greater biodegradation of oil on the plots
that had received fertilizer. Changes in the chemical composition of the oil demonstrated that the
biodegradation rate was doubled.

The rate of removal by natural processes of oil stranded in the active intertidal zones in the initial
ten days after oiling was rapid. After ten days, the removal rate slowed dramatically and oil
residue levels did not significantly change for some months. Four to five percent of the original
oil still remained in the active intertidal zone after one year (Owens et al., 2003). Oil placed in
the supratidal zone penetrated deeply into the coarse grain pebble sediments. About 30% of the
oil was lost in the first five days, but no further reduction occurred over the next year and
significant quantities remained, despite the relatively high local wave energy on this beach.

The in situ treatment applied during the Svalbard field trails had a distinct and verifiable effect
on the rate of oil removal from intertidal sediments. Oil persistence was reduced and secondary
effects appeared minimal as sediment location did not elevate toxicity in the nearshore
environment to unacceptable levels nor did the treatment methods result in significant alongshore
or offshore sediment oiling. Sediment relocation accelerated the removal of oil from the
shoreline through the formation of OMA (Owens et al., 2003). Sediment relocation can be used:
when sediment is stranded high on a beach above the zone of normal wave activity; when oil has
penetrated deeply into beach sediments below the zone of normal sediment reworking; when
natural attenuation would take an unacceptable length of time; when bulk oil has been removed
164

and cleaning or polishing of the remaining residue is desirable (high amenity beaches); prior to
storm events to expose oiled sediments; when in situ techniques are appropriate because
sediment removal is undesirable due to a lack of natural replenishment, or because waste transfer
or disposal is an issue; or where there are logistical constraints in remote or inaccessible
locations.

7.4 Field Trial Projects on Arctic Oil Spills in Ice

Dickins (2011) has provided an excellent overview of the history of research into the behaviour
of oil spills in ice covered waters and summarized what has been learned about the behaviour of
oil in ice from actual field experiments involving deliberate crude oil spills for research purposes
(Table 7).

As noted by Dickins (2011), experimental oil spills in ice began in the early 1970s and involved
fresh and emulsified crude oil spilled under solid ice. The first controlled oil spill in pack ice was
conducted by the contractors SL Ross & DF Dickins off the coast of Cape Breton in 1986. Crude
oil was spilled in pack ice to investigate the physical and chemical fate of spilled oil in the tested
ice conditions. This was followed by a number of focused studies on specific topics as described
below.

The most recent and certainly the most extensive Arctic oil spill research program has been the
SINTEF Oil in Ice JIP carried out between 2006 and 2009, with the field trials component
having been conducted in the Norwegian Barents Sea between 2008 and 2009
(www.sintef.no/Projectweb/JIP-Oil-In-Ice/). This JIP covered a number of subjects of interest to
industry, stakeholders and regulators.
Large-scale field experiments in the Barents Sea
Oil distribution and bioavailability
Fate and behaviour of oil spills in ice
ISB of oil spills in ice
Mechanical recovery of oil spills in ice
165

Use of dispersants and chemical herders on oil spills in ice
Oil spill response guide

Table 7 Experimental crude oil spills of a few barrels to hundreds of barrels conducted in sea ice, regardless
of latitude (Dickins, 2011).
Project Year Reference Location Environment Spills Size Cleanup
Behavior of oil
spills in the
Arctic
1970
Glaeser and
Vance, 1971
Chukchi
on and under
fast ice
1-2 bbl/5 spills burning
Crude Oil
Behavior on
Arctic Winter Ice
1971 McMinn, 1972 U.S. Arctic on fast ice
3 spills on
snow
not known
Interaction of
Crude Oil with
Arctic Sea Ice
1974/75 Norcor, 1975
Canada
Beaufort
under fast ice
340 bbl/9
spills
burning,
mechanical
Oil Behavior
Under MY Ice
1978-82
Comfort et al.,
1983
Canada High
Arctic
under old ice
11 bbl/ single
spill
none
Oil and Gas
Under Sea Ice
1979/80
Dickins and
Buist, 1981
Canada
Beaufort
under fast ice
116 bbl/3
spills
burning
Oil Migration in
Solid Sea Ice
1979/80
Nelson and
Allen, 1982
U.S. Beaufort under fast ice
18 spills of 1.5
to 18 gal each
mechanical
Interaction of
Crude Oil with
Solid First-year
Ice
1980/81
Nelson and
Allen, 1982
U.S. Beaufort on fast ice
3 spills
sprayed onto
snow 6 bbl
each
burning,
mechanical
Emulsions in Ice 1982 Buist et al., 1983
Canada
Beaufort
under fast ice
100 gal/2
spills
burning
Experimental
Spills of Crude
Oil in Pack Ice
1986
Buist and
Dickins, 1987
Canada East
Coast (Nova
Scotia)
between
floes and in
leads in pack
ice
18 bbl/3 spills burning
Marginal Ice
Zone
Experiment
1993
Singsaas et al.
1994; Vefsnmo
and
Johannessen,
1994
Barents Sea
75N
between
floes and in
leads in pack
ice
164 bbl none
Svalbard
Experimental
Spill
2006
Brandvik et al,
2006; Dickins et
al., 2008a
Svalbard under fast ice 21 bbl burning
Oil in Ice Field
Experiment 08
2008
Sorstrom et al.,
2010
Barents Sea
78N
in openings
within pack
ice
5 bbl/2 spills
burning,
herders
Oil in Ice Field
Experiment 09
2009
Sorstrom et al.,
2010
Barents Sea
78N
between
floes in close
pack ice
110 bbl/5
main spills
burning,
dispersants


The SINTEF field trials also demonstrated the importance of being able to validate small-scale
tank tests and meso-scale experiments with large-scale field trials and allowed researchers to
166

document the fate and behavior of spilled oil over time, albeit a relatively short time frame of
approximately two years (Brandvik and Faksness, 2009).

Trials to test the differences of spreading oil in open water and ice demonstrated that in most
situations, the presence of ice slowed the rate of spreading and oil was contained in a small area
(Table 8). A batch crude oil spill measuring 1600 m
3
(10,000 bbl) was carried out and final
average thickness and final area of the spilled oil was very different between open water and ice-
covered conditions. 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. These findings have implications for
response times and decision making during recovery. Oil contained in ice and snow will be
thicker and more easily burned or otherwise recovered.

Table 8 Spreading comparison for a 1600 m
3
(10,000 bbl) crude oil spill (SL Ross Environmental Research
Ltd. et al., 2010).
On Smooth Ice
Open Water Under Solid Mid-Winter Ice
Ice Snow
Final average oil
thickness (mm)
0.016 40 to 90
a
3 40
Final area (ha) 10,000 7 to 70
b
50 4
a
Maximum pool depth under ice depends on depth of under-ice depressions, which grow deeper as ice grows in winter
b
The range reflects the variable processes involved: final contaminated area depends on the available volume of under-ice
depressions and how they fill with oil


The conduct of large-scale field experiments has overwhelmingly demonstrated how natural
containment, reduced wave action and relatively slow weathering in the Arctic environment can
aid in spill response scenarios due in large part to the presence of ice and cold water. The
findings of these unique aspects of oil fate and behaviour following its release in the Arctic
environment extend the window of opportunity for mobilizing response activities and can
enhance the effectiveness of certain response measures such as ISB and skimming. However, the
unique Arctic environment poses challenges for traditional mechanical cleanup methods, further
reason to increase field trials and develop appropriate technologies.


167

Encouraging results have come from experimental oil spills that support further research and
probable exploration 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 (Dickins et al., 2000; Singsaas et al., 1994; Vefsnmo and Johannessen,
1994).
Ignitability is enhanced in cold temperatures and ice due to the greater persistence of
lighter and more volatile components of petroleum hydrocarbons(Srstrm et al., 2010).
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 around the world, thus
allowing for easier marine operations (Dickins, 2011).
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 (Dickins,
2011; Srstrm et al., 2010).
High ice concentrations (7/10 or more) tend to immobilize and encapsulate most spilled
oil quite rapidly, particularly from a subsea blowout (Dickins, 2011).
Ice encapsulated oil is effectively isolated from weathering (evaporation, dispersion,
emulsification) and allows for a longer window of opportunity to carry out cleanup
activities. Effective combustion at a later date is possible as the oil remains unweathered.
Most of the Arctic shoreline has a fringe of fast ice which acts as an effective natural
barrier against oil contamination on the coastline (Dickins, 2011).

Some response challenges and limitations have also been identified in the field trials.
Oil trapped on or under ice in moving pack ice is difficult to assess as crews cannot
maintain continuous operations with immediate means of evacuation.
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.
Crude oils with pour points of 0C or less tend to gel.

168


7.5 Field Trials on Enhanced Oil Dispersion with Mineral Fines

Field tests were conducted by the Canadian Coast Guard (Lee et al., 2011c) in the St. Lawrence
Estuary in the winter of 2008 to assess the feasibility of an oil spill response technique in ice-
infested waters based on the application of fine minerals in a slurry with mixing by propeller-
wash to promote the formation of OMA. This process promotes the physical dispersion of
mineral-fine stabilized oil droplets into the water column which are subsequently diluted below
toxicity threshold limits and biodegraded by natural bacteria. Following the release of the test
crude oil and the application of experimental treatments, time-series changes in oil
concentrations were monitored to quantify dispersion effectiveness. Field samples were also
recovered for laboratory microcosm studies on the biodegradation of petroleum hydrocarbons, by
monitoring O
2
consumption and CO
2
production (Micro-Oxymax Respirometer), and depletion
of specific hydrocarbon components. The GC/MS data shows that with hopane normalization,
more than 60% total petroleum hydrocarbon, 75 to 88% of total alkanes, and 55 to 65% TPAHs
were degraded after 50 days incubation at low temperature (0.5
o
C). The alkylated PAH was
degraded to a greater extent in the presence of clay minerals than in their absence. These field
and laboratory studies suggest that the proposed remediation strategy can be an environmentally
acceptable and cost-effective means to combat oil spill in ice-infested waters. In terms of the
Arctic, this technique offers several operational advantages including reduced numbers of
personnel required for its application, the lack of need for waste disposal sites, and cost
effectiveness.

7.6 Monitoring Arctic Offshore Oil and Gas Operations

Monitoring of the hydrocarbon impact of the Northstar and Liberty oil prospects began in 1999
as the Arctic Nearshore Impact Monitoring in the Development Area program, under the
auspices of the Minerals Management Service (now the Bureau of Ocean Energy Management,
Regulation and Enforcement), and continues as part of a long-term program to assess potential
spatial and temporal changes related to oil development and production near their associated
sites (Brown et al., 2011). Annual field surveys of sediment chemistry in the nearshore Beaufort
169

Sea examined PAHs, saturated hydrocarbons, steranes and triterpanes, metals, grain size, total
organic carbon, and radionuclides for age-dating. In general, the results indicate that no
significant contaminant inputs from Northstar development activities occurred, and that any
observed changes were within the bounds of natural variability for the study area. The
monitoring, analytical and interpretive methods are believed to be sufficiently sensitive to detect
changes that might occur from new input sources.

7.7 Future Oil Spill Field Trials in the Arctic

In light of the recent Deepwater Horizon blowout, fears expressed by the public of an accidental
oil spill in the Arctic environment may temper future exploration and development initiatives.
While some interest groups claim that industry has little or no experience in responding
effectively to Arctic oil spills, during the last two decades there have been a number of new oil
spill countermeasure technologies developed and tested in the laboratory. Some techniques such
as ISB that were used in the Gulf of Mexico are now being accepted as proven procedures for
operational use. There now is a need to have validated oil spill response tools for use in the
Arctic to address the requirements of contingency plans and to support oil spill responders. Field
trials on Arctic oil spill countermeasures are required. For environmental relevance, efficacy
studies on oil spill countermeasures cannot be conducted in the laboratory and mesocosm-scale
test facilities alone.

There is a need for government regulators to support the scientific communitys request to
undertake large-scale controlled oil spills in the Arctic to support the development and validation
of strategies for the protection of the marine environment that will be used in future contingency
plans for Arctic offshore drilling operations. The operational effectiveness of numerous new oil
spill cleanup technologies and response strategies remains largely unknown. Proof-of-concept
and operational guidelines must be established before they are accepted by the oil spill response
community. Typically, the only opportunity for oil spill countermeasure techniques to prove
their capabilities is during an actual accidental spill (spill of opportunity). Unfortunately, when
such events occur, the main focus of the activities is to recover the oil or prevent it from harming
wildlife or sensitive environmental areas and generally not on proving the effectiveness of the
170

different techniques employed. To complicate the situation further, regulators have been very
reluctant to accept the use of alternative countermeasure techniques, other than mechanical
recovery due to the lack of evidence for their performance under field conditions.

The results of the SINTEF JIP field experiments in 2008 and 2009 have recently been very well
received by a wide range of stakeholders, with these exceptions: the fact that only a limited
number of stakeholders witnessed the work, only one sea ice scenario was included, and only a
single location was covered. The scientific and oil response community have agreed that the
conduct of large-scale tank tests and field experiments is essential for improving our knowledge
and expertise in Arctic oil spill countermeasures.

Field testing for operational improvements and field demonstrations (with equipment
deployment) both have a role in illustrating the commitment of the oil and gas industry and
government regulators to insure the availability of an effective response capability in the Arctic.
To ensure better cost efficiency, equipment practice and demonstrations could be conducted in
conjunction with scientific experiments. Due to the global need for this information by nations
with Arctic interests and the costs involved for these studies in the Arctic environment, a study
of this magnitude should be cost-shared and managed under a collaborative international
program. To optimize the use of resources, collaborative multidisciplinary research needs to be
promoted. For example, tests on mechanical recovery with skimmers could follow field tests.
Validation of tracking and surveillance trials including remote sensing technologies can be
coordinated with the release of oil and subsequent oil behaviour tests.

Arctic field trials will provide an experimental framework to evaluate a number of issues of
concern, including these:
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/Arctic conditions;
development and verification of oil-in-ice drift and fate models;
improved mechanical response such as skimmers, pumping systems for viscous oils,
and the removal of oil from ice and water;
171

ISB in broken ice (development of fire booms, chemical herding agents, and ignition
systems);
oil dispersion enhancement by chemical dispersants/enhanced OMA formation (focus
on effectiveness, identification of controlling factors and development of application
guidelines;
bioremediation of oil stranded on shorelines;
characterization of water soluble components and biological effects (toxicity) on
Arctic species;
development of operational end-points for spill cleanup operations.

Results from recent field trials provide proof that carefully planned and executed experimental
spills can be carried out safely with no harm to the environment (Dickins, 2011; Lee et al.,
2011c; Srstrm et al., 2010). Unless methodologies can be validated in the field, they will not
be fully accepted by the oil responders, regulators and the public as operational tools. The
conduct of an international program to undertake field trials to support oil spill countermeasure
research fits in with the Government of Canadas Northern Strategy Framework, promotes
northern science and research, reinforces circumpolar cooperation, strengthens partnerships, and
encourages environmental protection. Potential participants include Canada, the United States,
Norway, France, Denmark, Sweden and Russia.

In terms of a location for field trials in the Arctic, there is a need to coordinate the conduct of
field trials in Canadian Arctic waters as the public does not fully accept data from other regions
of the world due to the perception that environmental conditions and sensitivity of indigenous
biota are different between geographical regions. This would require the coordination and
conduct of controlled oil spill release experiments in Canadian waters to validate emerging
techniques for oil spill response in Arctic marine waters with and without ice cover.

The Canadian Arctic is a unique environment, with hydrology, ecosystems and ice behaviour
which each pose challenges to understanding the behaviour of pollutants in the high North. The
presence of ice, the unique physical environments, and the fragility of the ecosystems reduce the
applicability of our knowledge of oil behaviour in temperate climates.
172


There is a considerable corpus of experimental data, but many of the results are anecdotal or
empirical in nature. As such, particularly for the historical trials of the earlier decades, it is not
easy to abstract these observations into behavioural models. This is changing with some of the
more systematic recent studies, but there remain notably large gaps (Khelifa, 2010) between the
research that has been accomplished and a basic understanding of the physical processes at work.
Empirical relations are often useful rules of thumb, but they have limited extensibility and
applicability to conditions other than those of the original experiment.

Current modelling capabilities for oil in ice are approximately two decades old and are typically
derived from open water models with factors included to account for the presence of ice. Basic
ice physics and behaviour are generally not captured in the analytical models.

Understanding of the weathering processes experienced by oil in the Arctic environment is
incomplete. Certain processes such as evaporation, biodegradation and emulsification, are
strongly affected by temperature. Many oils also have non-linear transition points (pour points)
in the range of Arctic temperatures, which have dramatic effects on oil behaviour, as seen in the
work by Buist et al. (2009).


173

8. Operational Waste Discharges

Two potential sources of petroleum hydrocarbons in the Arctic marine environment from
offshore oil and gas activities include produced water and drilling muds and fluids. A
comprehensive review on the potential impacts of Canadian east coast offshore oil and gas
exploration and production activities, including the discharge of operational wastes, has recently
been released by Fisheries and Oceans Canada (Lee et al., 2011a).

8.1 Produced Water

Produced water is the largest volume waste stream generated by offshore oil and gas activities. It
consists of formation water from the oil-bearing substrata brought to the surface with the oil and
gas, along with seawater injected into the reservoir to maintain wellhead pressure during
production, and in some cases condensed water (Ayers and Parker, 2001; Patin, 1999). Produced
water is usually hot saline water that contains a mixture of both organic and inorganic
compounds that are naturally present in the geologic formation, as well as treatment chemicals
used during drilling, production, stimulation and oil-water separation processes (Hawboldt
2010).

In 2003, an estimated 667 million metric tons of produced water were discharged offshore
throughout the world, including 21.1 million tons to offshore waters of North America, mostly
the U.S. Gulf of Mexico, and 358-419 million tons to offshore waters of Europe, mostly the
North Sea (OGP, 2004; OGP, 2005). These are underestimates of actual discharges, because
reporting of production to OGP (2004) ranged from 11 percent to 99 percent in the seven regions
of the world monitored. In terms of discharge capacity, it is generally reported that the total
amount of water produced from gas fields is much smaller than from oil production fields. While
many gas fields discharge less than 10 m
3
of produced water per day, most oil fields discharge
hundreds or even thousands of cubic metres of produced water per day (OGP, 2002).

174

Considerable concern has recently been raised over the disposal of produced water from
production operations as its release is continuous, in comparison to the episodic release of
contaminants associated with the disposal of muds and fluids during the drilling of exploration
and production wells.

The principal alternative to ocean discharge of treated produced water is underground injection.
However, the feasibility of this practice at offshore installations is dependent on a number of
site-specific factors including access to a suitable disposal formation, chemical interactions that
may result in precipitates that can plug the receiving formation, and cost. Furthermore, the net
environmental benefit of reinjection must be considered. On the basis of energy requirements, it
is estimated that 2.6-4.3g of CO
2
emissions are produced for each litre of produced water
reinjected into a sub-surface well (Shaw et al., 1999).

Chemical Composition

Produced water is a complex mixture of organic and inorganic chemicals in both dissolved and
particulate phases (Table 9). The physical and chemical properties of produced water vary
widely depending on the age, depth, and geochemistry of the hydrocarbon-bearing formation as
well as the chemical composition of the oil and gas phases in the reservoir. Furthermore, since no
two produced waters are alike, region specific studies may need to be conducted to address the
environmental risks from its discharge.

In Table 9, many of the high values reported for metals may be anomalous due to matrix
interferences from high concentrations of dissolved salts in produced water (Neff, 1987). The
highest values are extremely rare and concentrations of most constituents in most produced
waters, measured by modern, accurate analytical methods, fall into the lower part of the range.
175

Table 9 Concentration ranges (mg/L or parts per million) of several classes of naturally-occurring metals and
organic chemicals in produced water world-wide (Neff, 2002).
Organic Chemicals Concentration Range
Total organic carbon (0.1 >11,000
Total saturated hydrocarbons 17 30
Total benzene, toluene, ethylbenzene, and xylenes (BTEX) 0.068 578
Total polycyclic aromatic hydrocarbons (PAH) 0.04 3.0
Total steranes/triterpanes 0.14 0.175
Total phenols (primarily C0-C5-phenols) 0.4 23
Total organic acids 0.001 10,000
Inorganic Chemicals
Salinity (mostly sodium and chloride) <2000 - > 300,000
Sulfate 1.0 8,000
Sulfide 0 - 140
Nitrate 0.6 15.8
Ammonia 14 246
Orthophosphate 0.1 6.6
Arsenic 0.000004 0.32
Barium 0.001 2,000
Cadmium 0.0000005 0.49
Chromium 0.000001 0.39
Copper 0.000001 55
Iron 0.0001 465
Lead 0.000001 18
Manganese 0.0002 7.0
Mercury 0.000001 0.075
Nickel 0.000001 1.67
Zinc 0.000005 200
Total radium (pCi/L) 0 5,150


Petroleum Hydrocarbons

Petroleum hydrocarbons in both dissolved and dispersed (oil droplets) form are the organic
components of greatest environmental concern in produced water. Current regulatory guidelines
for produced water discharge in Canada are based on petroleum hydrocarbon content. In Canada,
the hydrocarbon content of produced water must be reduced to acceptable levels under the 2002
Offshore Waste Treatment Guidelines prior to discharge into the ocean. The minimum regulated
standard for the treatment and disposal of wastes associated with the routine operations of
offshore drilling and production installations in Canada is 30 mg/L as a 30-day weighted average
of oil in the discharge, coupled with a 24-hour arithmetic average not to exceed 60 mg/L.

176

Existing oil-water separators such as hydrocyclones are quite efficient in removing the oil
droplets within produced water. Thus, the petroleum hydrocarbons discharged to the ocean are
typically low molecular weight aromatic hydrocarbons and smaller amounts of saturated
hydrocarbons that are found within the dissolved phase of the produced water. Since there are no
cleanup procedures that are 100% effective, treated produced water still contains some dispersed
oil as droplets ranging in size from 1 to 10 m (Johnsen et al., 2004).

The most abundant hydrocarbons in produced water are the one-ring aromatic hydrocarbons,
BTEX and low molecular weight saturated hydrocarbons. BTEX may be present in produced
water from different sources at concentrations as high as 600 mg/L (Table 9). Benzene often is
the most abundant BTEX compound in produced water, followed by toluene. Because BTEX is
extremely volatile, it is lost rapidly during produced water treatment by air stripping and during
initial mixing of the plume in the ocean (Terrens and Tait, 1996). PAHs are the petroleum
hydrocarbons of greatest environmental concern in produced water because of their toxicity and
persistence in the marine environment (Neff, 1987; Neff, 2002). Concentrations of TPAH in
produced water typically range from about 0.04 to 3.0 mg/L and mainly consist of 2 and 3-ring
PAHs such as naphthalene, phenanthrene, and their alkyl homologues. Higher molecular weight,
4 through 6-ring PAHs rarely are detected in properly treated produced water. When present,
they are associated primarily with dispersed oil droplets (Johnsen et al., 2004).

Environmental Concerns over Discharges

Hawboldt et al. (2010) has recently raised concern over the potential impacts of produced water
discharges in the Arctic. While the quantity of oil discharged may be small, due to the
sensitivity of the Arctic environment there is concern that local impacts could have farther
reaching effects, especially with expansion of oil and gas development in the future. The harsh
conditions of the Arctic environment result in specific challenges when dealing with the
discharge of oil through produced water effluents. Sea water temperatures in the Arctic range
from -1 to 1C in the winter to 5 to 10C in the summer. The three most important mechanisms
when looking at the discharge of petroleum hydrocarbons and other produced water
contaminants are dilution, biodegradation and evaporation. However, in cold water both
177

biodegradation and evaporation rates are significantly reduced. Slower natural contaminant
reduction rates increase the probability of exposure and thus toxic effects. In addition,
contaminants that would normally dissolve in warmer waters may exist in a dispersed or less
soluble state in colder Arctic waters, further reducing their biodegradation and vaporization rates.
The discharge of hot produced water into cold receiving water may also result in a plume which
rises to the surface forming a sheen. Sheen formation is also more likely in the Arctic since oil in
produced water tends to stay in the dispersed phase rather than the dissolved phase due to
reduced solubility. The presence of a surface sheen may result in physical impacts on marine
organisms such as gill smothering in fish as well as oiling of seabirds and marine mammals.

The presence of ice can affect the fate of oil released into the environment in a number of ways.
Ice floating at the surface traps and accumulates dispersed oil. Ice prevents oil from mixing and
being diluted and thus oil may concentrate in one location, intensifying local impacts. Ice also
reduces the amount of open surface water and thus vaporization rates are reduced. There are a
number of sympagic species that may be at risk from the discharge of produced water, such as
fish that lay their eggs directly under the sea ice. These eggs or fish larvae can be harmed by
exposure to dispersed oil or by the lack of food such as plankton.

Due to the extreme climatic conditions in the Arctic, there are relatively few species resulting in
low biodiversity. Since food chains are relatively simple, an input of hydrocarbons which
directly affects one species may eventually impact the entire food chain. Migration of animals
into the Arctic region are a significant factor that determines the vulnerability of Arctic marine
ecosystems. There are a number of species that migrate to the Arctic for wintering, breeding,
feeding and molting. The consequence of the congregation of animals means that a small amount
of oil in a specific region may have significant ecological consequences. In addition, the long
seasonal changes in the Arctic results in a reduced period of productivity in the spring and
summer, which increases the risk of impacts from oily discharge. Reduced sunlight during the
Arctic winter could affect the removal of contaminants from the system since photo-oxidation
rates would be reduced.

178

Fate following Discharge into the Ocean

Treated produced waters from offshore platforms are typically discharged above or below the sea
surface once regulatory compliance concentrations are achieved. The location of subsurface
discharge pipes may range in depth from 10 to 100 m. Saline produced waters are usually as
dense, or more dense, than seawater and disperse below the sea surface, diluting rapidly upon
discharge into well-mixed marine waters. Low salinity produced water may form a plume on the
sea surface and dilute more slowly (Nedwed et al., 2004). Dispersion modelling studies of the
fate of produced water differ in specific details, but all predict a rapid initial dilution of discharge
by 30 to 100-fold within the first few tens of metres of the outfall. This is followed by a slower
rate of dilution at greater distances (Brandsma and Smith, 1996; Smith et al., 2004; Strmgren et
al., 1995; Terrens and Tait, 1993). Factors that affect the rate of dilution of produced water
include discharge rate and height above or below the sea surface, ambient current speed,
turbulent mixing regime, water column stratification, water depth, difference in density (as
determined by temperature and total dissolved solids concentration), and difference in chemical
composition between the produced water and ambient seawater.

Environmental Effects of Discharges

Based on the concentrations of toxic chemicals in most produced waters and predicted dispersion
rates for sites of concern, it is envisioned that there would be only limited potential for acute
toxicity beyond the immediate vicinity of rig sites. This is attributed to the sensitivity of the
biotests (primarily regulatory acute toxicity assays) and the rapid dispersion of the process
stream (Lee et al., 2005a). However, Holdway (2002) noted that to fully assess the potential
impact of produced water discharges, the chronic impacts associated with long-term exposures
must be quantified. Continual long-term chronic exposure may cause sublethal changes in
organisms including decreased community and genetic diversity, lower reproductive success,
decreased growth and fecundity, respiratory problems, behavioural and physiological disorders,
decreased developmental success and endocrine disruption.

179

Effects on Water-Column Organisms

Harmful biological effects in pelagic communities near open ocean produced water discharges
are expected to be minimal and localized, because of the rapid dilution and dispersion rates.
However, some produced waters contain chemicals that are toxic to sensitive marine species,
even at low concentrations. When discharge is in shallow, enclosed coastal waters, or when
discharge is of a low density in an area with low water turbulence and current speeds,
concentrations of produced water chemicals may remain high long enough to cause ecological
harm (Neff, 2002). The chemicals of greatest environmental concern in produced water, because
of their potential for bioaccumulation and toxicity, are metals and hydrocarbons. Highly
alkylated phenols (octyl and nonyl-phenols), though well known endocrine disruptors, have not
been detected in produced water at high enough concentrations to cause significant harm to water
column animals following initial dilution. It has been proposed that elevated nutrient (nitrate,
phosphate, ammonia, organic acids) concentrations may stimulate microbial and phytoplankton
growth in the receiving waters (Khelifa et al., 2003b; Rivkin et al., 2000). Some production
treatment chemicals are toxic. If they occur at high concentrations in produced water, the
discharge could cause localized harm.

Accumulation and Effects in Sediments

If water depths are shallow, some metals and higher molecular weight aromatic and saturated
hydrocarbons may accumulate in sediments near produced water discharges (Means et al., 1990;
Neff et al., 1989; Rabalais et al., 1991) possibly harming bottom living biological communities.
In well-mixed estuarine and offshore waters, elevated concentrations of saturated hydrocarbons
and PAH in surficial sediments may be observed out to a few hundred metres from a high-
volume produced water discharge. The concentrations of PAH in sediments near offshore
produced water discharges are related to the volume and density of produced water discharged,
the PAH concentration in it, water depth, and local mixing regimes. PAHs in sediments near
offshore platforms also may come from drilling discharges, particularly if oil based drilling muds
are used (Neff, 2005).

180

Barium, iron, and manganese are the metals most enriched in produced waters compared to their
concentrations in natural seawater. A phase transition typically occurs following the release of
produced water. The metals tend to precipitate out rapidly when produced water is discharged to
well-oxygenated surface waters that are high in sulphate. These particulate metals tend to settle
slowly out of the water column and accumulate to slightly elevated concentrations in surficial
sediments over a large area around the produced water discharge (Lee et al., 2005a; Neff, 2002).
In addition, the transport and concentration of inorganic constituents within produced water (e.g.
metals) to the surface microlayer may be promoted by the interaction between residual oil
droplets and metal precipitates (Lee et al., 2005a). Toxicity assessment using the Microtox
Test, a regulatory bioassay protocol based on inhibition of a primary metabolic function of a
bioluminescent bacterium, showed that unfiltered samples containing metal precipitates
generally had higher toxicity levels than filtered samples (Azetsu-Scott et al., 2007). Current
results from regulatory environmental effects monitoring programs generally show that natural
dispersion processes appear to control the concentrations of toxic metals in the water column and
sediments just slightly above natural background concentrations.

Aquatic Toxicity

Information on the toxicity of produced water on Arctic species is limited. Thus our assessment
of environmental risk from produced water discharges in the Arctic is primarily linked to studies
in other locations around the globe. As the concentration and composition of produced water
constituents may vary from one geological formation to another, and species sensitivity may
vary, it is recommended that site specific studies be conducted within the Arctic in the future as
the need arises.

Most treated produced water has a low to moderate toxicity. A small number of produced waters
are moderately toxic to mysids (a small shrimp-like crustacean) and sheepshead minnows
(Cyprinodon variegatus), with acute and chronic toxicities less than 0.1% (1,000 mg/L)
produced water. A few produced waters are practically nontoxic with acute and chronic toxicities
higher than 35 or 40%. Most produced waters have moderate toxicities, with acute and chronic
toxicities between about 2 and 10% for mysids and 5 to 20% for sheepshead minnows. Based on
181

earlier toxicity studies for produced waters from the Gulf of Mexico, Neff (1987) reported that
nearly 52% of all LC
50
concentrations were greater than 10% produced water, 37% were
between one and 9.9%, and 11% were less than one percent. These toxicity threshold limits are
consistent with those reported for Atlantic Canada. Mixed function oxidase enzyme activity was
highly induced in the liver, gills and heart of juvenile cod exposed to 5% produced water for 72
hours (Andrews et al., 2007). Histological changes were observed in the gills of fish exposed to
relatively high levels of produced water. In studies with Calanus finmarchicus, a major prey
species for fish in the North West Atlantic, no differences in mortality were found between
control and experimental animals exposed to 5% produced water for 48 hours (Payne et al.,
2001a).

In a comprehensive study on the acute effects of produced water recovered from a Scotian Shelf
offshore well on the early life stages of haddock (Melanogrammus aeglefinus), lobster (Homarus
americanus) and sea scallop (Placopecten magellanicus) in terms of survival, growth and
fertilization success, Querbach et al. (2005) noted that fed, stage I lobster larvae were the most
sensitive with an observed LC
50
of 0.9%. Feeding stage haddock larvae and scallop veligers
were the least sensitive with LC
50
values of 20 and 21% respectively. In terms of chronic
responses, the average size of scallop veligers was significantly reduced after exposure to
produced water concentrations greater than 10%.

There are poorly characterized species differences in the toxicity of produced waters to marine
organisms. When bioassays were performed with two or more marine taxa and the same sample
of produced water, crustaceans were generally more sensitive than fish (Jacobs et al., 1992;
Louisiana Dept. of Environmental Quality, 1990; Neff, 1987; Terrens and Tait, 1993).

Gamble et al. (1987) introduced produced water at a concentration equivalent to a 400 to 500-
fold dilution of the effluent (expected within 0.5 to 1.0 km from the Auk and Forties platforms in
the North Sea) into 300 m
3
mesocosm tanks containing natural assemblages of phytoplankton,
zooplankton, and larval fish. Bacterial biomass increased but phytoplankton production and
larval fish survival were unaffected. However, early life stages of copepods were sensitive to the
produced water and suffered large mortalities. The decrease in zooplankton abundance resulted
182

in an increase in the standing stock of phytoplankton and a reduction in the growth rates of the
fish larvae. In other mesocosm studies summarized by Stephenson et al. (1994), larval molluscs
and polychaete worms also were adversely affected. These mesocosm studies show that low
concentrations of produced water may have subtle effects on marine planktonic communities.
However, it should be pointed out that mesocosm studies represent conservative, worst-case
exposure scenarios, because produced water chemicals in the mesocosm enclosures do not
degrade and dilute as rapidly as they do in well-mixed ocean environments.

Bioaccumulation and Biomarkers as Evidence of Exposure

As mentioned earlier, biomarkers are quantifiable responses in an organism that often indicate
stress due to one or more environmental stimuli. Biomarkers that can be used are based on
various biological activities such as gene expression, protein synthesis, antibody production,
changes in blood chemistry, frequency of a particular deformity, or some other morphological or
biochemical change. Changes in a biomarker are not necessarily a toxic response, but an
indication that the organism is responding to a xenobiotic compound. The effect of PAHs has
been determined in numerous studies conducted in the laboratory and field using endpoints based
on biochemical, histopathological, immunological, genetic, reproductive and developmental
parameters (Payne et al., 2003a). Concentrations of PAHs were measured in the water column
and in blue mussels (Mytilus edulis) deployed at different distances from production platforms in
the Norwegian sector of the North Sea (Durell et al., 2006; Johnsen et al., 1998; Neff et al.,
2006b; Re Utvik et al., 1999). Direct measurements of PAH in the water gave inconsistent
results, because concentrations were too low and variable. However, the mussels did
bioaccumulate PAHs from the water, and concentrations decreased with distance down-current
from platforms.

PAH residues in mussel tissues were used to estimate PAH concentrations in surface waters
(Neff and Burns, 1996). Surface water TPAH concentrations ranged from 0.025 to 0.35 g/L
(parts per billion) within 1 km of platforms and reached background levels of 0.004-0.008 g/L
within 5-10 km of discharge, which is equivalent to about a 100,000-fold dilution of the PAH
concentration in the discharged water. Dilution modelling showed that most of the produced
183

water plume was restricted to the upper 15 to 20 m of the water column. Dilution was very rapid.
The Dose-related Risk and Effect Assessment Model predicted that the concentrations of PAH
and other chemicals in the produced water plume exhibited wide cyclic concentration variations
due to tidal and wind-driven current flows. Because of rapid dilution and fluctuating water-
column concentrations, the model predicted that potentially toxic concentrations and contact
times of PAH would not occur even in the near-field.

Brseth and Tollefsen (2004) found that concentrations of metals and PAH in soft tissues of the
caged mussels correlated well with distance from the discharge, with highest body burdens in
mussels closest to the platform. The PAH assemblage in mussel tissues was dominated by alkyl
homologues of naphthalene, phenanthrene, and dibenzothiophene, indicating that the PAHs were
from petroleum, likely from the produced water discharge. Biomarker responses in the mussels
provided equivocal evidence of exposure to produced water chemicals. The authors concluded
that mussels and cod deployed near a discharge probably were exposed to low concentrations of
produced water chemicals, below levels that might represent a health risk to water-column
organisms.

Elevated mixed function oxidase enzyme levels have been noted in fish larvae collected
downstream of the Hibernia field on the Grand Banks of Canada (Payne et al., 2003a). However,
induction may only be occurring near the well site with the induced larvae being transported
downstream by currents. Petro-Canada and Husky Energy have performed biomarker studies
with American plaice (Hippoclossoides platessoides) collected in the vicinity of the Terra Nova
and the White Rose offshore developments on the Grand Banks of Newfoundland (DeBlois et
al., 2005; Husky Energy, 2005). These studies showed that the overall health of the American
plaice was similar to the health of American plaice collected at distant reference sites. American
plaice are demersal (bottom-living) fish and may not have been exposed to the produced water
plume in the upper water column.

As part of the Biological Effects of Contaminants in Pelagic Ecosystems Program,
bioaccumulation and several biomarkers were measured in wild and caged marine animals along
a transect away from the Statfjord Platform in the North Sea (Hylland, 2006). Produced water
184

discharge is 74,100 m
3
/day from three platforms in the Statfjord field (Durell et al., 2006),
among the highest discharge rates of any offshore field in the world. Frlin and Hylland (2006)
measured EROD activity and bile metabolites in juvenile cod caged at several distances down-
current from one of the discharges. There were no significant trends in EROD activity in male
and female cod with distance from the discharge, though there was a trend for EROD activity in
female cod to increase with distance from the discharge, a trend opposite the expected one. In
spite of this, concentrations of alkylnaphthalene metabolites in fish bile were highest in cod near
the platform and decreased with distance from the platform. Alkylnaphthalenes are abundant in
produced water. There were no distance trends in concentrations of other PAH metabolites in
cod bile. The authors concluded that the cod were exposed to low levels of PAH from the
produced water discharges, but exposure levels were well below those that would pose a health
risk to fish living near the platforms.

A recent study by Hamoutene et al. (2010) investigated the effects of produced water on cod
immunity, feeding and general metabolism by exposing fish to diluted produced water at
concentrations of 0, 100 and 200 ppm for 76 days. No significant differences were observed in
weight gain or food intake. Similarly, serum metabolites, whole blood fatty acid percentages, and
mRNA expression of an appetite-regulating factor (cocaine and amphetamine regulated
transcript) in the brain remained unchanged between groups. Other than an irritant-induced
alteration in gill cells found in treated cod, resting immunity and stress response were not
affected by produced water. Changes in enzymatic activity of catalase and lactate dehydrogenase
were observed in livers but not in gills, suggesting an effect on oxidative metabolism subsequent
to hepatic detoxification processes. To evaluate potential effects of produced water discharges on
cod immunity, fish from the three groups were challenged by injection of Aeromonas
salmonicida lipopolysaccharides at the end of exposure. The lipopolysaccharide injection
affected respiratory burst activity of head-kidney cells, and circulating white blood cells ratios,
and increased serum cortisol in all groups. The most pronounced changes were seen in the group
exposed to the highest dose of produced water (200 ppm).


185

Alteration of Trophic Level Dynamics by Produced Water

A modelling study was carried out to assess potential perturbations in food web structure and
energy flow due to the discharge of produced water (Rivkin et al., 2000). The model predicted
significant increases in productivity and sedimentation fluxes over large spatial domains in
response to produced water derived ammonia and dissolved organic carbon.

Bacteria have very short generation times and respond rapidly to environmental changes. As
bacteria are involved in primary processes including the production of carbon, nutrient cycling,
biodegradation and biotransformation of contaminants their use in environmental effects
monitoring programs has been recommended (Lee and Tay, 1998; Wells et al., 1998). Studies
by Anderson et al. (2000) with naturally occurring bacteria indicated the potential for produced
water to both inhibit (short-term exposure at high concentrations) and enhance bacterial growth
(lower concentrations over an extended period).

In a recent study by Yeung et al. (2010) changes in indigenous microbial community structures
in response to the discharge of produced water from an offshore platform on the Grand Banks of
Canada were monitored by DGGE. The DGGE results showed that the production water did not
have a detectable effect on the bacterial populations in the surrounding water. Cluster analysis
showed greater than 90% similarity for all near surface water (2 m) samples, approximately 86%
similarity for all the 50 m and near bottom samples, and approximately 78% similarity for the
whole water column from top to bottom across a 50 km range, based on two consecutive yearly
sampling events. However, there were distinct differences in the composition of the bacterial
communities within the produced water compared to seawater near the production platform
(~50% similar), indicating that the effect from produced water may be restricted to the region
immediately adjacent to the platform. Specific microorganisms (Thermoanaerobacter of the
eubacteria, and Thermococcus and Archaeoglobus of the archaea) were detected as significant
components of the produced water. These particular signature microorganisms may be useful as
markers to monitor the dispersion of produced water into the surrounding ocean.

186

Ecological Risk of Produced Water Discharges

The Norwegian oil and gas industry advocates ecological risk assessment as the basis for
managing produced water discharges to the North Sea. Neff et al. (2006b) compared estimates of
ecological risks of PAHs from produced water to pelagic communities based on data from
hydrocarbon residues in soft tissues of blue mussels (M. edulis) deployed for a month near
offshore platforms and based on predictions of the Dose-related Risk and Effect Assessment
Model. The study was performed near produced water discharges in the Tampen and Ekofisk
Regions of the Norwegian Sector of the North Sea. Because PAHs are considered the most
important contributors to the ecological hazard posed by produced water discharges,
comparisons focused on this group of compounds. The two methods ranked stations at different
distances from produced water discharges in the same order and both identified 2 and 3-ring
PAHs as the main contributors to the ecological risk. Neither method identified a significant
ecological risk of PAH in the upper water column of the oil fields.

Due to chemical kinetic reactions that occur following the release of produced water in an anoxic
state into the open ocean, it has been suggested that the toxicity of produced water may change
over time following its discharge (Lee et al., 2005a). The significance of this process is clearly
illustrated in controlled dose-response experiments using natural microbial populations as the
test organisms. A typical toxicity dose-response curve (initial increase in productivity at low
concentrations of produced water due to addition of nutrients, followed by inhibition above a
threshold value) was observed with fresh produced water, that is, process water that has been
treated but not discharged. However, following aeration for 44 hours (to simulate equilibration in
the ocean following discharge) additions of the produced water over the same concentration
gradient elicited a stimulatory response. The difference after aeration was attributed to the loss of
low molecular weight hydrocarbons, and the precipitation of hydrolysis metals that may have
sequestered toxic metals and other chemicals. These results imply that accurate comparisons of
toxicological studies with similar endpoints (e.g., LC
50
) cannot be made unless sample collection
and handling protocols are standardized prior to toxicity testing.

187

Produced Water Treatment

Most environmental regulatory agencies in countries that have significant offshore oil and gas
production place limits on the concentrations of petroleum (usually measured as total oil and
grease) that can be present in produced water destined for ocean discharge. Table 10 gives
examples of limits on oil and grease imposed by different countries.

Table 10 National permissible concentrations of total oil and grease in produced water destined for ocean
disposal (Veil, 2006).
Country Monthly Average (mg/L) Daily Maximum (mg/L)
Canada 30 60
USA 29 42
OSPAR (NE Atlantic) 30 ---
Mediterranean Sea 40 100
Western Australia 30 50
Nigeria 40 72
Brazil --- 20

Produced water intended for ocean disposal usually is treated on the platform or at a shore
treatment facility to meet these regulatory limits. The objectives of oil-water-gas treatment on an
offshore platform are to produce stabilized crude oil and gas for pipeline or tanker transport to
shore facilities, and to generate a produced water that meets discharge requirements (if
discharged to the ocean) or is suitable for reinjection into the producing formation or another
geologic formation (Bothamley, 2004).

Within Canada, the formulation of regulatory guideline values for produced water discharge is an
adaptive process that promotes the development of improved environmental effects monitoring
programs that takes into consideration the level of environmental risk, and the best available
technology for mitigative measures and socio-economic benefits. Currently, waste discharges in
the Arctic are subject to strict regulatory standards, with the U.S. and Norway having more
mature regulatory systems compared to Russia, Greenland and Iceland (Hawboldt et al., 2010).
A summary of the treatment systems used in three Arctic regions (Beaufort Sea in Alaska,
Norwegian Barents Sea, and Sakhalin Island in Russia) can be found in Table 11. In all cases the
188

projects have adopted a zero-discharge limit. Although re-injection is a common practice, it is
associated with high energy usage and the creation of large amounts of air emissions. To meet
the future challenges for the treatment and disposal of produced water in the Arctic and the
establishment of improved regulatory guidelines, new technologies will need to be developed
and implemented.

Table 11 Summary of produced water treatment systems used by three Arctic oil and gas installations;
adapted from Hawboldt et al. (2010).
Installation/Oil and Gas Field Water
Depth (m)
Treatment Technology
BP Northstar/North Slope
Alaska
12 A single well was drilled and used for disposal of
all waste streams, including produced water and
drill cuttings.
Norwegian Barents Sea ~400 Existing operations are required to dispose of
produced water through re-injection. No new oil
and gas development is permitted due to
environmental concerns.
Piltum-Astokhskoye/Sakhalin
Island, Russia (12-15km
offshore)
1900-2500 Produced water is treated and used as a
supplement of the primary source seawater.
Azeri Chirag
Gunashli/Azerbaijan, Russia
125-175 An electrostatic coalescer is used to separate
95% of the produced water which is re-injected
into the geologic formation. The remaining 5% is
transported back to shore for further treatment.


Environmental Effects Monitoring and Research Needs

The general consensus of the scientific community is that any effects of produced water on
individual development sites in the open ocean are likely to be minor (Lee and Neff, 2009). The
toxicity threshold limits for acute effects are not likely to occur beyond the immediate vicinity of
the discharge pipe due to the effectiveness of natural dispersion processes driven by tides,
currents and wind. However, unresolved questions regarding aspects of produced water
composition, its fate and potential effect on the ecosystem remain. The effects of chronic toxicity
may only become evident after monitoring several life stages or generations. It is important to
189

acknowledge the consequences of long-term effects from offshore oil and gas facilities that may
have a 15-20 year lifecycle. Furthermore, cumulative effects linked to future expansion of
production operations must be considered. It is evident that additional information is needed to
improve the accuracy of existing risk assessment models for produced water discharge.
Multidisciplinary scientific studies are needed under an ecosystem based management approach
to provide information on the environmental fates (dispersion, precipitation, biological and
abiotic transformation) and effects of chronic, low-level exposures of the different chemicals in
produced water.

Numerical models need to be improved to better predict the fate and effects of chemical
constituents in produced water plumes that are rapidly dispersed. There is a need to develop
improved sample recovery and analytical techniques to support model validation needs. At
present many of the potential contaminants of concern in produced water cannot be detected in
the open ocean environment with standard analytical protocols. The future development of high
efficiency, cost-effective produced water treatment technologies is dependent on the
identification and monitoring of the primary target constituents of environmental concern (e.g.,
PAH, phenols, metals).

For a comprehensive protection plan, there is a need to support the development of improved
monitoring protocols to provide early warning of any potential problems related to sediment and
water quality (such as primary productivity), fish quality and fish health. Development of real-
time monitoring systems (contaminant specific sensors and data-transfer technologies) will
enhance our capacity to manage the ocean and its living resources. In consideration of natural
perturbations currently occurring within the ocean (e.g., climate change) and the impacts
potentially associated with other users of the oceans (marine transport, fisheries, etc.), an
ecosystem based integrated management approach must be taken to fully evaluate the risks of
produced water discharge into the oceans.

Interpretation of ecological risk from biological effect studies based on biomarker techniques
remains a challenge. Biomarkers may be used to indicate that 1) an organism has been exposed
to a specific chemical or group of chemicals; 2) an organism is affected by a contaminant and
190

responding to it; and 3) the organism has been damaged. However, as discussed by Gray (2002)
in an editorial comment entitled, Perceived and real risks: Produced Water from oil extraction,
the question at the site of concern is, What is the risk to populations in the field?

8.2 Drilling Muds

Drilling muds are used by the offshore oil and gas industry to cool and lubricate the drill bit, to
balance subsurface hydrostatic pressure, and to carry drill cuttings up to the surface through
drilling pipes. Drilling muds typically contain a base fluid (water-based, oil-based, or synthetic-
based), barite or bentonite (weighting agent), and chemical additives (emulsifiers, biocides,
lubricants, wetting agents, corrosion inhibitors, surfactants) to enhance the muds operational
properties (GESAMP, 1993). On the drill rig, the cuttings and muds are separated to recycle the
muds while the cuttings once treated may or may not be discharged into the sea. Once
discharged, drilling muds and cuttings undergo a number of physico-chemical processes in
marine environments: (1) advection, (2) dispersion, (3) aggregation, (4) settling, (5) deposition,
(6) consolidation, (7) erosion, (8) re-suspension, (9) re-entrainment, and (10) change in bed
elevation. The relative impacts of these processes on the fate of drilling wastes in marine
environments depend on the characteristics of drilling wastes and physical variables of the water
body such as water depth, currents (tidal and residual), waves and storms.

The biological effects of drilling wastes generally can be thought of as being caused primarily by
(1) chemical toxicity from hazardous pollutants and biodegradation products; (2) organic
enrichment of the seabed that may result in anoxic conditions; (3) physical smothering due to
accumulation; and (4) physical effects on tissues (causing reduction in growth and reproduction)
due to chronic exposure to very low concentrations of the drilling mud compounds bentonite and
barite (Cranford, 2006).

Most drilling of offshore oil and gas wells in the North Sea, the Gulf of Mexico and other
offshore production areas is achieved with water-based drilling muds (WBM). This is due to
strict regulations on discharges of oil-based mud (OBM) and synthetic-based mud (SBM) as a
result of their potential environmental impacts. SBM were designed to be less toxic and more
191

environmentally friendly than OBM. The use of SBM over WBM comes with concerns over the
risk of organic enrichment and the potential persistence of synthetic-based fluid biodegradation
products. However, these risks must be weighed against the higher levels of turbidity and metal
contaminations associated with bulk WBM disposal. Veil et al. (1995) noted that WBM use
typically generates between 1100 and 2000 cubic metres of muds and cuttings depending on the
depth and diameter of the well compared to 300-1300 cubic metres for SBM.

Although they were first used in the North Sea, the use of SBM in the North Sea has been
curtailed since 2001 when the OSPAR Commission (a consortium of fifteen Governments of the
western coasts and catchments of Europe, together with the European Community cooperated to
protect the marine environment of the Northeast Atlantic) stated that the discharge into the sea
of cuttings contaminated with synthetic fluids shall only be authorized in exceptional
circumstances. WBM, but not OBM or SBM, may be permitted for ocean discharge into
European offshore waters provided that they do not contain hazardous chemicals. All drilling
mud chemicals intended for offshore disposal must be tested for toxicity, bioaccumulation and
biodegradability. Norway recently introduced a zero harmful discharge policy. Although this was
mainly intended for produced water discharges, it was extended to drilling wastes as well.
However, the Norwegian government is re-evaluating the discharge of WBM since it may be less
environmentally harmful than the alternative disposal practices. Both WBM and SBM cuttings,
but not OBM cuttings, also are permitted for discharge into offshore waters of the U.S. Gulf of
Mexico, with requirements of the SBM material registered as biodegradable under anaerobic
conditions with indigenous microorganisms.

Current regulations in Canada (NEB et al., 2002) stipulate recovery and onshore disposal of the
cuttings generated using OBM. SBM remaining from a drilling mud change-over or drilling
program completion should be recovered and recycled, re-injected down-hole, or transferred to
shore. The cuttings produced with WBM are allowed to be discharged into the sea as well as
those produced with SBM following treatment with the best available technology. Best available
technology in some offshore regions internationally is believed to be capable of achieving a
concentration of 6.9 g/100 g oil on wet solids. This discharge limit may be modified in
individual circumstances where more challenging formations and drilling conditions are
192

experienced or areas of increased environmental risk are identified. It is anticipated that
performance will improve in the future as further advancements in technology and operating
procedures evolve.

Arctic Marine Food Webs and Toxicity

There is evidence that trophic level dynamics in the Arctic may diminish the toxic effects of
drilling muds. Arctic marine food webs are relatively simple compared to those in more
temperate climates. Metals and hydrocarbons are not likely to enter the food web or
bioaccumulate up the trophic levels. A number of short-term and long-term toxicity tests have
been performed on whole WBM as well as WBM ingredients with a wide variety of freshwater
and marine plants and animals, including many species from Arctic and cold-water environments
of Alaska, Canada and Norway. WBM containing ingredients permitted by current U.S. effluent
guidelines or on OSPARs PLONOR list (List of Substances/Preparations Used and Discharged
Offshore That Are Considered to Pose Little or No Risk to the Environment) were non-toxic to
all species. A number of microcosm and mesocosm experiments have confirmed that WBM
cuttings have no effects, or minimal and short-lived effects on zooplankton and benthic
communities. The lack of bioaccumulation or toxicity of WBM indicates their effects are highly
localized within close proximity of the discharge point, and are not being exported to the food
web.

No measurable adverse effects of routine WBM discharges on the water column environment
have been reported, because most are small amounts discharged intermittently over the course of
several months (Neff, 2005; NRC, 1983). When ecological effects are detected, it is often due to
physical disturbances caused by elevated concentrations of suspended particles in the water
column which may clog the gills or digestive tract of zooplankton or benthic filter-feeding
invertebrates. Drilling wastes may also bury some sessile or low mobility benthic species. By
changing the sediment grain size and texture it may make the substrate unsuitable for certain
native species. The biodegradation of settled organic material in the WBM may also stimulate
bacterial growth leading to a depletion of oxygen. Overall, these processes may cause changes in
193

species abundance, composition and diversity of the benthic community, although it is rare for
this to occur in sediments near WBM discharges.

It is known that the rate of benthic recovery depends on the thickness of the mud and cuttings
that accumulate on the sea floor, but recovery might also be slower in cold water environments
compared to more temperate locations. In general, the physical disturbances to the water column
and sediments from SBM discharges are similar in character and magnitude to the disturbances
caused by natural processes such as storms, and inputs of suspended sediments from Arctic
rivers during spring breakup and ice scour. Arctic species are well adapted to seasonal
disturbances and will recover rapidly from the relatively brief and intermittent disturbances
associated with exploratory drilling operations.
194

9. Case Study: The Exxon Valdez Oil Spill

The Exxon Valdez ran aground on Bligh Reef, Prince William Sound (PWS) on March 24, 1989,
and released approximately 42 million litres of ANS crude oil into the coastal marine
environment (Spies et al., 1996). Of this, approximately 40% washed ashore (Gilfillan et al.,
2000) and 783 km of coastline were oiled (Neff et al., 1995). The Alaskan coast is known for its
rich marine life of seabirds, marine mammals, whales and seals, as well as supporting a multi-
million dollar fishery of pink salmon, pacific herring and groundfish. The rugged coastline is
also a backdrop for tourism which, along with the fishery, is a key component of the Alaskan
economy (Spies et al., 1996). The immediate impact of the spill was the death of about 250,000
sea birds, 22 killer whales, 2800 sea otters, 300 harbour seals, and untold numbers of fish eggs
(Guterman, 2009).

Several strategies were employed to treat the Exxon Valdez crude oil (EVCO). Some beaches
were left untreated while others subjected to cleanup. Initially, oil was removed from the
shoreline with aggressive washing techniques, but eventually less destructive methods were used
(Wells et al., 1995). Cleanup methods included low pressure warm-water washes, high pressure
hot water washes, the dispersant Corexit 7664, and the beach cleaner Corexit 9580 M2. While all
treatments were suspected of causing some deleterious effects, the most damaging was
application of high pressure hot water, with the areas subjected to this treatment exhibiting
impacts six years after the spill (Lees et al., 1996).

General reviews of the effects of the Exxon Valdez oil spill (EVOS) have been published at
regular intervals (Paine et al., 1996; Peterson, 2001), with the most recent of by Peterson et al.
(2003) and Harwell and Gentile (2006). Peterson et al. (2003) stated that as of 2003, there were
detectable effects of the EVOS on the PWS ecosystem, but that they were no longer the result of
acute toxicity. Chronic exposure to sublethal concentrations of EVCO was now the concern, and
these effects were being observed in many species in PWS. The authors stated that there had
been a paradigm shift, moving away from the effects of short-term acute toxicity on individual
species to an ecosystem-based synthesis of short and long-term impacts. Harwell and Gentile
(2006) examined about 500 peer-reviewed articles in order to assess whether the marine
195

ecosystem had recovered by 2006. They applied the concept of a VEC, including birds (bald
eagles, harlequin ducks), marine mammals (otters, orcas, seals), invertebrates (clams, mussels)
and other components referred to as higher levels of organization (e.g. intertidal and subtidal
communities, the trophic structure of PWS, and water quality). By examining the literature on
effects related to VECs, they came to the conclusion that although the EVOS and associated
cleanup had profoundly affected the ecosystem from the time of the spill until a few months to a
few years later, longer lasting effects were masked by anthropogenic stressors in PWS that were
not associated with the EVOS.

The long-term impacts of the EVOS are still in dispute. While the waters of PWS are now clear
and shorelines bristle with life, oil still lingers, hidden beneath rocks and below the surface of
beaches (Guterman, 2009). To date, the controversial environmental assessment studies
following the incident have been the longest, largest and most expensive ever done. Consensus
has not been reached between various parties of government, the public and industry over the
level of long-term effects and their environmental significance. Much of the dispute revolves
around whether or not industry should be held responsible for funding additional monitoring,
research and remediation operations (Guterman, 2009; Landis, 2007). The problem with
quantifying impact assessment, toxicity, and recovery of an ecosystem is that all three processes
depend on a number of factors (Shigenaka, 2005) that are influenced by the investigator, the
metrics used, the benchmarks for determining what is affected, and the statistical approached
employed. As a result, any definitive answer to a question like, Has the PWS ecosystem
recovered from the EVOS? would be next to impossible to achieve. In addition, the definition
of recovery is problematic. Peterson et al. (2003) seem to regard it as synonymous with a return
to pre-spill conditions, while Harwell and Gentiles (2006) utilization of a more value-driven
approach is based on the assumption that an ecosystem will reach equilibrium (Landis, 2007). In
actuality, there are many definitions available (Parker and Wiens, 2005). Inevitably a bias a will
develop, depending on which definition is chosen, as this will be a policy-driven process
(Landis, 2007).

The chemistry, toxicity and fate of EVCO, plus the effect on birds, mammals, fish, invertebrates,
and the intertidal zones for the first several years after the spill has been well documented. Books
196

such as Proceedings of the Exxon Valdez Oil Spill Symposium (Rice et al., 1996) and Exxon
Valdez Oil Spill: Fate and Effects in Alaskan Waters (Wells et al., 1995) cover the topics well,
and several reviews (Harwell and Gentile, 2006; Paine et al., 1996; Peterson, 2001; Peterson et
al., 2003) are good summaries of fate and effects up to the time they were published.

9.1 Fate of the Oil

Fractionation of the spilled oil started soon after its release from the hull of the Exxon Valdez.
Physical, chemical and biological processes sub-divided the oil into components that travelled at
different rates depending on their chemical properties and environmental conditions (Spies et al.,
1996). Volatile components evaporated as the slick was transported by wind and waves. Three
days after the spill, a storm occurred in PWS, which further fractionated the oil through
evaporation, dissolution, physical dispersion, emulsification, adsorption onto suspended particles
and grounding on the shoreline (Spies et al., 1996).

Residual oil continues to be detected on beaches in PWS. Short et al. (2004) discovered oil on 71
of 91 beaches randomly selected according to their oiling history in 2001, and arrived at a
conservative estimate of 55,600 kg of EVCO remaining sub-surface, putting the beaches where
the oil still resides at continued risk of biological exposure. This estimate of residual oil has been
supported by the calculations of (Pella and Maselko, 2007) who estimated that PWS had 41,000
m
2
of surface oiled areas and 71,000 m
2
of sub-surface oiled area in 2001, representing
approximately 50,000 kg of oil. Taylor and Reimer (2008) found EVCO present in and on 39
beaches using the shoreline cleanup assessment team field observation program procedures. The
oil was detected as asphalt pavements in the upper intertidal to supratidal zones, while
subsurface oil persisted as an approximately 3 cm thick band about 5-10 cm below a pebble and
gravel veneer. They also discovered that at six previously oiled sites there was no longer any oil
detected and concluded that, based on their calculations, only 0.5% of the shorelines of PWS
contained residual oil. They postulated that the persistence of the oil was due to it being
incorporated into finer sediments, isolating the oil from active weathering in shoreline wave
shadows with limited wave action (Taylor and Reimer, 2008).

197

Exxon-funded researchers did not dispute the presence of residual EVCO in PWS. They did,
however, question the amount and the condition in which it is found. By 1992, the area occupied
by sub-surface oil was estimated to be about 12,000 m
2
(Neff et al., 1995). In addition, a more
recent study (Boehm et al., 2008) of 22 sites that had been most heavily oiled in 1989 concluded
that in over 90% of the samples collected, there was no oil. In addition, those samples that still
contained EVCO were characterized by heavily (>70%) weathered oil that was located in mid to
upper intertidal zones (Boehm et al., 2008), away from areas of biological activity. Boehm et al.
(2008) concluded that most of the biologically-available oil in PWS had been eliminated through
natural weathering. This idea was disputed by (Short et al., 2006) who discovered residual
EVCO on 14 shorelines, with most of the oil located near the mid-intertidal zone and about half
of the sub-surface oil located in the lower intertidal zone rich in biological activity. As a result,
they stated that there was a significant likelihood of sea otters and harlequin ducks encountering
sub-surface oil.

The conditions contributing to the creation of persistent EVCO on PWS beaches have been
investigated in detail. Beaches exposed to oil mousse, an emulsion of EVCO and water (see
Section 3.4 Factors Influencing Oil Fate (Weathering), Emulsification), contained persistent oil
(Short et al., 2007), while beaches that were not exposed to the mousse did not have as much
residual oil (Irvine et al., 2006). Oil mousse appeared to resist weathering, resulting in the
prolongation of the effect of toxic PAHs (Short et al., 2007) and was compositionally similar to
11-day old EVCO (Irvine et al., 2006). In addition, physical processes associated with persistent
oil included boulder armour, where finer sediments trapped beneath boulder lags resulted in the
persistence of lightly-weathered oil in a high-energy environment (Irvine et al., 2006). Added to
this, a permeable surface sediment can act as a temporary storage layer, and then funnel oil to
more permanent storage in underlying low-permeability sediment when the water table drops
(Xia et al., 2010). This lower layer in which the oil can become isolated exhibits nearly anoxic
conditions, reducing the possibility of biodegradation or weathering (Li and Boufadel, 2010).
When combined with tide-induced exchange of groundwater in mid-intertidal to high intertidal
zones, this would result in the persistence of oil in low intertidal regions (Xia et al., 2010).

198

Subsurface oil from EVOS still exists in scattered patches within shoreline sediments of low
permeability (Boehm et al., 2011). The oil in these sediments is protected from weathering
(erosion, dissolution, biodegradation) because of surface boulder and cobble armour. As such,
the oil is not accessible or bioavailable to wildlife that forage on the shore (Boehm et al., 2011).
This is also the reason why PAH levels in the water of shorelines identified as having the most
residual oil are similar to background: the trapped oil is found in sediment that is 100 to 1000
times less permeable than the surrounding sediment (Pope et al., 2011). Laboratory experiments
showed that there was limited contact between groundwater and oiled sediments, and if any PAH
entered the aqueous phase, it would be readily washed out and diluted by tidal action (Pope et al.,
2011).

Geospatial models have been used to locate where subsurface oil from the Exxon Valdez is most
likely to persist after 19 years following the spill (Michel et al., 2011). Field data collected from
314 sites (2001 to 2007) were used to identify geomorphologic and hydrologic factors involved
in the persistence of the subsurface oil. The relationships between the data and the physical
factors controlling oil persistence were not clearly understood in all cases. The results indicated
several unsurveyed shoreline locations as potentially having subsurface oil, which could be
quantitatively prioritized for investigation.

Samples of beach substrate containing subsurface 19 year-old oil from EVOS were tested in
microcosms for biodegradability (Venosa et al., 2010). With nutrient amendment, oil in the
substrate was found to be significantly (p < 0.01) more biodegradable than natural attenuation,
no matter what the degree of weathering. The reason that the subsurface oil had persisted was
predominantly due to a lack of oxygen in the interstices of the sediment from where the oil was
retrieved. Nitrogen was found to be a secondary factor limiting biodegradation. Biologically
available nitrogen is often relatively low in marine environments, and hydrocarbon loadings of
>10 ppm in a closed laboratory microcosm will likely result in low rates and extent of
biodegradation (Lee et al., 2011d). This was overcome by the addition of nutrients.
Biodegredation occurred in the natural attenuation microcosms because of the high
concentrations of nitrogen in the substrate, and the aerobic test conditions (Venosa et al., 2010).

199

9.2 Effects of the Oil

Bioavailability

One important question concerning the residual oil is whether or not it poses a hazard to the
environment. If the remaining oil is biologically inert, meaning that it is either no longer
releasing compounds, or any compounds that are released are non-toxic, then it is considered to
be biologically unavailable. PAHs are considered to be the primary source of stranded oil
toxicity, especially those of higher molecular weight. The exposure of intertidal animals to
bioavailable PAHs from residual EVOS oil was first noted by (Roberts et al., 1999), and
continued to be observed over the next decade (Deepthike et al., 2009; Springman et al., 2008;
Thomas et al., 2007). Venosa et al. (2010) have demonstrated that 19-year-old weathered EVCO
can still be biodegraded. The fact that this oil is still biodegradable could be due to several
factors including the absence of biodegrading bacteria, an electron acceptor, and nutrients
necessary to facilitate microbial degradation in the zones where the oil is trapped. The
significance of these observations have been questioned by other scientists who contend that the
oil remaining from the EVOS is buried below zones frequented by the biota, or is weathered to
the extent that the residues exude PAHs at low concentrations (Boehm et al., 2007a; Neff et al.,
2006a), or release PAHs that are not toxic (Page et al., 2002). These scientists also contend that
PAHs from the EVOS residual oil are now at such low levels (Neff et al., 2006a), that they are
masked by PAHs released by non-EVOS sources (Page et al., 1996; Page et al., 2006). This
contention is supported by the results from the Long-Term Environmental Monitoring Program,
indicating that TPAH concentrations in the PWS ecosystem are now at background levels (Payne
et al., 2008).

Shoreline Flora and Fauna

Peterson et al. (2003) reported that clams and other invertebrates living in subtidal and intertidal
areas in PWS with contaminated sediments remained contaminated for more than a decade.
However, recent literature (later than 2004) cannot be found on the response of the PWS
ecosystem to EVCO, and even though the initial stranding of oil on the shoreline had an effect on
200

the biology of the infauna, epifauna and macroalgae (Dean and Jewett, 2001; Driskell et al.,
1996), it was the methods used for shoreline cleanup (primarily high-pressure, hot water
washing) which had the most devastating effect. The cleanup methods resulted in thermal shock
in the intertidal zone (Driskell et al., 1996; Lees et al., 1996). All sites in PWS exhibited signs of
recovery by the early 1990s (Driskell et al., 1996; Gilfillan et al., 1995a; Gilfillan et al., 1995b;
Houghton et al., 1995). However, the recovery of cleaned sites lagged behind sites that were left
to natural attenuation (Driskell et al., 1996; Lees et al., 1996; Skalski et al., 2001). Habitat-
specific recovery of shallow, subtidal communities following the EVOS were investigated (Dean
and Jewett, 2001). Impacts were greatest in sheltered bays that were subject to heavy oiling.

In subtidal areas of eelgrass (Zostera marina), the long-term effects of EVOS were identified as
decreased mean density of shoots and flowering shoots (Dean et al., 1998). Over a period of up
to 10 years after the spill, impacts were greatest in nearshore, subtidal (<11 m deep) communities
of sheltered bays that had been heavily oiled, in which it was observed that the effects of oiling
and cleanup were more pronounced on the eelgrass beds than in kelp habitats comprised mostly
of Agarum clathratum and Laminaria saccharina (Dean and Jewett, 2001). One year after the
spill, concentrations of TPAHs were higher in the eelgrass beds than among the kelp. Also, there
were more groups of organisms at lower densities in eelgrass than in kelp beds in the oiled sites
compared with control sites.

Dean and Jewett (2001) noted that recovery was slower for eelgrass than kelp, with
approximately 80% of the impacted groups of organisms in eelgrass beds appearing unrecovered
in 1995, while most of the impacted groups in kelp beds recovered in two years. It was suggested
that the protracted recovery of the eelgrass beds was probably because of them having more
sensitive species such as crustaceans, in addition to the higher concentration of TPAHs, and the
persistence of the oil within the sheltered, soft sediments as opposed to the sand and boulder
substrate of the kelp beds.



201

Invertebrates

Both mussels and clams are sources of food for foraging sea otters and harlequin ducks, and
mussels were used as part of the Long-Term Environmental Monitoring Program. The program
was started in 1993 to assess the presence of lingering oil (as PAHs and saturated hydrocarbons)
in the tissues of mussels from sites in PWS and the Gulf of Alaska. Prolonged effects of PAH
exposure on mussels were still detectable in 2002 (Thomas et al., 2007), but have subsequently
been judged to be at background levels since 2004 (Boehm et al., 2004; Payne et al., 2008;
Reynolds and Braman, 2009), presenting no further risk to foraging sea otters or harlequin ducks
(Neff et al., 2011).

Fish

The long-term effects of the EVOS on fish in PWS are the subject of ongoing dispute. Initial
effects on fish species were thought to be the result of reduction in the number of adult fish
returning to PWS in later years due to oiling of juveniles (Geiger et al., 1996; Willette, 1996) or
exposure of embryos or larvae to oil or its residues (Brown et al., 1996; Carls and Thedinga,
2010; Rice et al., 2001). Benthic fish species appeared to suffer less with reports of no pervasive
deleterious effects (Armstrong et al., 1995), although some studies reported exposure to organic
contaminants through activation of the CYP1A gene in subtidal fish species through 1991
(Collier et al., 1996). Pelagic fish displayed no CYP1A induction from 1999-2000 (Huggett et
al., 2003).

Pacific herring (Clupea pallasi), a major commercial species in PWS, suffered a severe decline
in 1993. Some attributed this to a poor phytoplankton bloom in 1992 (leaving the fish hungry
and vulnerable to disease), while others attributed the population crash to the spill (Guterman,
2009; Thorne and Thomas, 2008) through mechanisms such as the sensitivity of embryos to oil
(Carls et al., 2000) and cardiotoxic effects (Incardona et al., 2009).

The other commercial species examined for effects from the EVOS was the pink salmon
(Oncorhynchus gorbuscha). The mechanism of damage to this fish is thought to have been
202

through exposure of eggs and embryos to toxic PAHs released from oil residues in areas where
the salmon spawn (Carls and Thedinga, 2010; Heintz et al., 1999). Principal issues concerning
continued effects from exposure to PAHs from EVCO residues are that PAH concentrations in
PWS represent ambient background levels (Payne et al., 2008), being below that which cause
effects in pink salmon embryos (Brannon et al., 2006a; Brannon et al., 2007; Brannon et al.,
2006b) and that residues from EVOS are not the only sources of PAH in PWS (Burns et al.,
1997; Huggett et al., 2006; Page et al., 1996; Page et al., 2006; Payne et al., 2008; Roberts et al.,
2006).

Birds

Bird mortalities due to EVCO were in the hundreds of thousands due to direct oiling from
floating oil (Esler and Iverson, 2010; Iverson and Esler, 2010). The long-term effects (chronic,
sublethal) mentioned in the review of Peterson et al. (Peterson et al., 2003) are the cause of
continued debate.

Harlequin ducks (Histrionicus histronicus) were studied extensively due to their close
association to marine intertidal habitats and their feeding on benthic invertebrates that can
bioaccumulate oil residues. This made the ducks particularly vulnerable to the effects of residual
oil (Iverson and Esler, 2010). While an estimated 25% of the PWS duck population died as a
result of acute exposure during the immediate aftermath of the spill (Iverson and Esler, 2010),
74% of ducks collected in 1989 and 1990 were found to have hydrocarbon metabolites in their
systems, implying further lethal or sublethal damages (Esler et al., 2002).

Researchers found that while the survival rates of harlequin duck populations from oiled and
unoiled areas were indistinguishable from one another 11-14 years after the EVOS (Esler and
Iverson, 2010; Iverson and Esler, 2010), results from application of the biomarker CYP1A
indicate that the ducks continued to be exposed to oil residues up to 20 years after the spill (Esler
et al., 2010).

203

The continued exposure of ducks to oil residues from the Exxon Valdez is disputed. In 1995, the
effects of naturally weathered EVCO were tested on mallard ducks (Anas platyrhynchos), with
deleterious effects noted only at the highest concentrations. This indicated that weathered oil was
substantially less toxic to mallard ducks and their developing embryos than unweathered oil
(Stubblefield et al., 1995). After considering the PWS ecosystem as a whole, Wiens (2007)
concluded that it was highly unlikely that residual oil from the EVOS was continuing to have an
effect on harlequin ducks through the food web (ingestion of contaminated mussels). In
addition, given the possibility that only a small percentage of PWS shoreline contains residual
oil, the likelihood of wildlife coming in contact with the oil is remote (Neff et al., 2011). Finally,
other researchers have noted that the induction of the CYP1A gene in harlequin ducks and sea
otters can occur through exposure to other contaminants such as polychlorinated biphenols, and
not just oil residues (Huggett et al., 2006; Neff et al., 2011; Ricca et al., 2010). This assertion
has been counter-balanced by (Short et al., 2008), who found that the only source of CYP1A-
inducing contaminants in rainbow trout was from oil residues remaining after the EVOS.

Marine Mammals

A tally of mortalities from the EVOS on marine mammals yielded a body count of 1011 sea
otters, 19 harbour seals, 12 Steller sea lions, and 37 cetaceans. The numbers were considered
misleading due to the fact that, in all likelihood, not all victims were found, and not all died due
to the effects of EVCO (St. Aubin and Geraci, 1994). Other estimates suggested that up to 2800
sea otters and at least 302 harbour seals died in PWS (Peterson, 2001). During the few years
following the spill, the deleterious effects of EVCO on marine mammals were primarily due to
direct oiling, inhalation of fumes, or direct ingestion of oil rather than any prolonged exposure
through feeding on contaminated food (von Ziegesar et al., 1994).

Continued long-term effects of EVCO residues have been most closely followed in sea otters
(Enhydra lutris). Similar to harlequin ducks, sea otters can come into contact with buried oil
through foraging and exposure to contaminated prey (Peterson, 2001; Short et al., 2006). As
with harlequin ducks, the principal biomarker is CYP1A gene expression (Bodkin et al., 2002).
CYP1A induction was observed in sea otters until 1998, in contrast to the expression of the
204

CYP1A gene as EROD activity in harlequin ducks as late as 2009. The long-term effects are
questioned by a number of scientists who contend that the incidence of residual oil on the shores
of PWS is low, and consequently the likelihood of sea otters contacting residual oil or
contaminated food is also low (Boehm et al., 2007b; Neff et al., 2011).

Effects on cetaceans (whales, dolphins and porpoises) were difficult to document directly. There
was no observable effect from the EVOS on the population of humpback whales (von Ziegesar et
al., 1994). In the case of killer whales (Orcinus orca), they did not avoid the oiled areas or avoid
surfacing in oil slicks resulting in the likelihood for contact with surface oil and the consumption
of oiled prey (Matkin et al., 1994). While it was noticed after the spill that 14 members of one
pod were missing, death directly attributable to exposure to EVCO could not be proven (St.
Aubin and Geraci, 1994).

The only cetacean group that continued to be monitored over the long-term was the killer whale
pods studied immediately post-spill (Matkin et al., 2008). Two killer whale pods were tracked
five and then 16 years after the spill. During the 16 years, neither pod had returned to their pre-
spill populations. One pod, (identified as AT1, which lost nine members) is feared to be on the
brink of extinction due to the loss of reproductive females. The other (AB) appears to be
recovering, but at a slow rate.

9.3 Present Status of Injured Resources and Services

The 2010 update on injured resources and services (EVOS Trustee Council, 2010) provides the
current status of resources and services affected by the EVOS. The list is not comprehensive, but
gives an account of selected items that were identified for the Restoration Plan, and was used for
directing public restoration funds (Table 12). Although the recovery goal is for a resource to be
ameliorated to a condition that it would have been had the spill not occurred, this might not be
possible because of changes in the environment that have occurred irrespective of the spill.


205

The categories for recovery status are as follows and are taken directly from the EVOS Trustee
Council (2010) publication.
Not Recovering: Resources that are Not Recovering continue to show little or no clear
improvement from injuries stemming from the oil spill. Recovery objectives have not been
met.
Recovering: Recovering resources are demonstrating substantive progress toward recovery
objectives, but are still adversely affected by residual impacts of the spill or are currently
being exposed to lingering oil. The amount of progress and time needed to attain full
recovery varies depending on the species.
Recovered: Recovery objectives have been met, and the current condition of the resource is
not related to residual effects of the oil spill.
Very Likely Recovered: While there has been limited scientific research on the recovery
status of these resources in recent years, prior studies suggest that there had been substantial
progress toward recovery in the decade following the spill. In addition so much time has
passed since any indications of some spill injury, including exposure to oil, it is unlikely that
there are any residual effects of the spill.
Recovery Unknown: For resources in the unknown category, data on life history or the
extent of injury from the spill is limited. Moreover, given the length of time since the spill, it
is unclear if new or further research will provide information that will help in
comprehensively assessing the original injury of determining the residual effects of the spill
such that a better evaluation of recovery can occur.

The recovery objectives are described as, Specific, measurable parameters that, when achieved,
signal the recovery of an injured resource or service.
206


Table 12 Status in 2010 of resources and services injured by the Exxon Valdez oil spill in 1989. Human
services are those which were negatively impacted because of their connection with impacted resources
(EVOS Trustee Council, 2010).
RESOURCE STATUS
Archaeological Resources Recovered
Bald Eagles Recovered
Barrow's Goldeneye Recovering
Black Oystercatchers Recovering
Clams Recovering
Common Loons Recovered
Common Murres Recovered
Cormorants Recovered
Cuffhroat Trout Very likely recovered
Designated Wilderness Recovering
Dolly Varden Recovered
Harbour Seals Recovered
Harlequin Ducks Recovering
lntertidal Communities Recovering
Killer Whales resident AB pod Recovering
Killer Whales AT1 population Not recovering
Kittlitz's Murrelets Unknown
Marbled Murrelets Unknown
Mussels Recovering
Pacific Herring Not recovering
Pigeon Guillemots Not recovering
Pink Salmon Recovered
River Otters Recovered
Rockfish Very likely recovered
Sea Otters Recovering
Sediments Recovering
Sockeye Salmon Recovered
Subtidal Communities Very likely recovered
HUMAN SERVICES
Commercial Fishing Recovering
Passive Use Recovering
Recreation and Tourism Recovering
Subsistence Recovering


207

9.4 Lessons Learned and Issues to Consider

It appears that the opinions of scientists representing the public sector and government on the
recovery of the PWS ecosystem diverged quickly from those held by industry stakeholders soon
after the spill. While all parties acknowledged the immediate, acute effects on many species of
birds, sea mammals, invertebrates and shoreline flora, one party concluded that effects were
inconsequential anywhere from a few months to a few years post-spill while the other concluded
that the effects had shifted from being acute (causing mortalities), to chronic, long-term,
sublethal effects which persisted decades after the spill. This difference in opinion is largely the
result of
differences in the definitions and monitoring endpoints for PWS ecosystem recovery
(Parker and Maki, 2003; Parker and Wiens, 2005; Skalski et al., 2001);
incomplete baseline studies of the ecosystem prior to oil exploration;
natural variation, for example Eckert (2009) noted an average temporal variability of
85% (1 to 447%) in a time series spanning more than 2 years of 226 species (birds,
mammals, fish, algae and invertebrates) in the Gulf of Alaska from sites that had never
been oiled, which makes it difficult to assess impacts of the EVOS;
PAH sources in PWS in addition to the EVOS (Deepthike et al., 2009; Page et al., 1996;
Page et al., 2010); and
normative science, which is the application of scientific results in a piecemeal and
selective manner to support pre-existing policy rather than using the whole suite of
results to develop new policy.

208

10. Future Research Needs

There are a number of identified areas for future research to improve our understanding of the
fate and effects of oil spills in the Arctic and the efficacy of oil spill response protocols.

10.1 Oil Detection

Effective application of oil spill countermeasures is dependent on our ability to detect and track
oil both in and under ice cover. There is a need to improve methodologies to locate and assess
extent of oil within an ice field to improve on simple daylight visual capability. Proposed
programs for sensor development and validation include technologies such as laser fluorosensor
for exposed spills between floes and on ice, and ground-penetrating radar for oil under ice,
including rough ice.


10.2 Oil Fate and Behaviour

Due to the recent blowout in the Gulf of Mexico, and the intent for deep-sea recovery of
petroleum resources within the Arctic there is much interest to improve our knowledge on the
fate and behaviour of oil released from a sub-sea blowout. Furthermore, higher resolution
oceanographic meta-data (metocean data) is now required for the development of integrated oil
spill trajectory models to support Arctic operations. There is also a need to understand the
influence of environmental variables such as the presence and the influence of various levels of
ice coverage on drifting oil and the interaction of oil with suspended particulate material to form
OMA. The development of predictive operational models relies upon the quality of available
data on spreading, advection, evaporation, emulsification, natural dispersion, oil-mineral particle
interaction, dissolution, and biodegradation.


209

10.3 Biological Effects

For comparative purposes and to support future natural resource damage assessment procedures,
there is a need to develop standardized protocols for toxicity testing. Risk assessments will
require a toxicity database for valued ecosystem components (VECs). There is a need to conduct
pre-production assessments of native populations and habitat characteristics to establish baseline
conditions at future sites slated for exploration and production. In terms of environmental effects
monitoring, our focus should be shifted from acute to sublethal effect studies focusing on key
polar species and features including sensitive life stages. There is an identified need for studies
on temperature and its affect on the metabolism of cold-blooded organisms which may alter
biochemical responses, and studies on the life histories of Arctic organisms to determine
differences in the timing of exposure to oil. The solar regime in the high Arctic supports
investigation on the influence of photodegradation processes on oil toxicity. While oil spill
response protocols have been developed for the Arctic, there is a need to understand the
consequences of the technologies such as oil spill treatment agents. Integration of toxicity data
and the development of predictive risk assessment models are essential for the development and
application of predictive risk assessment models.

10.4 Mechanical Recovery

Improvements in mechanical recovery are highly dependent on improving the oil encounter
rate. Emphasis has been on improving skimmer effectiveness in cold, icy conditions, the
development of tactics for shallow water and overflood conditions, the use of chemical herders,
and oil deflection techniques. Field trials are required to test mechanical recovery systems in a
full-scale ice field with representative thick ice, including effects of vessel-ice interactions, and
weather conditions. There is also a need to find a solution for the disposal of recovered oily
wastes.


210

10.5 In Situ Burning

In situ burning has now been validiated to be an operational oil spill response following its recent
demonstration in the Gulf of Mexcico. There is a need for the development of operational
guidelines for future use of the technology. Several reseach priority areas have been identified
including
the testing of chemical herders to enhance ignitability of oil in loose pack ice conditions;
improvement of fire boom reliability;
investigation of the fate and toxicity of burn residues to marine species; and
the improvement of ignition systems (e.g. plasma torch).

10.6 Enhanced Dispersion

Dispersion of spilled oil and ISB are considered the spill response countermeasures holding the
most promise of success under Arctic conditions. There is a need to develop operational
guidelines for traditional product formulations and methods currently in use, and to assess the
efficacy of improved cold water or low salinity chemical dispersant formulations. Research,
involving full-scale field trials that include tracking and monitoring techniques, is needed to
elucidate the long-term fate and effects of chemically dispersed oil. Sub-surface injection of
chemical oil dispersants at the wellhead requires the development of application systems. The
significance of OMA formation in spill response operations both in open water, within ice floes,
and on shoreline environments (such as surf-washing) requires further study.

10.7 Biodegradation and Natural Attenuation

There is a need to fully understand the chemical and biological processes controlling the rates of
natural and enhanced recovery associated with microbial degradation of oil, with a view towards
developing in situ methods of biodegradation with minimal mechanical containment and
recovery of oily wastes, their transportation, or disposal. There is a need for studies using
advanced genomic techniques to assess the significance of natural oil degradation processes in
211

the Arctic marine environment. Our level of knowledge concerning the diversity of oil catabolic
genes and levels of expression in Arctic marine systems is practically nil.

10.8 Development of Predictive Models

Improvements in predictive modelling are required to understand the fate, transport and effects
of oil released in the Arctic to support future risk assessment needs. The models will require
real-time high resolution meta-data of oceanographic properties to provide emergency response
support.


10.9 Field Trials
The consensus of the international scientific community is that field trials are needed to advance
our development of currently emerging oil spill countermeasures for use in the Arctic such as
ISB, chemical oil dispersants, herders and enhanced OMA dispersion. It is recommended that
field trials be conducted under varying spill scenarios and environmental conditions, particularly
spill thickness and volume, blowout compared to batch-type release, tanker release, sub-sea
pipeline release. Experimental parameters for ice conditions should include ice thickness,
concentration, brash ice, marginal ice, dynamic pack ice, sea state, shallow and deep water.
There is also an interest to evaluate the use of azimuth drive icebreakers to provide mixing
energy and to control the spreading of oil in ice fields. Field trials will also provide a means for
the training of oil spill personnel, an opportunity to test various surveillance techniques and to
validate oil spill trajectory models.

212

11. Acknowledgements

The authors thank Dr. Lionel Camus for providing recent sources of information from his
laboratory (Akvaplan niva, Norway), on the sensitivity of Arctic marine organisms to petroleum
hydrocarbons. The views expressed in this report are those of the authors and do not necessarily
represent the opinions or positions of the National Energy Board of Canada which sponsored this
independent review.

213

References

Aas, E., B. Liewenborg, B. E. Grsvik, L. Campus, G. Jonsson, J. F. Brseth and L. Balk. 2003.
DNA adduct levels in fish from pristine areas are not detectable or low when analysed
using the nuclease P1 version of the
32
P-postlabelling technique. Biomarkers 8: 445-460.
ACIA. 2004. Impacts of a Warming Arctic: Arctic Climate Impact Assessment. Cambridge
University Press. 140 pp.
Albers, P. H. 2002. Petroleum and Individual Polycyclic Aromatic Hydrocarbons. In D. J.
Hoffman, B. A. Rattner, G.A. Burton Jr. and J. Cairns Jr. (eds), Petroleum and Individual
Polycyclic Aromatic Hydrocarbons. Lewis Publishers/CRC Press, Boca Raton, FL. pp
341-372.
Allen, A. A. 1978. Case study: oil recovery beneath ice. In Proceedings of the 10th Annual
Offshore Technology Conference, Dallas, TX. pp. 261-266.
AMAP. 2008. Arctic Oil and Gas 2007. Arctic Monitoring and Assessment Program (AMAP),
Oslo, Norway.
American Petroleum Institute. 2009. Analysis of U.S. Spillage. API Publication No. 356.
American Petroleum Institute, Washington, DC. 71 pp.
Anderson, M. R., R. B. Rivkin and P. Warren. 2000. The influence of produced water on natural
populations of marine bacteria. In Proceedings of the 27th Annual Toxicity Workshop.
Canadian Technical Report on Fisheries and Aquatic Sciences, 2331:91-98.
Andrews, C. D., J. Payne, L. Fancey, J. Hanlon and K. Lee. 2007. Potential of produced water
discharges to affect fish health. In Proceedings of the International Produced Water
Conference, Environmental Risks and Advances in Mitigation Technologies, October 17-
18, 2007. St. Johns, Newfoundland, Canada p.43.
ArcOD. 2010. Census of Marine Life. Arctic Ocean Diversity. University of Rhode Island,
Office of Marine Programs. [Online]. Available: http://www.coml.org/ [April 18, 2011].
Armstrong, D. A., P. A. Dinnel, J. M. Orensanz, J. L. Armstrong, T. L. McDonald, R. F.
Cusimano, R. S. Nemeth, M. L. Landolt, J. R. Skalski, R. F. Lee and R. J. Huggett. 1995.
Status of selected groundfish and crustacean species in Prince William Sound following
the Exxon Valdez oil spill. American Society for Testing and Materials (ASTM), Ann
Arbor, MI.
Arndt, C. E. and K. M. Swadling. 2006. Crustacea in Arctic and Antarctic sea ice: distribution,
diet and life history strategies. Advances in Marine Biology 51: 213-232.
ASTM International. 2006. Standard test method for API gravity of crude petroleum and
petroleum products (Hydrometer Method). ASTM International Report D287-92, West
Conshohocken, PA. 3 pp.
Atlas, R. M. 1981. Microbial degradation of petroleum hydrocarbons: an environmental
perspective. Microbiological Reviews 45: 180-209.
Atlas, R. M. 1984. Petroleum Microbiology. MacMillan Publishing Company.
Atlas, R. M. 1995. Petroleum biodegradation and oil spill bioremediation. Marine Pollution
Bulletin 31: 178-182.
Atlas, R. M. and R. Bartha. 1972. Degradation and mineralization of petroleum in seawater:
Limitation by nitrogen and phosphorus. Biotechnology and Bioengineering 14: 309-317.
Atlas, R. M. and R. Bartha. 1992. Hydrocarbon biodegradation and oil-spill bioremediation.
Advances in Microbial Ecology 12: 287-338.
214

Atlas, R. M., E. A. Schofield, F. A. Morelli and R. E. Cameron. 1976. Effects of petroleum
pollutants on Arctic microbial populations. Environmental Pollution 10: 35-43.
Aunaas, T., A. Olsen and K. E. Zachariassen. 1991. The effects of oil and oil dispersants on the
amphipod Gammarus oceanicus from arctic waters. Polar Research 10: 619-630.
Aurand, D. and G. Coelho. 2005. Cooperative Aquatic Toxicity Testing of Dispersed Oil and the
Chemical Response to Oil Spills: Ecological Effects Research Forum (CROSERF).
Ecosystem Management & Associates, Inc. Lusby, MD. Technical Report 07-03.
Aydin, K. Y., V. V. Lapko, V. I. Radchenko and P. A. Livingston. 2002. A Comparison of the
Eastern Bering and Western Bering Sea Shelf and Slope Ecosystems Through the Use of
Mass-Balance Food Web Models. U.S. Department of Commerse. 78 pp.
Ayers, R. C. and M. Parker. 2001. Produced Water Waste Management. Canadian Association of
Petroleum Producers (CAPP). Calgary, Alberta, Canada.
Azetsu-Scott, K., P. Yeats, G. Wohlgeschaffen, J. Dalziel, S. Niven and K. Lee. 2007.
Precipitation of heavy metals in produced water: Influence on contaminant transport and
toxicity. Marine Environmental Research 63: 146-167.
Bado-Nilles, A., B. Gagnaire, H. Thomas-Guyon, S. L. Floch and T. Renault. 2008. Effects of 16
pure hydrocarbons and two oils on haemocyte and haemolymphatic parameters in the
Pacific oyster, Crassostrea gigas (Thunberg). Toxicology in Vitro 22: 1610-1617.
Baklien, ., R. Lange and L.-O. Reiersen. 1986. A comparison between the physiological effects
in fish exposed to lethal and sublethal concentrations of a dispersant and dispersed oil.
Marine Environmental Research 19: 1-11.
Baror, Y. 1990. The effect of adhesion on survival and growth of microorganisms. Experientia
46: 823-826.
Barron, M. G., M. G. Carls, R. Heintz and S. D. Rice. 2004. Evaluation of fish early life-stage
toxicity models of chronic embryonic exposures to complex polycyclic aromatic
hydrocarbon mixtures. Toxicological Sciences 78: 60-67.
Barron, M. G., M. G. Carls, J. W. Short and S. D. Rice. 2003. Phototoxicity 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.
Bartha, R. 1986. Biotechnology of petroleum pollutant biodegradation. Microbial Ecology 12:
155-172.
Bassin, N. J. and T. Ichiye. 1977. Flocculation behavior of suspended sediments and oil
emulsions. Journal of Sedimentary Petrology 47: 671-677.
Bech, C. and P. Sveum. 1991. Spreading of oil in snow: a field experiment. In Proceedings of the
14th Arctic and Marine Oilspill Program (AMOP) Technical Seminar. Environment
Canada, Ottawa, Ontario: 57-71.
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: 9-29.
Belore, R. C., K. Trudel, J. V. Mullin and A. Guarino. 2009. Large-scale cold water dispersant
effectiveness experiments with Alaskan crude oils and Corexit 9500 and 9527
dispersants. Marine Pollution Bulletin 58: 118-128.
Berge, J., G. Johnsen, F. Nilsen, B. Gulliksen and D. Slagstad. 2005. Ocean temperature
oscillations enable reappearance of blue mussels Mytilus edulis in Svalbard after a 1000
year absence. Marine Ecology Progress Series 303: 167-175.
215

Bertrand, J. C., M. Almallah, M. Acquaviva and G. Mille. 1990. Biodegradation of hydrocarbons
by an extremely halophilic archaebacterium. Letters in Applied Microbiology 11: 260-
263.
Bird, K. J., R. R. Charpentier, D. L. Gautier, D. W. Houseknecht, T. R. Klett, J. K. Pitman, T. E.
Moore, C. J. Schenk, M. E. Tennyson and C. J. Wandrey. 2008. Circum-Arctic Resource
Appraisal: Estimates of Undiscovered Oil and Gas North of the Arctic Circle. United
States Geological Survey, USGS Fact Sheet 3049.
Birtwell, I. K., R. Fink, D. Brand, R. Alexander 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.
Bjerkemo, O. K. 2011. Behavior of oil and HNS spilled in Arctic waters - an Arctic Council
Project. In Proceedings of the 2011 International Oil Spill Conference, Portland OR,
USA. 12 pp.
Blanchard, A. L., H. Nichols and C. Parris. 2010. 2008 Environmental Studies Program in the
Chukchi Sea: Benthic Ecology of the Burger and Klondike Survey areas. ConocoPhillips
Alaska, Inc. and Shell Exploration & Production Company, Anchorage, Alaska.
Blenkinsopp, S. A., G. Sergy, K. Doe, G. Wohlgeschaffen, K. Li. and M. Fingas. 1996. Toxicity
of the weathered crude oil used at the Newfoundland Offshore Burn Experiment (NOBE)
and the resultant burn residue. Spill Science and Technology Bulletin 3: 277-280.
Blouin, M. and K. Lee. 2007. Oil Mineral Aggregate Formation in Ice-Infested Waters. In
Proceedings of the 2007 International Oil & Ice Workshop. Minerals Management
Service, Herndon, VA.
Bluhm, B. A., I. R. MacDonald and C. Debenham. 2005. Macro- and megabenthic communities
in the high Arctic Canada Basin: Initial findings. Polar Biology 28: 218-231.
Bobra, A. M. and M. F. Fingas. 1986. The behaviour and fate of Arctic oil spills. Water Science
and Technology 18: 13-23.
Bobra, M. 1992. A Study of the Evaporation of Petroleum Oils, A Study of the Evaporation of
Petroleum Oils. Consultchem, Ottawa.
Bodkin, J. L., B. E. Ballachey, T. A. Dean, A. K. Fukuyama, S. C. Jewett, McDonald, M. L.,
D.H., C. E. OClair and G. R. VanBlaricom. 2002. Sea otter population status and the
process of recovery from the 1989 Exxon Valdez oil spill. Marine Ecology Progress
Series 241: 237-253.
Boehm, P. D., J. M. Neff and D. S. Page. 2007a. Assessment of polycyclic aromatic hydrocarbon
exposure in the waters of Prince William Sound after the Exxon Valdez oil spill: 1989-
2005. Marine Pollution Bulletin 54: 339-356.
Boehm, P. D., D. S. Page, J. S. Brown, N. J.M., J. R. Bragg and R. M. Atlas. 2008. Distribution
and weathering of crude oil residues on shorelines 18 years after the Exxon Valdez spill.
Environmental Science & Technology 42: 9210-9216.
Boehm, P. D., D. S. Page, J. S. Brown, J. M. Neff and W. A. Burns. 2004. Polycyclic aromatic
hydrocarbon levels in mussels from Prince William Sound, Alaska, USA. Environmental
Toxicology and Chemistry 23: 2916-2929.
Boehm, P. D., D. S. Page, J. M. Neff, E. Gundlach and J. S. Brown. 2011. Long-term Fate and
Persistence of Oil from the Exxon Valdez Oil Spill: Lessons Learned or History Repeated
(Poster). In 2011 International Oil Spill Conference, Portland, OR, USA. 1pp.
216

Boehm, P. D., D. S. Page, J. M. Neff and C. B. Johnson. 2007b. Potential for sea otter exposure
to remnants of buried oil from the Exxon Valdez oil spill. Environmental Science &
Technology 41: 6860-6867.
Boehm, P. D., M. S. Steinhauer, B. F. D.R. Green, B. Humphrey, D. L. Fiest and W. J. Cretney.
1987. Comparative fate of chemically dispersed and beached crude oil in subtidal
sediments of the arctic nearshore. Arctic 40: 133-148.
Bomba, J. G. and R. J. Brown. 2011. Well, you've drilled 'em. Now can you produce em?
Panarctic Oils Ltds test program - the Drake F76 Project. In Proceedings of the 2011
Arctic Technology Conference, Houston TX Feb 7-9, 10 pp.
Brseth, J. F. and K.-E. Tollefsen. 2004. Water column monitoring 2003. Summary Report RF-
Akvali Report RF-63319.
Bothamley, M. 2004. Offshore processing options vary widely. Oil and Gas Journal 102: 47-55.
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 spill. Ground Water
Monitoring and Remediation 31: 59-67.
Boufadel, M. C., P. Reeser, M. T. Suidan, B. A. Wrenn, J. Cheng, X. Du, T. H. L. Huang and A.
D. Venosa. 1999. Optimal nitrate concentration for the biodegradation of n-heptadecane
in a variably-saturated sand column. Environmental Technology 20: 191-199.
Boufadel, M. C., Y. Sharifi, B. Van Aken, B. A. Wrenn 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.
Bowman, J. and J. Deming. 2010. Elevated bacterial abundance and exopolymers in saline frost
flowers and implications for atmospheric chemistry and microbial dispersal. Geophysical
Research Letters 37: L13501, 5 pp.
Boyd, J. N., J. H. Kucklick, D. K. Scholz, A. H. Walker, R. G. Pond and A. Bostrom. 2001.
Effects of Oil and Chemically Dispersed Oil in the Environment. Health and
Environmental Sciences Department Cape Charles, Virginia.
Bradstreet, M. S. W. and W. E. Cross. 1982. Trophic relationships at High Arctic ice edges.
Arctic 35: 1-12.
Bragg, J. R., R. C. Prince, E. J. Harner and R. M. Atlas. 1994. Effectiveness of bioremediation
for the Exxon Valdez oil spill. Nature 368: 413-418.
Bragg, J. R. and S. H. Yang. 1995. Clay-oil flocculation and its effects on the rate of natural
cleansing in Prince William Sound following the Exxon Valdez oil spill. In Wells, P.G.,
Butler, J.N. and Hughes, J.S. (eds.), Exxon Valdez Oil Spill - Fate and Effects in Alaskan
Waters. American Society for Testing and Materials, Philadelphia, PA.
Brakstad, O. G. and K. Bonaunet. 2006. Biodegradation of petroleum hydrocarbons in seawater
at low temperatures (0-5 degrees C) and bacterial communities associated with
degradation. Biodegradation 17: 71-82.
Brakstad, O. G., K. Bonaunet, T. Nordtug and O. Johansen. 2004. Biotransformation and
dissolution of petroleum hydrocarbons in natural flowing seawater at low temperature.
Biodegradation 15: 337-346.
Brakstad, O. G., I. Nonstad, L.-G. Faksness and P. J. Brandvik. 2008. Responses of microbial
communities in Arctic sea ice after contamination by crude petroleum oil. Microbial
Ecology 55: 540-552.
Brandsma, M. G. and J. P. Smith. 1996. Dispersion modeling perspectives on the environmental
fate of produced water discharges. In: M. Reed and S. Johnsen, Eds., Produced Water 2.
217

Environmental Issues and Mitigation Technologies. Environmental Science Research.
Volume 52. Plenum Press, New York, pp. 215-224.
Brandvik, P. J. 2007. Behaviour and weathering of oil spills under Arctic conditions and
implications for response. In Proceedings of the 2007 International Oil & Ice Workshop.
Minerals Management Service, Herndon, VA.
Brandvik, P. J., P. S. Daling and J. L. Myrhaug. 2010a. Mapping weathering properties as a
function of ice conditions: A combined approach using a flume basin verified by large
scale field experiments. In Proceedings of the 33rd Arctic and Marine Oilspill Program
(AMOP) Technical Seminar. Environment Canada, Ottawa, Ontario, pp 701-724.
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.
Brandvik, P. J., L.-G. Faskness, D. Dickins and J. Bradford. 2006a. Weathering of Oil Spills
Under Arctic Conditions: Field Experiments with Different Ice Conditions Followed by
In-situ Burning. Presented at the Third Annual NATO/CCMS Oil Spill Response
Workshop In, Dartmouth, NS, Canada.
Brandvik, P. J. and F. Leirvik. 2008. Weathering Properties of the Trestakk Oil. SINTEF
Materials and Chemistry Report, Project no. 800849, Trondheim, Norway.
Brandvik, P. J., J. F. Rasmussen and N. R. Bosberg. 2010b. Predicting ignitability of oil spills as
a function of oil type and weathering. Laboratory studies verified by field experiments. In
Proceedings of the 33rd Arctic and Marine Oilspill Program (AMOP) Technical
Seminar. Halifax, Canada. Environment Canada, Ottawa, Ontario.
Brandvik, P. J., J. F. Rasmussen, R. Danilof and F. Leirvik. 2010c. A laboratory burning cell
used to quantify in situ ignitability of oils as a function of oil type and weathering. In
Proceedings of the 33rd Arctic Marine Oilspill Program (AMOP) Technical Seminar.
Halifax, Canada. Environment Canada, Ottawa, Ontario. in press.
Brandvik, P. J., K. R. Sorheim, I. Singsaas and M. Reed. 2006b. Short State-of-the-Art Report on
Oil Spills in Ice-Infested Waters. Oil In Ice JIP Report no.1, SINTEF Materials and
Chemistry, Trondheim, Norway. 63 pp.
Brannon, E. L., K. M. Collins, J. S. Brown, J. M. Neff, K. R. Parker 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., K. M. Collins, M. A. Cronin, L. L. Moulton, K. R. Parker and W. Wilson. 2007.
Risk of weathered residual Exxon Valdez oil to pink salmon embryos in Prince William
Sound. Environmental Toxicology and Chemistry 26: 780-786.
Brannon, E. L., A. W. Maki, L. L. Moulton and K. R. Parker. 2006b. Results from a sixteen year
study on the effects of oiling from the Exxon Valdez on adult pink salmon returns. Marine
Pollution Bulletin 52: 892-899.
Broje, V., G. Merrell, J. Taylor and M. Macrander. 2011. Comprehensive contingency planning
for Arctic offshore operations. In Proceedings of the 2011 International Oil Spill
Conference, Portland, OR, USA. 9 pp.
Brown, E. D., T. T. Baker, J. E. Hose, R. M. Kocan, G. D. Marty, M. D. McGurk, B. L. Norcross
and J. Short. 1996. Injury to the early life history stages of pacific herring in Prince
William Sound after the Exxon Valdez oil spill. In Proceedings of the 1996 Exxon Valdez
Oil Spill Symposium. American Fisheries Society, Bethesda, Maryland: 931.
218

Brown, J., L. Cook and P. D. Boehm. 2011. Monitoring Petroleum Hydrocarbons in the
Sediments of the Nearshore Beaufort Sea During Construction and Development of the
Northstar Production Island, 1999 through 2006 (Poster). In 2011 International Oil Spill
Conference Portland, OR, USA. 1 pp.
Brown, M. and J. Bowman. 2001. A molecular phylogenetic survey of sea-ice microbial
communities (SIMCO). FEMS Microbiology and Ecology 35: 267-275.
Brude, O. W., T. Nordtug, L. Sverdrup, . Johansen and A. Melby. 2010. Oppdatering av faglig
grunnlag for forvaltningsplanen for Barentshavet og omrdene utenfor Lofoten (HFB):
Konsekvenser av akutt utslipp for fisk. Det Norske Veritas, Hvik, Norway. 94 pp.
Buckley, J. R., B.R. De Lange Boom and E. M. Reimer. 1987. The physical oceanography of the
Cape Hatt Region, Eclipse Sound, N.W.T. Arctic 40: 20-33.
Buist, I. 2003. Window-of-opportunity for in situ burning. Spill Science and Technology Bulletin
8: 341-346.
Buist, I., R. Belore, A. Guarino, D. Hackenberg, D. Dickins and Z. Wang. 2009. Empirical
weathering properties of oil in ice and snow. In Proceedings of the 32nd Arctic and
Marine Oilspill Program (AMOP) Technical Seminar on Environmental Contamination
and Response, Vancouver, BC, 9-11 June. Environment Canada, Ottawa, ON: 67-107.
Buist, I., D. Dickins, L. Majors, K. Linderman, J. Mullin and C. Owens. 2003. Tests to determine
the limits to in situ burning of thin oil slicks in brash and frazil ice. In Proceedings of the
26th Arctic and Marine Oil Spill Program (AMOP) Technical Seminar, Victoria, BC, 10-
12 June. Environment Canada, Ottawa, ON: 629-648.
Buist, I. and T. Nedwed. 2011. Using herders for rapid in situ burning Of oil spills on open
water. In Proceedings of the 2011 International Oil Spill Conference, Portland, OR,
USA. 7 pp.
Buist, I., S. Potter and T. Nedwed. 2011. Herding agents to thicken oil spills in drifti ice for in
situ burning: New developments. In Proceedings of the 2011 International Oil Spill
Conference, Portland, OR, USA. 12 pp.
Buist, I., S. Potter, T. Nedwed 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, 4-8 May. IOSC, Savannah, GA: 673-680.
Buist, I., S. Potter, L. Zabilansky, P. Meyer and J. Mullin. 2006. Mid-scale test tank research on
using oil herding surfactants to thicken oil slicks in pack ice: an update. In Proceedings of
the 29th Arctic and Marine Oilspill Program (AMOP) Technical Seminar, Vancouver,
BC, 6-8 June. Environment Canada, Ottawa, ON: 691-709.
Buist, I., K. Trudel, J. Morrison and D. Aurand. 1995. Laboratory studies of the properties of in-
situ burn residues. In Proceedings of the 1995 International Oil Spill Conference, Long
Beach, CA, USA: 4245-4252.
Buist, I. A. and I. Bjerkelund. 1986. Oil in pack ice preliminary results of three experimental
spills. In Proceedings of 9th Arctic and Marine Oil Spill Program (AMOP) Technical
Seminar. Edmonton, Canada. Environment Canada, Ottawa, Ontario: 379-397.
Buist, I. A., S. Joyce and D. F. Dickens. 1987. Oil spill in leads: tank tests and modeling,
Ottawa, ON.
Buist, I. A., W. M. Pistruzak and D. F. Dickens. 1981. DOME Petroleums oil and gas undersea
study. In Proceedings of the 4th Arctic Marine Oilspill Program (AMOP) Technical
Seminar. Edmonton, Canada. Environment Canada, Ottawa, Ontario: 647-686.
219

Buist, I. A., S. G. Potter and D. F. Dickens. 1983. Fate and behaviour of water-in-oil emulsions
in ice. In Proceedings of 6th Arctic and Marine Oilspill Program (AMOP) Technical
Seminar. Edmonton, Canada. Environment Canada, Ottawa, Ontario: 263-279
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: 172-183

Bunch, J. N., and T. Cartier. 1984. Effects of Petroleum Releases on the Microheterotrophic
Flora of Arctic Sediments Effects After Two Years. Baffin Island Oil Spill (BIOS)
Project Report no. 83-5. Canadian Department of the Environment, Edmonton, AB.
Bunch, J. N., C. Bdard and T. Cartier. 1983. Abundance and Activity of Heterotrophic Marine
Bacteria in Selected Bays at Cape Hatt, N.W.T.: Effects of Oil Spills, 1981. Canadian
Manuscript Report of Fisheries and Aquatic Sciences no. 1708, Baffin Island Oil Spill
(BIOS) project. Fisheries and Aquatic Sciences Ste. Anne de Bellevue, Quebec.
Burns, W. A., P. J. Mankiewicz, A. E. Bence, D. S. Page and K. R. Parker. 1997. A Principal-
Component and Least-Squares method for allocating polycyclic aromatic hydrocarbons
in sediment to multiple sources. Environmental Toxicology and Chemistry 16: 1119-
1131.
Busdosh, M. 1981. Long-term effects of the water soluble fraction of Prudhoe Bay crude oil on
survival, movement and food search successs of the arctic amphipod Boeckosimus
(Onisimus ) affinis. Marine Environmental Research 5: 167-180.
Busdosh, M. and R. M. Atlas. 1977. Toxicity of oil slicks to Arctic amphipods. Arctic 30: 85-92.
Byford, D. C., P. R. Laskey and A. Lewis. 1984. Effect of low temperature and varying energy
input on the droplet size distribution of oils treated with dispersants. In Proceedings of
the 7th Annual Arctic and Marine Oilspill Program (AMOP) Technical Seminar.
Environment Canada, Ottawa, Ontario: 208-228
Cairn. 2011. Greenland, Discovering Hidden Value. Cairn Energy PLC 2009-2011. [Online].
Available:http://www.cairnenergy.com/operations/. [2011, June 28].
Camus, L., S. R. Birkely, M. B. Jones, J. F. Brseth, B. E. Grsvik, B. Gulliksen, O. J. Lnne, F.
Regoli and M. H. Depledge. 2003. Biomarker responses and PAH uptake in Mya
truncata following exposure to oil-contaminated sediment in an Arctic fjord (Svalbard).
The Science of the Total Environment 308: 221-234.
Camus, L., B. J.F., M. B. Jones, F. Regoli and M. Depledge. 2002a. Heart rate and total
oxyradical scavenging capacity of the Arctic spider crab, Hyas araneus, following
exposure to crude oil via sediment and injection. Aquatic Toxicology 61: 1-13.
Camus, L., M. B. Jones, J. F. Brseth, B. E. Grsvik, F. Regoli and M. H. Depledge. 2002b.
Total oxyradical scavenging capacity and cell membrane stability of haemo- cytes of the
Arctic scallop, Chlamys islandicus, following benzo()pyrene exposure. Marine
Environmental Research 54: 425-430.
Camus, L., G. Olsen and J. Nahrgang. 2010. Biomonitoring tools and risk assessment in the
Arctic. In Proceedings of the 33rd Arctic and Marine Oilspill Program (AMOP)
Technical Seminar on Environmental Contamination and Response, v. 2. Environment
Canada. Ottawa, ON. pp. 1253-1266.
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.
Camus, L., G. R. Olsen and M. Carroll. 2008. Arctic Ecotoxicological Studies Review. Akvaplan-
niva AS Report 3975. Akvaplan-niva AS, 9296 Troms, Norway. 54 pp.
220

Canfield, D. E., E. Kristensen and B. Thamdrup. 2005. Microbial Ecosystems. In Southward,
A.J., Tyler, P.A., Young, C.M. and Fuiman, L.A. (eds.), Advances in Marine Biology:
Aquatic Geomicrobiology. Elsevier Academic Press, San Diego, California 92101-4495,
USA.
Carls, M. G., M. M. Babcock, P. M. Harris, G. V. Irvine, J. A. Cusick and S. D. Rice. 2001a.
Persistence of oiling in mussel beds after the Exxon Valdez oil spill. Marine
Environmental Research 51: 167-190.
Carls, M. G., R. Heintz, A. Moles, S. D. Rice and J. W. Short. 2001b. Long-term biological
damage: What is known, and how should that influence decisions on response,
assessment, and restoration? In Proceedings of the 2001 International Oil Spill
Conference, Tampa, FL, USA: 399-403.
Carls, M. G., J. E. Hose, R. E. Thomas and S. D. Rice. 2000. Exposure of Pacific Herring to
weathered crude oil: Assessing effects on ova. Environmental Toxicology and Chemistry
19: 1649-1659.
Carls, M. G. and S. Korn. 1985. Sensitivity of Arctic Marine Amphipods and Fish to Petroleum
Hydrocarbons. Canadian Technical Report of Fisheries and Aquatic Sciences No. 1368:
11-26.
Carls, M. G., S. D. Rice and J. E. 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. and J. F. Thedinga. 2010. Exposure of pink salmon embryos to dissolved
polynuclear aromatic hydrocarbons delays development, prolonging vulnerability to
mechanical damage. Marine Environmental Research 69: 318-325.
Carroll, J., G. H. Olsen, M. Smit and H. B. Mostad. 2010. Species sensitivities to acute oil
exposure - are arctic and temperate dwelling organisms similar or different? - Arctic
species sensitivities to acute oil exposure- final report 2010. Accessed 18 Jun 2011:
http://www.vista.no/areas/environment/species-sensitivities-to-acute-oil-exposure-are-
arctic-and-temperate-dwelling-organisms-similar-or-different/reports/species-
sensitivities-to-acute-oil-exposure-are-arctic-and-temperate-dwelling-organisms-similar-
or-different-arctic-species2019-sensitivities-to-acute-oil-exposure-final-report-2010.
CBC. 2011. Greenland offshore drilling is safe: Cairn Energy. CBCnews, 25 May 2011.
[Online]. Available: http://www.cbc.ca/news/canada/north/story/2011/05/25/cairn-
energy-greenland-drilling.html. [2011, June 28].
CGPC. 2006. Natural Gas Potential in Canada 2005, Volume 3: Frontier Basins of Canada.
The Canadian Gas Potential Committee. [Online]. Available:
http://www.centreforenergy.com. [2011, June 27].
Chen, E. C., B. E. Keevil and R. O. Ramsevier. 1976. Behaviour of Oil Spilled in Ice Covered
Rivers. Scientific Series No.61, Environment Canada, Ottawa, ON.
Chen, Z., K. G. Osadetz, A. F. Embry, H. Gao and P. K. Hannigan. 2000. Petroleum potential in
western Sverdrup Basin, Canadian Arctic Archipelago. Bulletin of Canadian Petroleum
Geology 48: 323-338.
Christiansen, J., L. Karamushko and J. Nahrgang. 2010. Sub-lethal levels of waterborne
petroleum may depress routine metabolism in polar cod Boreogadus saida; (Lepechin,
1774). Polar Biology 33: 1049-1055.
221

Christiansen, J. S. and S. G. George. 1995. Contamination of food by crude oil affects food
selection and growth performance, but not appetite, in an Arctic fish, the polar cod
(Boreogadus saida). Polar Biology 15: 277-281.
Cloutier, D., C. L. Amos, P. R. Hill and K. Lee. 2002. Oil erosion in an annular flume by
seawater of varying turbidities: A critical bed shear stress approach. Spill Science &
Technology Bulletin 8: 83-93.
Cloutier, D. and B. Doyon. 2008. OMA Formation in Ice-Covered Brackish Waters: Large-Scale
Experiments. In W. F. Davidson, K. Lee and A. Cogswell (eds.), Oil Spill Response: A
Global Perspective. Springer. pp. 71-88.
Cloutier, D., S. Gharbi and B. Michel. 2005. On the oil-mineral aggregation process: A
promising response technology in ice-infested waters. In Proceedings of the 2005
International Oil Spill Conference, Miami Beach, FL, USA.
CNLOPB. 2010. Spill Statistics Summary (1997-2009). Canada-Newfoundland and Labrador
Offshore Petroleum Board. Accessed 18 Jul 2011:
www.cnlopb.nl.ca/pdfs/spill/sumtab.pdf
CNSOPB. 2010. Spills to the Sea. Canada-Nova Scotia Offshore Petroeum Board. Accessed 18
Jul 2011: www.cnsopb.ns.ca/environment_incident_statistics.php.
Colcomb, K., D. Bedborough, T. Lunel, R. Swannel, P. Wood, J. Rusin, N. Bailey, C. Halliwell,
L. Davis, M. Sommerville, A. Dobie, D. Michell, M. McDonagh, S. Shimwell, B. Davies,
D. Harries and K. Lee. 1997. Shoreline cleanup and waste disposal issues during the Sea
Empress Incident. In Proceedings of the 1997 International Oil Spill Conference, Fort
Lauderdale, FL, USA: 195-203.
Colcomb, K., M. Peddar, D. Salt and A. Lewis. 2005. Determination of the limiting oil viscosity
for chemical dispersion at sea. In Proceedings of the 20005 International Oil Spill
Conference, Miami Beach, FL, USA: 11506-11511.
Collier, T. K., C. A. Krone, M. M. Krahn, J. E. Stein, S.-L. Chan and U. Varanasi. 1996.
Petroleum exposure and associated biochemical effects in subtidal fish after the Exxon
Valdez oil spill. In Proceedings of the 1996 Exxon Valdez Oil Spill Symposium, Bethesad,
MD, USA.
Comfort, G. and W. Purves. 1982. The Behaviour of Crude Oil Spilled Under Multi-Year Ice.
Environment Canada, Ottawa, ON, Canada.
Connelly, T. L., C. M. Tilburg and P. L. Yager. 2006. Evidence for psychrophiles outnumbering
psychrotolerant marine bacteria in the springtime coastal Arctic. Limnology and
Oceanography 51: 1205-1210.
Costa, D. P. and G. L. Kooyman. 1982. Oxygen consumption, thermoregulation, and the effect
of fur oiling and washing on the sea otter, Enhydra lutris. Canadian Journal of Zoology
60: 2761-2767.
Cranford, P. J. 2006. Scallops and marine contaminants. In Shumway, S. and Parsons, J. (eds.),
Scallops: Biology, Ecology and Aquaculture (2nd ed.). Elsevier Applied Science,
London.
Cretney, W. J., D. R. Green, B. R. Fowler, B. Humphrey, F. R. Engelhardt, R. J. Norstrom, M.
Simon, D. L. Fiest and P. D. Boehm. 1987a. Hydrocarbon biogeochemical setting of the
Baffin Island Oil Spill experimental sites. III. Biota. Arctic 40: 71-79.
Cretney, W. J., D. R. GREEN, B. R. FOWLER, B. HUMPHREY, D. L. FIEST and P. BOEHM.
1987b. Hydrocarbon biogeochemical setting of the Baffin Island Oil Spill experimental
sites. I. Sediments. Arctic 40: 51-65.
222

Cretney, W. J., D. R. Green, B. R. Fowler, B. Humphrey, D. L. Fiest and P. D. Boehm. 1987c.
Hydrocarbon biogeochemical setting of the Baffin Island Oil Spill experimental sites. II.
Water. Arctic 40
66-70.
Cross, W. E. 1982a. In-situ studies of the effects of oil and dispersed oil on primary productivity
of ice algae and on under-ice amphipod communities. Special studies - 1981 study
results, Baffin Island Oil Spill Working Report 81-10. 61 pp.
Cross, W. E. 1982b. Under-ice biota at the pond inlet ice edge and in adjacent fast ice areas
during spring. Arctic 35: 13-27.
Cross, W. E. 1987. Effects of oil and chemically treated oil on primary productivity of high
Arctic ice algae studied in situ. Arctic 40: 266-276.
Cross, W. E. and B. Humphrey. 1987. Monitoring the Long-term Fate and Effects of Spilled Oil
in an Arctic Marine Subtidal Environment. Environmental Studies Research Funds
Report no. 075, Ottawa, ON. 120 pp.
Cross, W. E. and C. M. Martin. 1983. In-situ Studies of the Effects of Oil and Chemically Treated
Oil on Primary Productivity of Ice Algae and on Under-ice Microfauna and Macrofaunal
Communities. Special Studies - 1982 study results, Baffin Island Oil Spill Working
Report 82-7. 103 pp.
Cross, W. E. and C. M. Martin. 1987. Effects of oil and chemically treated oil on nearshore
under-ice meiofauna studied in situ. Arctic 40: 258-265.
Cross, W. E., C. M. Martin and D. H. Thomson. 1987a. Effects of experimental releases of oil
and dispersed oil on Arctic nearshore macrobenthos. II. Epibenthos. Arctic 40: 201-201.
Cross, W. E. and D. H. Thomson. 1987. Effects of experimental releases of oil and dispersed oil
on Arctic nearshore macrobenthos. I. Infauna. Arctic 40: 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: 211-219.
Daling, P. S., O. M. Aamo, A. Lewis and T. Strom-Kristiansen. 1997. SINTEF/IKU Oil-
weathering model: Predicting oil's properties at sea. In Proceedings of the 1997
International Oil Spill Conference, Fort Lauderdale, FL, USA: 297-307.
Darnis, G., D. G. Barber and L. Fortier. 2008. Sea ice and the onshore-offshore gradient I pre-
winter zooplankton assemblages in southern Beaufort Sea. Journal of Marine Systems 74:
994-1011.
Davis, J. E. and S. S. Anderson. 1976. Effects of oil pollution on breeding grey seals. Marine
Pollution Bulletin 7: 115-118.
Dean, T. A. and S. C. Jewett. 2001. Habitat-specific recovery of shallow subtidal communities
following the Exxon Valdez Oil Spill. Ecological Applications 11: 1456-1471.
Dean, T. A., M. S. Stekoll, S. C. Jewett, R. O. Smitha and J. E. Hose. 1998. Eelgrass (Zostera
marina L.) in Prince William Sound, Alaska: effects of the Exxon Valdez oil spill. Marine
Pollution Bulletin 36: 201-210.
DeBlois, E. M., C. Leeder, K. C. Penny, M. Murdoch, M. D. Paine, F. Power and U. P. Williams.
2005. Terra Nova Environmental Effects Monitoring Program: from Environmental
Impact Statement Onward. In S.L. Armworthy, P.J. Cranford and K. Lee (eds.), Offshore
Oil and Gas Environmental Effects Monitoring: Approaches and Technologies. Battelle
Press, Columbus, OH, USA, pp. 475-491.
223

Deepthike, H. U., R. Tecon, G. v. Kooten, J. R. v. d. Meer, H. Harms, M. Wells and J. Short.
2009. Unlike PAHs from Exxon Valdez crude oil, PAHs from Gulf of Alaska coals are
not readily bioavailable. Environmental Science & Technology 43: 5864-5870.
Delille, D., A. Bassres and A. Dessommes. 1997. Seasonal variation of bacteria in sea ice
contaminated by diesel fuel and dispersed crude oil. Microbial Ecology 33: 97-105.
Delille, D., A. Bassres and A. Dessommes. 1998. Effectiveness of bioremediation for oil-
polluted Antarctic seawater. Polar Biology 19: 237-241.
Delille, D. and B. Delille. 2000. Field observations on the variability of crude oil impact on
indigenous hydrocarbon-degrading bacteria from sub-Antarctic intertidal sediments.
Marine Environmental Research 49: 403-417.
Delille, D., E. Pelletier, A. Rodriguez-Blanco and J.-F. Ghiglione. 2009. Effects of nutrient and
temperature on degradation of petroleum hydrocarbons in sub-Antarctic coastal seawater.
Polar Biology 32: 1521-1528.
Delvigne, G. A. L. and C. E. Sweeney. 1988. Natural dispersion of oil. Oil and Chemical
Pollution 4: 281-310.
Delvigne, G. A. L., J. A. Van del Stel and C. E. Sweeney. 1987. Measurements of Vertical
Turbulent Dispersion and Diffusion of Oil Droplets and Oil Particles. US Department of
the Interior, Minerals Management Service, Anchorage, Alaska. 501 pp.
Deming, J. 2010. Sea ice bacteria and viruses. In D.N. Thomas and G.S. Dieckmann (eds) Sea
Ice - An Introduction to its Physics, Chemistry, Biology and Geology, 2nd Edition.
Blackwell Science Ltd., Oxford. pp. 247-282.
Deppe, U., H.-H. Richnow, W. Michaelis and G. Antranikian. 2005. Degradation of crude oil by
an arctic microbial consortium. Extremophiles 9: 461-470.
Desai, J. D. and I. M. Banat. 1997. Microbial production of surfactants and their commercial
potential. Microbiology and Molecular Biology Reviews 61: 47-&.
Deshpande, N., S. Chandrasekar, G. A. Sorial and J. W. Weaver. 2005. Dispersant effectiveness
on oil spills - impacts on environmental factors. MS, University of Cincinnati,
Engineering : Environmental Engineering, 2007CM2005/S:30 (Unpublished). pp 1-10.
Deslauriers, P. C. 1979. Observations of oil behaviour in ice floes and the 1977 Ethel H. spill. In
Proceeding of the 1979 Workshop on Oil, Ice, and Gas. University of Toronto, Institute
of Environmental Studies, Publication No. EE-14, pp.87-94.
Diaz, E. (ed) 2008. Microbial Biodegradation: Genomics and Molecular Biology. Caister
Academic Press, Norfolk, UK. 402 pp.
Dickens, D. F., J. Bradford and L. Steinborn. 2008. Detection of Oil on and Under Ice: Phase III
Evaluation of Airborne Radar System Capabilities in Selected Arctic Spill Scenarios,
Detection of Oil on and Under Ice: Phase III Evaluation of Airborne Radar System
Capabilities in Selected Arctic Spill Scenarios. Report 588, Mineral Management
Service, United States, 56p.
Dickins, D. F. 1987. Ice conditions at Cape Hatt, Baffin Island. Arctic 40: 34-41.
Dickins, D. F. 1994. Oil behaviour in ice-covered waters. In Proceedings of the Conference on
Oil Spill Response in Broken Ice. DF Dickins Assoc., La Jolla CA USA.
Dickins, D. F. 2011. Behavior of Oil Spills in Ice and Implications for Arctic Spill Response,
Behavior of Oil Spills in Ice and Implications for Arctic Spill Response. DF Dickins
Associates LLC. Prepared for the Arctic Technology Conference, Houston, TX, USA, 7-
9 Febfruary 2011.
224

Dickins, D. F., P. J. Brandvik, J. Bradford, L. G. Faksness, L. Liberty 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 in Proceedings of
the 2008 International Oil Spill Conference, Savannah, GA, USA: 681-688.
Dickins, D. F., G. Hearon, K. Morris, K. Ambrosius and W. Horowitz. 2011. Mapping sea ice
overflood using remote sensing: Alaskan Beaufort Sea. Cold Regions Science and
Technology 65: 275-285.
Dickins, D. F. and J. A. Hellebust. 1981. Return to Balaena Bay: Long Term Effects of a Large
Scale Crude Oil Spill Under Arctic Sea Ice. Gulf Canada Resources Inc. Calgary, AB. .
Dickins, D. F., E. Thornton and W. J. Cretney. 1987. Design and operation of oil discharge
systems and characteristics of oil used in the Baffin Island Oil Spill Project. Arctic 40:
100-108.
Dickins, D. F., K. Vaudrey and S. Potter. 2000. Oil Spills in Ice Discussion Paper, A Review of
Spill Response, Ice Conditions, Oil Behavior and Monitoring. Alaska Clean Seas,
Prudhoe Bay, AK
Dome Petroleum Ltd. 1981. Oil and Gas Under Sea Ice. Report CV-1, Canadian Offshore Oil
Spill Research Association (COOSRA), Vol. I and II, Calgary, AB, 286 p.
Driskell, W. B., A. K. Fukuyama, J. P. Houghton, D. C. Lees, A. J. Mearns and G. Shigenaka.
1996. Recovery of Prince William Sound intertidal infauna from Exxon Valdez oiling and
shoreline treatments, 1989 through 1992 In Proceedings of the 1996 Exxon Valdez Oil
Spill Symposium, Bethesda, MD, USA: 931.
Drolet, R., L. Fortier, D. Ponton and M. Gilbert. 1991. Production of fish larvae and their prey in
subarctic southeastern Hudson Bay. Marine Ecology Progress Series 77: 105-118.
Drozdowski, A., S. Nudds, C. G. Hannah and H. Niu. 2011. State of the Art Review of Oil Spill
Trajectory Modelling in the Presence of Ice. A report prepared for the National Energy
Board of Canada, June 2011, Dartmouth, NS, Canada. 75 pp.
Drummond, K. J. 2006. Canada's Discovered oil and gas resources North of 60*. Search and
discovery article #10102. Modified from extended abstract for presentation at AAPG
annual convention in Calgary, 2005., Calgary, AB, Canada.
Duesterloh, S., J. W. Short and M. G. Barron. 2002. Photoenhanced toxicity of weathered Alaska
North Slope crude oil to the calanoid copepods Calanus marchallae and Metridia
okhotensis. Environmental Science & Technology 36: 3953-3959.
Durell, G., S. Johnsen, T. R. Utvik, T. Frost and J. Neff. 2006. Oil well produced water
discharges to the North Sea. Part I: Comparison of deployed mussels (Mytilus edulis),
semi- permeable membrane devices, and the DREAM Model predictions to estimate the
dispersion of polycyclic aromatic hydrocarbons. Marine Environmental Research 62:
194-223.
Dutta, T. K. and S. Harayama. 2000. Fate of crude oil by the combination of photooxidation and
biodegradation. Environmental Science & Technology 34: 1500-1505.
Eckert, G. 2009. A synthesis of variability in nearshore Alaskan marine populations.
Environmental Monitoring and Assessment 155: 593-606.
Edmonton Journal. 2010. Interest in Baffin Bay oil high. Edmonton Journal, 10 April 2010.
Eimhjellen, K. and K. Josefson. 1984. Microbiology 2: Biodegradation of stranded oil - 1983
study results. Baffin Island Oil Spill Project Working Report 83-6. Ottawa:
Environmental Protection Service, Environment Canada. 58 pp.
225

Elasheva, O. M., T. N. Shabalina, K. M. Badyshtova, N. P. Meloshenko, V. A. Tyshchenko, I. N.
Smirnov and I. I. Zanozina. 2001. Rheological properties of Russian Crude Oil.
Chemistry and Technology of Fuels and Oils 37: 114-117.
Ellertsen, B., P. Fossum, P. Solemdal and S. Sundby. 1989. Relation between temperature and
survival of eggs and first-feeding larvae of northeast Arctic cod (Gadus morhua L.). In
Proceedings of the Meetings of the International Council for the Exploration of the Sea
(ICES): 209-219.
Energytoday. 2010. Canada to grant new offshore oil exploration leases in Beaufort Sea.
[Online]. Available: http://www.offshoreenergytoday.com/canada-to-grant-new-offshore-
oil-exploration-leases-in-beaufort-sea/. [2011, June 27].
Engelhardt, F. R. 1983. Petroleum effects on marine mammals. Aquatic Toxicology 4: 199-217.
Engelhardt, F. R. 1994. Limitations and innovations in the control of environmental impacts
from petroleum industry activities in the Arctic. Marine Pollution Bulletin 29: 334-341.
Engelhardt, F. R., J. R. Geraci and T. G. Smith. 1977. Uptake and clearance of petroleum
hydrocarbons in the ringed seal, Phoca hispida. Journal of the Fisheries Research Board
of Canada 34: 1143-1147.
Environment Canada. 2001a. Oil Properties Database. [Online]. Available: http://www.etc-
cte.ec.gc.ca/databases/oilproperties/. [2011, April 28].
Environment Canada. 2001b. Spilltox Database. [Online]. Available: http://www.etc-
cte.ec.gc.ca/databases/Spilltox/default.aspx . [2011, April 28].
Eriksson, M., J.-O. Ka and W. W. Mohn. 2001. Effects of low temperature and freeze-thaw
cycles on hydrocarbon biodegradation in Arctic tundra soil. Applied and Environmental
Microbiology 67: 5107-5112.
ESL Environmental Sciences Limited. 1982. The Biological Effects of Hydrocarbon Exploration
and Production Related Activities, Disturbances and Wastes on Marine Flora and Fauna
of the Beaufort Sea Region. Dome Petroleum Limited, Beaufort Environmental Impact
Statement, Calgary, AB, Canada.
Esler, D., T. D. Bowman, K. A. Trust, B. E. Ballachey, T. A. Dean, S. C. Jewett and C. E.
O'Clair. 2002. Harlequin duck population recovery following the Exxon Valdez oil spill:
progress, process, and constraints. Marine Ecology Progress Series 241: 271286.
Esler, D. and S. A. Iverson. 2010. Female harlequin duck winter survival 11 to 14 years after the
Exxon Valdez oil spill. Journal of Wildlife Management 74: 471-478.
Esler, D., K. A. Trust, B. E. Ballachey, S. A. Iverson, T. L. Lewis, D. J. Rizzolo, D.M.Mulcahy,
A. K. Miles, B. R. Woodin, J. J. Stegeman, J. D. Henderson and B. W. Wilson. 2010.
Cytochrome P4501A biomarker indication of oil exposure in Harlequin Ducks up to 20
years after the Exxon Valdez oil spill. Environmental Toxicology and Chemistry 29:
1138-1145.
EVOS Trustee Council. 2010. Exxon Valdez oil spill restoration plan 2010 update injured
resources and services. Exxon Valdez Oil Spill Trustee Council, Anchorage, AK. 46 pp.
Faksness, L.-G., J. F. Brseth, T. Baussant, B. H. Hansen, D. Altin, A. H. S. Tandberg, A.
Ingvarsdottir, N. Aarab and T. Nordtug. 2011. The Effects of Different Oil Spill Cleanup
Technologies on Body Burden and Biomarkers in Arctic Marine Organisms A
Laboratory Study. SINTEF Materials and Chemistry, Trondheim, Norway.
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.
226

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.
Faksness, L.-G., R. L. Daae, P. J. Brandvik, F. Leirvik and J. F. Brseth. 2010. Oil distribution
and bioavailability field experiment FEX 2009. Oil in Ice JIP Report no. 33, SINTEF
Materials and Chemistry Trondheim, Norway, 73 pp.
Faskness, L.-G. and P. Brandvik. 2005. Dissolution of Water Soluble Components from Oil
Spills Encapsulated in Ice In in Proceedings of the 28th Arctic and Marine Oilspill
Program (AMOP) Technical Seminar, Environment Canada, Ottawa, Ontario. pp. 59-73.
Feder, H. M., D. W. Norton and J. B. Geller. 2003. A review of apparent 20th century changes in
the presence of mussels (Mytilus trossulus) and macroalgae in Arctic Alaska, and of
historical and paleontological evidence used to relate mollusc distributions to climate
change. Arctic 56: 391-407.
Federal Interagency Solutions Group. 2010. Oil Budget Calculator: Deepwater Horizon,
November 2010. Federal Interagency Solutions Group, Oil Budget Calculator Science
and Engineering Team. Accessed 16 Jul 2011:
http://www.restorethegulf.gov/sites/default/files/documents/pdf/OilBudgetCalc_Full_HQ
-Print_111110.pdf.
Fingas, M. 1992. The behaviour of oil in ice. In Combatting Marine Oil Spills in Ice and Cold
Climates, HELLCOM Seminar, Helsinki, Finland.
Fingas, M. 1994. Evaporation of Oil Spills. Journal of the American Society of Civil Engineers
(in Press).
Fingas, M. 2005. Stability and Resurfacing of Dispersed Oil, Stability and Resurfacing of
Dispersed Oil. Prepared for Prince William Sound Regional Citizens' Advisory Council
(PWSRCAC). Environmental Technology Center, Environment Canada, Ottawa, ON.
Fingas, M. F. 1999. The evaporation of oil spills: Dvelopment and implementation of new
prediction methodology. In Proceedings of the 1999 International Oil Spill Conference,
Seattle, WA, USA.
Fingas, M. F. 2010. Review of the North Slope Oil Properties Relevant to Environmental
Assessment and Prediction. Prince William Sound Regional Citizens' Advisory Council,
Anchorage, Alaska.
Fingas, M. F., G. Halley, F. Ackerman, R. Nelson, M. Bissonnette, N. Laroche, Z. Wang, P.
Lambert, K. Li, P. Jokuty, G. Sergy, E. J. Tennyson, J. V. Mullin, L. Hannon, R. Turpin,
P. Campagna, W. Halley, J. Latour, R. Galarneau, B. Ryan, D. V. Aurand and R. R.
Hiltabrand. 1995. The Newfoundland Offshore Burn Experiment: NOBE In in
Proceedings of the 1995 International Oil Spill Conference, Long Beach, CA, USA:
123132.
Fingas, M. F. and B. P. Hollebone. 2003. Review of behaviour of oil in freezing environments.
Marine Pollution Bulletin 47: 333-340.
Finke, N., V. Vandieken and B. B. Jorgensen. 2007. Acetate, lactate, propionate, and isobutyrate
as electron donors for iron and sulfate reduction in Arctic marine sediments, Svalbard.
FEMS Microbiology Ecology 59: 10-22.
Fiocco, R. J., G. DeMarco, R. R. Lessard, P. S. Daling and G. P. Canevari. 1999. Chemical
dispersibility study of heavy bunker fuel oil In Environment Canada Arctic and Marine
Oil Spill Program Technical Seminar (AMOP) Proceedings: 173-186.
227

Forbes, D. L. e. 2011. State of the Arctic Coast 2010 Scientific Review and Outlook.
International Arctic Science Committee, Land-Ocean Interactions in the Coastal Zone,
Arctic Monitoring and Assessment Programme, International Permafrost Association.,
Helmholtz-Zentrum, Geesthacht, Germany, 178 pp. [Online]. Available:
http://arcticcoasts.org. [2011, June 2].
Forbes, V. E., M. S. H. Andreassen and L. Christensen. 2001. Metabolism of the polycyclic
aromatic hydrocarbon, fluoranthene, by the polychaete, Capitella capitata species I.
Environmental Toxicology and Chemistry 20: 1012-1021.
Frlin, L. and K. Hylland. 2006. Hepatic cytochrome P4501A concentration and activity in
Atlantic cod caged in two North Sea pollution gradients. In K. Hylland, T. Lang and D.
Vethaak (eds.), Biological Effects of Contaminants in Marine Pelagic Ecosystems.
SETAC Press, Pensacola, FL.
Forsyth, J. V., Y. M. Tsao and R. D. Blem. 1995. Bioremediation: when is augmentation
needed? In Hinchee, R.E. et al. (eds.), Bioaugmentation for Site Remediation. Battelle
Press, Columbus, OH, pp. 1-14.
Fraker, M. A. 1984. Balaena mysticetus: Whales, Oil and Whaling in the Arctic. Sohio Alaska
Petroleum Company and BP Alaska Exploration Inc.
Fraser, T. P. and M. Wicks III. 1995. Estimation of maximum stable oil droplet sizes at sea
resulting from natural dispersion and from use of a dispersant. In Proceedings of the 18th
Arctic and Marine Oilspill Program (AMOP) Technical Seminar. Environment Canada,
Ottawa, Ontario: 313-316.
Fretheim, A., E. McLanahan and S. Gudmundsdottir. 2011. Arctic Council Arctic Marine
Shipping Assessment (AMSA). In Proceedings of the 2011 International Oil Spill
Conference, Portland, OR, USA. 8 pp.
Friberg, S. E. 2007. Some emulsion features. Journal of Dispersion Science and Technology 28:
1299-1308.
Gamble, J. C., J. M. Davies, S. J. Hay and F. K. Dow. 1987. Mesocosm experiments on the
effects of produced water discharges from offshore oil platforms in the northern North
Sea. Sarsia 72: 383-386.
Garrett, R. M., I. J. Pickering, C. E. Haith and R. C. Prince. 1998. Photooxidation of crude oils.
Environmental Science & Technology 32: 3719-3723.
Gautier, D. L., K. J. Bird, R. R. Charpentier, A. Grantz, D. W. Houseknecht, T. R. Klett, T. E.
Moore, J. K. Pitman, C. J. Schenk, J. H. Schuenemeyer, K. Srensen, M. E. Tennyson, Z.
C. Valin and C. J. Wandrey. 2009. Assessment of undiscovered oil and gas in the Arctic.
Science 324: 1175-1179.
Geiger, H. J., B. G. Bue, S. Sharr, A. C. Wertheimer 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 In Proceedings of the 1996 Exxon Valdez Oil Spill Symposium.
American Fisheries Society, Bethesda, MD, USA: 931.
George, I., B. Stenuit and S. Agathos. 2010. Application of metagenomics to bioremediation. In
Diana Marco (ed.), Mathagenomics: Theory, Methods and Applications. Caister
Academic Press Norfolk, UK, 119 pp.
George, J. D. 1971. The effects of pollution by oil and oil-dispersants on the common intertidal
polychaetes, Cirriformia tentaculata and Cirratulus cirratus. Journal of Applied Ecology
8: 411-420.
228

George, S. G., J. S. Christiansen, B. Killie and J. Wright. 1995. Dietary crude oil exposure
during sexual maturation induces hepatic mixed function oxygenase (CYPIA) activity at
very low environmental temperatures in Polar cod Boreogadus saida. Marine Ecology
Progress Series 122: 307-312.
Geraci, J. R. 1990. Physiologic and toxic effects on cetaceans. In J. R. Geraci and D. J. St. Aubin
(eds), Physiologic and toxic effects on cetaceans. Academic Press, San Diego, pp. 167-
197.
Geraci, J. R. and D. J. S. Aubin. 1982. A Study of the Effects of Oil on Cetaceans. Final report
prepared for U.S. Department of the Interior, Bureau of Land Management, Washington,
D.C. 15 pp.
Geraci, J. R. and T. G. Smith. 1976. Direct and indirect effects of oil on ringed seals (Phoca
hispida) of the Beaufort Sea. Journal of the Fisheries Research Board of Canada 33:
1976-1984.
Geraci, J. R. and D. J. St. Aubin. 1985. Expanded Studies of the Effects of Oil on Cretaceans.
final report Part 1. U.S. Department of the Interior, Minerals Management Service,
Washington, D.C.
Gerdes, B., R. Brinkmeyer, G. Dieckmann and E. Helmke. 2005. Influence of crude oil on
changes of bacterial communities in Arctic sea-ice. FEMS Microbial Ecology 53: 129-
139.
GESAMP. 1993. Impact of Oil, Individual Hydrocarbons and Related Chemicals on the Marine
Environment, Including Used Lubricant Oils, Oil Spill Control Agents and Chemicals
Used Offshore. Joint Group of Experts on the Scientific Aspects of Marine
Environmental Protection, Report Study 50. 178pp.
Gilfillan, E. S., D. S. Page, E. J. Harner and P. D. Boehm. 1995a. Shoreline Ecology Program for
Prince William Sound, Alaska, Following the Exxon Valdez Oil Spill: Part 3--Biology. In
Peter G. Wells, James N. Butler and Jane S. Hughes (eds.), Exxon Valdez Oil Spill: Fate
and Effects in Alaskan Waters. American Society for Testing and Materials, Philadelphia.
Gilfillan, E. S., D. S. Page, J. M. Neff, K. R. Parker and P. D. Boehm. 2000. 1999 shoreline
conditions in the Exxon Valdez Oil Spill zone in Prince William Sound. In Proceedings of
the 23rd Arctic and Marine Oil Spill Program (AMOP) Technical Seminar. Vancouver,
Canada. Environment Canada, Ottawa, Ontario. pp. 281294.
Gilfillan, E. S., T. H. Suchanek, P. D. Boehm, E. J. Harner, D. S. Page and N. A. Sloan. 1995b.
Shoreline impacts in the Gulf of Alaska following the Exxon Valdez oil spill. In Peter G.
Wells, James N. Butler and Jane S. Hughes (eds.), Exxon Valdez Oil Spill: Fate and
Effects in Alaskan Waters. American Society for Testing and Materials, pp. 444-481.
Gjsteen, J. 2004. A model for oil spreading in cold waters. Cold Regions Science and
Technology 38: 117-125.
Gjsteen, J. K. . and S. Lset. 2004. Laboratory experiments on oil spreading in broken ice.
Cold Regions Science and Technology 38: 103-116.
Gjsteen, J. K. O., S. Lset and O. T. Gudmestad. 2003. The ability to model oil spills in broken
ice In Proceedings of the 17th International Conference on Port and Ocean Engineering
under Arctic Conditions, Trondheim, Norway.
Glaeser, J. L. and G. Vance. 1971. A Study of the Behaviour of Oil Spills in the Arctic. United
State Coast Guard, NTIS Report AD 717 142, Washington, DC, USA. 53 pp.
Glover, N. W. and D. F. Dickins. 1999. Response plans for Arctic oil and ice encounters. In
Proceedings of the 1999 International Oil Spill Conference, Seattle, WA, USA.
229

Godard, C. A. J., R. M. Smolowitz, J. Y. Wilson, R. S. Payne and J. J. Stegeman. 2004.
Induction of cetacean cytochrome P4501A1 by b-naphthoflavone exposure of skin biopsy
slices. Toxicological Sciences 80: 268-275.
Goksyr, A., T. S. Solberg and B. Serigstad. 1991. Immunochemical detection of cytochrome
P450IA1 induction in cod larvae and juveniles exposed to a water soluble fraction of
North Sea crude oil. Marine Pollution Bulletin 22: 122-127.
Goodale, D. R., M. A. M. Hyman and H. E. Winn. 1981. Cetacean responses in association with
REGAL SWORD oil spill. Chapter XI. In R. K. Edel, M. A. Hyman and M. F. Tyrell
(eds), Cetacean responses in association with REGAL SWORD oil spill. Chapter XI.
University of Rhode Island, pp. XI-I to Xl-15.
Goodman, R. H., A. G. Holoboff, T. W. Daley, P. Waddell, L. D. Murdock and M. Fingas. 1987.
A technique for the measurement of under-ice roughness to determine oil storage
volumes. In Proceedings of the 1987 International Oil Spill Conference, Baltimore, MD,
USA: 395-398.
Gosselin, M., M. Levasseur, P. A. Wheeler, R. A. Horner and B. C. Booth. 1997. New
measurements of phytoplankton and ice algal production in the Arctic Ocean. Deep Sea
Research Part II: Topical Studies in Oceanography 44: 1623-1644.
Grant, A., L. EM, L. K and M. JD. 1986. PISCES IV research submersible finds oil on Baffin
shelf. Current Research, Part A. Geological Survey Canada 86-1A: 65-69.
Gray, J. S. 2002. Perceived and real risks: Produced water from oil extraction. Marine Pollution
Bulletin 44: 1171-1172.
Greer, C. W., L. G. Whyte and T. D. Niederberger. 2010. Microbial Communities in
Hydrocarbon-Contaminated Temperate, Tropical, Alpine, and Polar Soils. In K. N.
Timmis (ed) Microbial Communities in Hydrocarbon-Contaminated Temperate, Tropical,
Alpine, and Polar Soils. Springer Berlin Heidelberg, pp. 2313-2328.
Griffiths, D. J., N. A. ritsland and T. ritsland. 1987. Marine Mammals and Petroleum
Activities in Norwegian Waters. Fisheridirektoratets Havforskningsinstitutt, Norway.
Gunette, C. C., G. A. Sergy, E. H. Owens, R. C. Prince and K. Lee. 2003. Experimental design
of the Svalbard shoreline field trials. Spill Science & Technology Bulletin 8: 245-256.
Gunette, C. C., P. Sveum, C. M. Bech and I. A. Buist. 1995. Studies of in-situ burning of
emulsions in Norway. In Proceedings of the 1995 International Oil Spill Conference,
Long Beach, CA, USA: 8110-8125.
Gulliksen, B. and O. J. Lnne. 1991. Sea ice macrofauna in the Antarctic and the Arctic. Journal
of Marine Science 2: 53-61.
Guo, Q. N., H. L. Li, M. C. Boufadel and Y. Sharifi. 2010. Hydrodynamics in a gravel beach and
its impact on the Exxon Valdez oil. Journal of Geophysical Research - Oceans 115:
C12077.
Guterman, L. 2009. Exxon Valdez turns 20. Science 323: 1558-1559.
Guyomarch, J., S. Le Floch and F. X. Merlin. 2002. Effect of suspended mineral load, water
salinity and oil type on the size of oil-mineral aggregates in the presence of chemical
dispersant. Spill Science & Technology Bulletin 8: 95-100.
Haines, J. R. and R. M. Atlas. 1982. In situ microbial degradation of Prudhoe Bay crude oil in
Beaufort Sea sediments. Marine Environmental Research 7: 91-102.
Hamel, J.-F. and A. Mercier. 2005. Arctic coastal ecology. In M. L. Schwartz (ed.),
Encyclopedia of Coastal Science. Springer, Netherlands, pp. 45-49.
230

Hamoutene, D., H. Volkoff, C. Parrish, S. Samuelson, G. Mabrouk, A. Mansour, A. Mathieu, T.
King and K. Lee. 2010. Effect of produced water on innate immunity, feeding and
antioxidant metabolism in Atlantic cod (Gadus morhua). In K. Lee
J. Neff (eds.), Environmental Risks and Mitigation Technologies for Produced Water. Springer
Publishing Company (in Press), New York, NY.
Hannam, M. L., S. D. Bamber, A. J. Moody, T. S. Galloway and M. B. Jones. 2010.
Immunotoxicity and oxidative stress in the Arctic scallop Chlamys islandica: Effects of
acute oil exposure. Ecotoxicology and Environmental Safety 73: 1440-1448.
Hannam, M. L., S. D. Bamber, J. A. Moody, T. S. Galloway and M. B. Jones. 2009. Immune
function in the Arctic scallop, Chlamys islandica, following dispersed oil exposure.
Aquatic Toxicology 92: 187-194.
Hannisdal, A., P. V. Hemmingsen, A. Silset and J. Sjoblom. 2007. Stability of water/crude oil
systems correlated to the physicochemical properties of the oil phase. Journal of
Dispersion Science and Technology 28: 639-652.
Hansen, B. H., D. Altin, S. F. Rrvik, I. B. verjordet, A. J. Olsen and T. Nordtug. 2011.
Comparative study on acute effects of water accommodated fractions of an artificially
weathered crude oil on Calanus finmarchicus and Calanus glacialis (Crustacea:
Copepoda). Science of the Total Environment 409: 704-709.
Hara, K., H. Yamaguchi and G. Sagawa. 2008. A numerical simulation of spilled oil behaviour
of ice-covered sean in the Sea of Okhotsk. In Proceedings of the 23rd International
Symposium on the Okhotsk Sea and Sea Ice, Mombetsu, Hokkaido, pp. 108-113.
Harayama, S., Y. Kasai and A. Hara. 2004. Microbial communities in oil-contaminated seawater.
Current Opinion in Biotechnology 15: 205-214.
Harris, K. A., L. M. Nichol and P. S. Ross. 2011. Hydrocarbons in sea otters in British
Columbia. Environmental Toxicology and Chemistry (accepted).
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.
Hatlen, K., L. Camus, J. Berge, G. H. Olsen and T. Baussant. 2009. Biological effects of water
soluble fraction of crude oil on the Arctic sea ice amphipod Gammarus wilkitzkii
Chemical Ecology 25: 151-162.
Hawboldt, K., B. Chen, W. Thanyamanta, S. Egli and A. Gryshchenko. 2010. Review of
Produced Water Management and Challenges in Harsh/Arctic Environments. Submitted
to the American Bureau of Shipping (ABS), August 2010.
Hawkins, J. R. 2010. Research Needs in the Beaufort Sea: Unique Challenges of Exploring in
Deepwater Regions. Presentation at the Canada-United States Northern Oil and Gas
Research Forum, Calgary AB, 7 pp. [Online]. Available:
http://www.arcus.org/files/meetings/279/271/presentations/wed08301000hawkins.pdf.
[2011, June 27].
Hazen, T. C., E. A. Dubinsky, T. Z. DeSantis, G. L. Andersen, Y. M. Piceno, N. Singh, J. K.
Jansson, A. Probst, S. E. Borglin, J.L. Fortney, W. T. Stringfellow, M. Bill, M. E.
Conrad, L. M. Tom, K. L. Chavarria, R. Alusi, R. Lamendella, D. C. Joyner, C. Spier, J.
Baelum, M. Auer, M. L. Zemla, R. Chakraborty, E. L. Sonnenthal, P. Dhaeseleer, H.
Ying, N. Holman, S. Osman, Z. Lu, J. D. V. Nostrand, Y. Deng, J. Zhou and O. U.
Mason. 2010. Deep-sea oil plume enriches indigenous oil-degrading bacteria. Science
330: 204-208.
231

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.
Heintz, R. A., J. W. Short 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.
Heiss-Blanquet, S., Y. Benoit, C. Marchaux and F. Monot. 2005. Assessing the role of alkane
hydroxylase genotypes in environmental samples by competitive PCR. Journal of
Applied Microbiology 99: 1392-1403.
Hill, P. S., A. Khelifa and K. Lee. 2002. Time scale for oil droplet stabilization by mineral
particles in turbulent suspensions. Spill Science & Technology Bulletin 8: 73-81.
Hjermann, D. ., A. Melsom, G. E. Dingsr, J. M. Durant, A. M. Eikeset, L. P. Red, G.
Ottersen, G. Storvik and N. C. Stenseth. 2007. Fish and oil in the LofotenBarents Sea
system: synoptic review of the effect of oil spills on fish populations. Marine Ecology
Progress Series 339: 283-299.
Hjorth, M. and T. G. Nielsen. 2011. Oil exposure in a warmer Arctic: potential impacts on key
zooplankton species. Marine Biology 158: 1339-1347.
Hodson, P. V., T. Cross, A. Ewert, S. Zambon and K. Lee. 2002. Evidence for the bioavailability
of PAH from oiled beach sediments in situ In Proceedings of the 25th Arctic and Marine
Oil Spill Program (AMOP) Technical Seminar. Environment Canada, Ottawa, pp. 379-
388.
Hodson, P. V., C. W. Khan, G. Saravanabhavan, L. Clarke, R. S. Brown, B. Hollebone, J. S. Z.
Wang, K. Lee and T. King. 2007. Alkyl PAH in crude oil cause chronic toxicity to early
life stages of fish In Proceedings of the 30th Arctic and Marine Oilspill Program
(AMOP) Technical Seminar. Environment Canada, Ottawa, Ontario: 291-300.
Hodson, P. V., P. J. Kloepper-Sams, K. R. Munkittrick, W. L. Lockhart, D. A. Metner, P. L.
Luxon, I. R. Smith, M. M. Gagnon, M. Servos and J. F. Payne. 1991. Protocols for
measuring mixed function oxygenases of fish liver. Canadian Technical Report on Fish
and Aquatic Sciences 1829: 1-51.
Holdway, D. A. and D. Heggie. 2002. The acute and chronic effects of wastes associated with
offshore oil and gas production on temperate and tropical marine ecological processes.
Marine Pollution Bulletin 183: 73-78.
Hood, D. W., W. E. Sheils and E. J. Kelley. 1973. Environmental Studies of Port Valdez. Inst.
Mar. Sci. Occas. Publ. 3. 495 pp.
Hop, H. and W. M. Tonn. 1998. Gastric evacuation rates and daily rations of Arctic cod
(Boreogadus saida) at low temperatures. Polar Biology 19: 293-301.
Hop, H., W. M. Tonn and H. E. Welch. 1997. Bioenergetics of Arctic cod (Boreogadus saida) at
low temperatures. Canadian Journal of Fisheries and Aquatic Sciences 54: 1772-1784.
Hopcroft, R. R., C. Clarke, R. J. Nelson and K. A. Raskoff. 2005. Zooplankton communities of
the Arctics Canada Basin: The contribution by smaller taxa. Polar Biology 28: 198-206.
Houghton, J. P., R. H. Gilmour, D. C. Lees, W. B. Driskell and S. C. Lindstrom. 1995. Prince
William Sound intertidal biota - Good news and bad news five years later In in
Proceedings of the 18th Arctic and Marine Oilspill Program (AMOP) Technical Seminar.
Environment Canada, Ottawa, Ontario: 1075-1094.
Hsiao, S. I. C. 1978. Effects of crude oils on the growth of Arctic marine phytoplankton.
Environmental Pollution (1970) 17: 93-107.
232

Hsiao, S. I. C., D. W. Kittle and M. G. Foy. 1978. Effects of crude oils and the oil dispersant
corexit on primary production of arctic marine phytoplankton and seaweed.
Environmental Pollution (1970) 15: 209-221.
Huang, C. P. and H. A. Elliott. 1977. The stability of emulsified crude oils as affected by
suspended particles. In Wolfe, D.A. (ed.), Fate and Effects of Petroleum Hycrocarbons in
Marine Organisms and Ecosystems. Pergamon, New York, NY, pp. 413-420.
Huang, J. C. 1983. A review of the state-of-the-art of oil spill fate/behavior models. In
Proceedings of the 1983 International Oil Spill Conference. American Petroleum
Institute, Pp. 313-322, San Antonio, TX.
Hubbard, D. 2009. The failure of risk management: why it's broken and how to fix it. John Wiley
& Sons. 281 pp.
Huggett, R. J., J. M. Neff, J. J. Stegeman, B. Woodin, K. R. Parker and J. S. Brown. 2006.
Biomarkers of PAH exposure in an intertidal fish species from Prince William Sound,
Alaska: 2004-2005. Environmental Science & Technology 40: 6513-6517.
Huggett, R. J., J. J. Stegeman, D. S. Page, K. R. Parker, B. Woodin and J. S. Brown. 2003.
Biomarkers in fish from Prince William Sound and the Gulf of Alaska: 1999- 2000.
Environmental Science & Technology 37: 4043-4051.
Humphey, B., P. D. Boehm, M. C. Hamilton and R. J. Norstrom. 1987. The fate of chemically
dispersed and untreated crude oil in Arctic benthic biota. Arctic 40: 149-161.
Huntington, H. P. 2009. A preliminary assessment of threats to arctic marine mammals and their
conservation in the coming decades. Marine Policy 33: 77-82.
Hurley, G. V. and J. Ellis. 2004. Environmental effects of exploratory drilling offshore Canada:
EEM data and literature review. Final Report. Canadian Environmental Assessment
Agencys Regulatory Advisory Committee (RAC), 61 pp. + App.
Husky Energy. 2005. White Rose Environmental Effects Monitoring Program 2005. Vol. 1 and 2.
2005 Environmental Effect Monitoring Program Report, Husky Energy. St. John's, NL.
Hutcheson, M. S. and G. W. Harris. 1982. Sublethal Effects of a Water-soluble Fraction and
Chemically Dispersed Form of Crude Oil on Energy Partitioning in Two Arctic Bivalves.
Environment Canada Environmental Protection Service. 34 pp.
Hylland, K., T. Lang, and D. Vethaak (Eds.). 2006. Biological Effects of Contaminants in
Marine Pelagic Ecosystems. SETAC Press, Pensacola, FL. 474 pp.
Ikvalko, J. and B. Gerdes. 2004. Biodiversity of Arctic sea ice biota and possible effects of oil
spills during oil transportation. In The ACIA International Scientific Symposium on
Climate Change in the Arctic. Extended Abstracts. Reykjavik, Iceland, 9-12 November
2004. pp. 16-18.
Iken, K., B. A. Bluhm and R. Gradinger. 2005. Food web structure in the high Arctic Canada
Basin: Evidence from 13C and 15N analysis. Polar Biology 28: 238-249.
IMO (International Maritime Organisation). 1988. Manual for Oil Pollution, Section IV
Combating Oil Spills. pp. 1939.
INAC. 1995. Petroleum Exploration in Northern Canada: a Guide to Oil and Gas Exploration
and Potential. Department of Indian and Northern Affairs Canada (now, Aboriginal
Affairs and Northern Development Canada), Ottawa ON. 110 pp.
INAC. 2010. Beaufort PEMT Valued Component Description References. Indian and Northern
Affairs Canada. [Online]. Available: http://www.ainc-inac.gc.ca/nth/og/pemt/ref-eng.asp.
[2011, June 1].
233

Incardona, J. P., M. G. Carls, H. L. Day, C. A. Sloan, J. L. Bolton, T. K. Collier 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.
Ingebrigtsen, K., J. S. Christiansen, . Lindhe and I. Brandt. 2000. Disposition and cellular
binding of
3
H-benzo(a)pyrene at subzero temperatures: studies in an aglomerular arctic
teleost fish the polar cod(Boreogadus saida). Polar Biology 23: 503-509.
Irvine, G. V., D. H. Mann and J. W. Short. 2006. Persistence of 10-year old Exxon Valdez oil on
Gulf of Alaska beaches: The importance of boulder-armoring. Marine Pollution Bulletin
52: 1011-1022.
ITG, I. T. G. 1983. Oil Spill Response in the Arctic. Part 2: Field Demonstrations in Broken Ice.
Shell Oil Company and AMOCO Production Company, Archorage, AK, USA.
ITOPFL. 1986. Response to Marine Oil Spills. International Tanker Owners Pollution Federation
Limited, Whitherby & Co. Ltd.
ITOPFL. 2002. The Fate of Marine Oil Spills. International Tanker Owners Pollution Federation
Limited. [Online]. Available: http://www.itopf.com/_assets/documents/tip2.pdf. [2011,
April 27].
IUCN Otter Specialist Group. 2010. Enhydra lutris (Linnaeus, 1758), the Sea Otter, Enhydra
lutris (Linnaeus, 1758), the Sea Otter.
Ivanov, B., A. Bezgreshnov, N. Kubyshkin 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.
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.
Jacobs, R. P. W. M., R. O. H. Grant, J. Kwant, J. M. Marquenie and E. Mentzer. 1992. The
Composition of Produced Water from Shell Operated Oil and Gas Production in the
North Sea. In J.P. Ray and F.R. Engelhardt (eds.), Produced Water:
Technological/Environmental Issues and Solutions. Plenum Press, New York, NY, pp.
13-21.
Jaraula, C. M. B., F. Kenig, P. T. Doran, J. C. Priscu and K. A. Welch. 2008. SPME-GCMS
study of the natural attenuation of aviation diesel spilled on the perennial ice cover of
Lake Fryxell, Antarctica. Science of the Total Environment 407: 250-262.
Jensen, H. V. 1994. The 1993 Norwegian Experimental Spill. In Proceedings of the Conference
on Oil Spill Response in Broken Ice. DF Dickins Assoc., La Jolla, CA, USA.
Jensen, L. K. and J. Carroll. 2010. Experimental studies of reproduction and feeding for two
Arctic-dwelling Calanus species exposed to crude oil. Aquatic Biology 10: 261-271.
Jensen, M. H., T. G. Nielsen and I. Dahllf. 2008. Effects of pyrene on grazing and reproduction
of Calanus finmarchicus and Calanus glacialis from Disko Bay, West Greenland.
Aquatic Toxicology 87: 99-107.
Johnsen, S., T. I. Re, G. Durell and M. Reed. 1998. Dilution and bioavailability of produced
water components in the northern North Sea. A combined modeling and field study. SPE
46578. In Proceedings of the 1998 SPE International Conference on Health, Safety and
Environment in Oil and Gas Exploration and Production. Society of Petroleum
Engineers, Richardson, TX: 1-11.
234

Johnsen, S., T. I. R. Utvik, E. Garland, B. d. Vals and J. Campbell. 2004. Environmental fate and
effects of contaminants in produced water. SPE 86708. Paper presented at the Seventh
SPE International Conference on Health, Safety, and Environment in Oil and Gas
Exploration and Production. Society of Petroleum Engineers, Richardson, TX. 9 pp.
Jokuty, P. 2001. Properties of crude oil and oil products (not just another pretty database). In
Proceedings of the 2001 International Oil Spill Conference, Tampa, FL, USA: 975-981.
Jones, J. A. 2005. An Introduction to Factor Analysis of Information Risk (FAIR): A Framework
for Understanding, Analyzing, and Measuring Information Risk. Risk Management
Insight. 76 pp. [Online]. Available
http://riskmanagementinsight.com/media/documents/FAIR_Introduction.pdf. [2011, June
18].
Jonsson, H., R. C. Sundt, E. Aas and S. Sanni. 2010. The Arctic is no longer put on ice:
Evaluation of Polar cod (Boreogadus saida) as a monitoring species of oil pollution in
cold waters. Marine Pollution Bulletin 60: 390-395.
Jnsson, M., A. Abrahamson, B. Brunstrm, I. Brandt, K. Ingebrigtsen and E. H. Jrgensen.
2003. EROD activity in gill filaments of anadromous and marine fish as a biomarker of
dioxin-like pollutants. Comparative Biochemistry and Physiology Part C: Toxicology &
Pharmacology 136: 235-243.
Jrgensen, E. H. and J. Wolkers. 1999. Effect of temperature on the P4501A response in winter-
and summer-acclimated Arctic char (Salvelinus alpinus) after oral benzo[a]pyrene
exposure. Canadian Journal of Fisherines and Aquatic Sciences 56: 1370-1375.
Junge, K., H. Eicken and J. Deming. 2004. Bacterial activity at -2 to -20C in Arctic wintertime
sea ice. Applied and Environmental Microbiology 70: 550-557.
Junge, K., J. Imhoff, J. Staley and J. Deming. 2002. Phylogenetic diversity of numerically
important bacteria in Arctic sea ice. Microbial Ecology 43: 315-328.
Junge, K., C. Krembs, J. Deming, A. Stierle and H. Eicken. 2001. A microscopic approach to
investigate bacteria under in-situ conditions in sea-ice samples. Annals of Glaciology 33:
304-310.
Juurmaa, K. 2006. Arctic Operational Platform, ARCOP Final Report. Aker Finnyards Inc., 330
pp.
Kelly, B. P., J. L. Bengtson, P. L. Boveng, M. F. Cameron, S. P. Dahle, J. K. Jansen, E. A.
Logerwell, J. E. Overland, C. L. Sabine, G. T. Waring and J. M. Wilder. 2010. Status
Review of the Ringed Seal (Phoca hispida). U.S. Department of Commerce. 250 pp.
Khan, R. A. and J. W. Kiceniuk. 1988. Effect of petroleum aromatic hydrocarbons on
monogeneids parasitizing Atlantic cod Gadus morhua. Bulletin of Environmental
Contamination and Toxicology 41: 94-100.
Khelifa, A. 2010. A summary review of modelling oil in ice. In Proceedings of the 33rd Arctic
and Marine Oilspill Program (AMOP) Technical Seminar on Environmental
Contamination and Response. Environment Canada, Ottawa, Ontario. pp. 587-608.
Khelifa, A., L. O. Ajijolaiya, P. MacPherson, K. Lee, P. S. Hill, S. Gharbi and M. Blouin. 2005a.
Validation of OMA formation in cold brackish and sea waters. In Proceedings of the 28th
Arctic and Marine Oilspill Program (AMOP) Technical Seminar on Environmental
Contamination and Response. Environment Canada, Ottawa, Ontario. pp.527-538.
Khelifa, A. and C. Brown. 2010. Fate of Oil-SPM Aggregates in the St-Lawrence Estuary. In
Proceedings of the 33rd Arctic and Marine Oilspill Program (AMOP) Technical Seminar
235

on Environmental Contamination and Response. Environment Canada, Ottawa, Ontario.
pp.609-622.
Khelifa, A., P. S. Hill and K. Lee. 2003a. A stochastic model to predict the formation of oil-
mineral aggregates. In: 26th Proceedings of Arctic and Marine OilSpill Program
Technical Seminar, Victoria, British Columbia, Canada, pp. 893-910.
Khelifa, A., M. Pahlow, A. Vezina, K. Lee and C. Hannah. 2003b. Numerical investigation of
impact of nutrient inputs from produced water on the marine planktonic community. In
Proceedings of the 26th Arctic and Marine Oilspill Program (AMOP) Technical Seminar.
Victoria, Canada. Environment Canada. Ottawa, Ontario. pp. 323-334.
Khelifa, A., P. Stoffyn-Egli, P. S. Hill and K. Lee. 2005b. Effects of salinity and clay type on oil-
mineral aggregation. Marine Environmental Research 59: 235-254.
Kjeilen-Eilertsen, G., J. M. Jersak and S. Westerlund. 2011. Developing treatment products for
increased microbial degradation of petroleum oil spills across open-water surfaces. In
Proceedings of the 2011 Arctic Technology Conference, Houston TX Feb 7-9, 9 pp.
Kooyman, G. L., R. W. Davis and M. A. Castellini. 1977. Thermal conductance of immersed
pinniped and sea otter pelts before and after oiling with Prudhoe Bay crude. In
D.A.Wolfe (ed.), Fate and Effects of Petroleum Hydrocarbons in Marine Organisms and
Ecosystems. Pergamon Press, Oxford, UK, pp. 151-157.
Kovacs, A., R. M. Morey, D. F. Cundy and G. Decoff. 1981. Pooling of oil under sea ice. In
Proceedings of the 6th International Conference on Port and Ocean Engineering under
Arctic Conditions, Quebec, Canada. pp. 912-922.
Krembs, C. and J. Deming. 2011. What do we know about organisms that thrive in Arctic sea
ice? Sea ice: a refuge for life in polar seas? National Oceanic and Atmospheric
Administration. Arctic Theme Page [Online]. Available:
http://www.arctic.noaa.gov/essay_krembsdeming.html [2011, May 25].
Krembs, C., H. Eicken and J. W. Deming. 2011. Exopolymer alteration of physical properties of
sea ice and implications for ice habitability and biogeochemistry in a warmer Arctic.
PNAS Proceedings of the National Academy of Sciences [Online]. Published ahead of
print, available: http://www.pnas.org/content/early/2011/02/11/1100701108.abstract.
[2011, February 22].
Krembs, C., R. Gradinger and M. Spindler. 2000. Implications of brine channel geometry and
surface area for the interaction of sympagic organisms in Arctic sea ice. Experimental
Marine Biology and Ecology 1: 55-80.
Labbe, D., R. Margesin, F. Schinner, L. G. Whyte and C. W. Greer. 2007. Comparative
phylogenetic analysis of microbial communities in pristine and hydrocarbon-
contaminated Alpine soils. FEMS Microbiology Ecology 59: 466-475.
Lagerloef, G. S. E., C. Swift and D. Le Vine. 1995. Sea surface salinity: The next remote sensing
challenge. Oceanography 8: 44-50.
Laidre, K. L., I. Stirling, L. F. Lowry, . y. Wiig, M. P. Heide-Jrgensen and S. H. Ferguson.
2008. Quantifying the sensitivity of Arctic marine mammals to climate-induced habitat
change. Ecological Applications 18: 97-125.
Lally, J. 2011. U.S. Coast Guard marine environmental response in the Arctic: Challenges and
initiatives. In Proceedings of the 2011 International Oil Spill Conference, Portland, OR,
USA. 9 pp.
Lampela, K. and K. Jolma. 2011. Mechanical oil spill recovery in Ice; Finnish approach. In
Proceedings of the 2011 International Oil Spill Conference, Portland, OR, USA. 9 pp.
236

Landes, K. 1973. Mother Nature as an oil polluter. The American Association of Petroleum
Geologists (AAPG) Bulletin 57: 637-641.
Landis, W. G. 2007. The Exxon Valdez oil spill revisited and the dangers of normative science.
Integrated Environmental Assessment and Management 3: 439-441.
Langworthy, D. E., R. D. Stapleton, G. S. Sayler and R. H. Findlay. 1998. Genotypic and
phenotypic responses of a riverine microbial community to polycyclic aromatic
hydrocarbon contamination. Applied and Environmental Microbiology 64: 3422-3428.
Leahy, J. G. and R. R. Colwell. 1990. Microbial degradation of hydrocarbons in the
environment. Microbiology and Molecular Biology Reviews 54: 305-315.
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: 3-8.
Lee, K., S. L. Armsworthy, S. E. Cobanli, N. A. Cochrane, P. J. Cranford, A. Drozdowski, D.
Hamoutene, C. G. Hannah, E. Kennedy, T. King, H. Niu, B. A. Law, Z. Li, T. G.
Milligan, J. Neff, J. F. Payne, B. J. Robinson, M. Romero and T. Worcester. 2011a.
Consideration of the Potential Impacts on the Marine Environment Associated with
Offshore Petroleum Exploration and Development Activities. DFO. Can. Sci. Advis. Sec.
Res. Doc. 2011/060: xii + 134p.
Lee, K., K. Azetsu-Scott, S. E. Cobanli, J. Dalziel, S. Niven, G. Wohlgeschaffen and P. Yeats.
2005a. Overview of potential impacts of produced water discharges in Atlantic Canada.
In S.L. Armworthy, P.J. Cranford and K. Lee (eds) Offshore Oil and Gas Environmental
Effects Monitoring: Approaches and Technologies. Battelle Press, Columbus, OH. pp.
319-342.
Lee, K., T. King, B. Robinson, Z. Li, L. Burridge, M. Lyons, D. Wong, K. Mackeigan, S.
Courtenay, S. Johnson, M. Boudreau, P. Hodson, C. Greer and A. Venosa. 2011b.
Toxicity effects of chemically-dispersed crude oil on fish. In Proceedings of the 2011
International Oil Spill Conference, Portland, OR, USA: 17.
Lee, K., Z. Li, M. C. Boufadel, A. D. Venosa and M. Scott Miles. 2009a. Wave tank studies on
dispersant effectiveness as a function of energy dissipation rate and particle size
distribution. Final report submitted to NOAA/CRRC/UNH. Pp67+ appendix, Wave tank
studies on dispersant effectiveness as a function of energy dissipation rate and particle
size distribution. Final report submitted to NOAA/CRRC/UNH. Pp67+ appendix.
Lee, K., Z. Li, T. King, P. Kepkay, M. C. Boufadel and A. D. Venosa. 2008. Wave tank studies
on formation and transport of OMA from the chemically dispersed oil. In Davidson,
W.F., Lee, K. and Cogswell, A (eds.), Oil Spill Response: A Global Perspective.
Springer, Netherlands, pp. 159-177.
Lee, K., Z. Li, H. Niu, P. Kepkay, Y. Zheng, M. Boufadel and Z. Chen. 2009b. Enhancement of
oil-mineral-aggregate formation to mitigate oil spills in offshore oil and gas activities.
Final Report (Contract No. M07PC13035) submitted to Minerals Management Service.
91p. March 30, 2009, Enhancement of oil-mineral-aggregate formation to mitigate oil
spills in offshore oil and gas activities. Final Report (Contract No. M07PC13035)
submitted to Minerals Management Service. 91p. March 30, 2009, pp. 1-99.
Lee, K., Z. Li, H. Niu, P. Kepkay, Y. Zheng, M. Boufadel and Z. Chen. 2009c. Enhancement of
oil-mineral-aggregate formation to mitigate oil spills in offshore oil and gas activities.
Final Report (Contract No. M07PC13035) submitted to Minerals Management Service.
91p. March 30, 2009.
237

Lee, K., Z. Li, B. Robinson, P. E. Kepkay, M. Blouin and B. Doyon. 2011c. Field trials of in-situ
oil spill countermeasures in ice-infested waters. In Proceedings of the 2011 International
Oil Spill Conference, Portland, OR, USA.
Lee, K., Z. Li, B. Robinson, P. E. Kepkay, X. Ma, S. Cobanli, T. King, M. Blouin and B. Doyon.
2009d. In-situ remediation of oil spills in ice-packed waters: Enhanced dispersion and
biodegradation of petroleum hydrocarbons. In Proceedings of the 10th International In
Situ and On-Site Bioremediation Symposium.
Lee, K., T. Lunel, P. Wood, R. Swannel and P. Stoffyn-Egli. 1997. Shoreline cleanup by
acceleration of clay-oil flocculation process. In The 1997 International Oil Spill
Conference. American Petroleum Institute, Washington, DC, Publication no. 4651: 235-
240.
Lee, K., T. Nedwed and R. C. Prince. 2011d. Lab tests on the biodegradation rates of chemically
dispersed oil must consider natural dilution. In Proceedings of the 2011 International Oil
Spill Conference, Portland, OR, USA. 12 pp.
Lee, K. and J. Neff. 2009. Environmental Risks and Advances in Mitigation Technologies In in
Proceedings of the 2009 International Produced Water Conference, October 17 18,
2007. ESRF Technical Report Series (In Press).
Lee, K. and P. Stoffyn-Egli. 2001. Characterization of oil-mineral aggregates. In Proceedings of
the 2001 International Oil Spill Conference, Tampa, FL, USA: 991-996.
Lee, K., P. Stoffyn-Egli and E. H. Owens. 2001. Natural dispersion of oil in a freshwater
ecosystem: Desaguadero Pipeline Spill, Bolivia. In Proceedings of the 2001 International
Oil Spill Conference, Tampa, FL, USA: 1445-1448.
Lee, K., P. Stoffyn-Egli and E. H. Owens. 2002. The OSSA II pipeline oil spill: Natural
mitigation of a riverine oil spill by oil-mineral aggregate formation. Spill Science &
Technology Bulletin 7: 149-154.
Lee, K., P. Stoffyn-Egli, G. H. Tremblay, E. H. Owens, G. A. Sergy, C. C. Guenette and R. C.
Prince. 2003. Oil-mineral aggregate formation on oiled beaches: Natural attenuation and
sediment relocation. Spill Science & Technology Bulletin 8: 285-296.
Lee, K., P. Stoffyn-Egli, P. Wood and T. Lunel. 1998. Formation and structure of oil-mineral
fines aggregates in coastal environments. In Proceedings of the 21st Arctic and Marine
Oilspill Program (AMOP) Technical Seminar. Environment Canada, Ottawa, Ontario, pp.
911-921.
Lee, K. and K. L. Tay. 1998. Measurement of microbial exoenzyme activity in sediments for
environmental impact assessment. In P.G. Wells, K. Lee and C. Blaise (eds.), Microscale
Aquatic Toxicology: Advances Techniques and Practice. CRC Press. pp 219-236.
Lee, K., A. D. Venosa and F. X. Merlin. 2005b. Marine oil spill bioremediation: field studies for
development of operational spill response strategies. In Proceedings of the International
Council for the Exploration of the Sea (ICES) CM. Theme Session on Oil Spills in Marine
Ecosystems: Impacts and Remediation. The Council, Copenhagen, Denmark.
Lee, K., A. M. Weise and S. St-Pierre. 1996. Enhanced oil biodegradation with mineral fine
interaction. Spill Science & Technology Bulletin 3: 263-267.
Lee, R. F. 2003. Photo-oxidation and photo-toxicity of crude and refined oils. Spill Science and
Technology Bulletin 8: 157-162.
Lees, D. C., J. P. Houghton and W. B. Driskell. 1996. Short-term effects of several types of
shoreline treatment on rocky intertidal biota in Prince William Sound. In Proceedings of
the 1996 Exxon Valdez Oil Spill Symposium, Bethesda, MD, USA: 329348.
238

Lewis, A. 2000. Behavior of Oil Spilled in Ice A Literature Survey Focused on Modeling,
Behavior of Oil Spilled in Ice A Literature Survey Focused on Modeling. Report
prepared by Applied Science Associates (ASA) for ExxonMobil, Project Number ASA
00-149.
Lewis, A., P. S. Daling, T. Stram-Kristiansen, A. B. Nordvik and R. J. Fiocco. 1995. Weathering
and chemical dispersion of oil at sea. In Proceedings of the 1995 International Oil Spill
Conference Long Beach, CA, USA: 157-164.
Li, H. and M. C. Boufadel. 2010. Long-term persistence of oil from the Exxon Valdez spill in
two-layer beaches. Nature Geoscience 3: 96-99.
Li, Z., P. Kepkay, K. Lee, T. King, M. C. Boufadel and A. D. Venosa. 2007. Effects of chemical
dispersants and mineral fines on oil dispersion in a wave tank under breaking waves.
Marine Pollution Bulletin 54: 983-993.
Li, Z., K. Lee, T. King, M. C. Boufadel and A. D. Venosa. 2009a. Evaluating chemical
dispersant efficacy in an experimental wave tank: 2, significant factors determining in-
situ oil droplet size distribution. Environmental Engineering Science 26: 1407-1418.
Li, Z., K. Lee, T. King, M. C. Boufadel and A. D. Venosa. 2009b. Evaluating crude oil chemical
dispersion efficacy in a flow-through wave tank under regular non-breaking wave and
breaking wave conditions. Marine Pollution Bulletin 58: 735-744.
Li, Z., K. Lee, T. King, M. C. Boufadel and A. D. Venosa. 2010. Effects of temperature and
wave conditions on dispersion efficacy of heavy fuel oil in an experimental wave tank.
Marine Pollution Bulletin 60: 1550-1559.
Li, Z., K. Lee, T. King, P. Kepkay, M. C. Boufadel and A. D. Venosa. 2009c. Evaluating
chemical dispersant efficacy in an experimental wave tank: 1, dispersant effectiveness as
a function of energy dissipation rate. Environmental Engineering Science 26: 1139-1148.
Logerwell, E. A. and M. C. Baker. 2011. A Conceptual Model for Determining Oil Fate and
Effects on Habitat and Wildlife in the Arctic. National Oceanic and Atmospheric
Administration (in Press).
Louisiana Dept. of Environmental Quality. 1990. Spread Sheet Containing Volumes, Salinity,
Radium Isotopes, and Toxicity of Produced Water from Coastal Louisiana. Obtained
from M.T. Stephenson, Texaco, Inc, Bellaire, TX.
Lovley, D. R., J. D. Coates, D. A. Saffarini and D. J. Lonergan. 1997. Dissimilatory iron
reduction. In Winkelmann, G. and Carrano, C. J. (eds.), Transition Metals in Microbial
Metabolism. Harwood Academic Publishers, Geneva, Switzerland, pp. 187-215.
Lukin, A., J. Sharova, L. Belicheva and L. Camus. 2011. Assessment of fish health status in the
Pechora River: Effects of contamination. Ecotoxicology and Environmental Safety 74:
355-365.
Lunan, D. 2005. Back to the Beaufort. Oilweek, Oct 2005. [Online]. Available:
http://www.oilweek.com/articles.asp?ID=232. [2011, June 28].
Lunel, T., K. Lee, R. Swannell, P. Wood, J. Rusin, N. Bailey, C. Halliwell, L. Davies, M.
Sommerville, A. Dobie, D. Mitchell and M. McDonagh. 1996. Shoreline cleanup 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. Environment Canada, Ottawa, Ontario: 1521-1540.
Lunel, T., R. Swannell, J. Rusin, P. Wood, N. Bailey, C. Halliwell, L. Davies, M. Sommerville,
A. Dobie, D. Mitchell, M. McDonagh and K. Lee. 1997. Monitoring the effectiveness of
239

response operations during the Sea Empress incident: A key component of the successful
counter-pollution response. Spill Science & Technology Bulletin 2: 99-112.
Luz, A., V. Pellizari, L. Whyte and C. Greer. 2004. A survey of indigenous microbial
hydrocarbon degradation genes in soils from Antarctica and Brazil. Canadian Journal of
Microbiology 50: 323-333.
Ly, J. M. and O. K. Bjerkemo. 2011. Prevention, preparedness, and response to shipping
incidents in the Barents Sea. In Proceedings of the 2011 International Oil Spill
Conference, Portland, OR, USA. 9 pp.
Lyons, M. C., D. L. Wong, I. Mulder, K. Lee and L. E. Burridge. 2011. The influence of water
temperature on induced liver EROD activity in Atlantic cod (Gadus morhua) exposed to
crude oil and oil dispersants. Ecotoxicology and Environmental Safety 74: 904-910.
MacNeil, M. R. and R. H. Goodman. 1987. Oil Motion During Lead Closure. Environmental
Studies Revolving Funds Report No. 053, Ottawa, ON. 13 pp.
Maki, H., T. Sasaki and S. Harayama. 2001. Photo-oxidation of biodegraded crude oil and
toxicity of the photo-oxidized products. Chemosphere 44: 1145-1151.
Malcolm, J. D. and A. B. Cammaert. 1981. Transport and deposition of oil and gas spills under
sea ice. In Proceedings of the Fourth Arctic and Marine Oilspill Program (AMOP)
Technical Seminar. Edmonton, AB. Environment Canada, Ottawa, Ontario: 45-73.
Malins, D. C. 1977. Effects of Petroleum on Arctic and Subarctic Environments and Organisms.
V. 1: Nature and Fate of Petroleum. V. 2: Biological Effects. Academic Press, New York,
491 pp.
Margesin, R., D. Labbe, F. Schinner, C. W. Greer and L. G. Whyte. 2003. Characterization of
hydrocarbon-degrading microbial populations in contaminated and pristine alpine soils.
Applied and Environmental Microbiology 69: 3085-3092.
Martin, S., P. Kauffman and P. E. Welander. 1976. A Laboratory Study of the Dispersion of
Crude Oil Within Sea Ice Growth in a Wave Field. University of Washington Report
No.69. Seattle, WA, USA.
Matkin, C. O., G. M. Ellis, M. E. Dahlheim and J. Zeh. 1994. Chapter 8: Status of killer whales
in Prince William Sound, 1985-1992. In Thomas R. Loughlin (ed.), Marine Mammals
and the Exxon Valdez. Academic Press, San Diego, CA, 395 pp.
Matkin, C. O., E. L. Saulitis, G. M. Ellis, P. Olesiuk and S. D. Rice. 2008. Ongoing population-
level impacts on killer whales Orcinus orca following the Exxon Valdez oil spill in Prince
William Sound, Alaska. Marine Ecology Progress Series 356: 269-281.
Matthews, D. 2011. The prospects and the perils of Beaufort Sea oil: how Canada is dealing with
its high north. IAGS Journal of Energy Security. [Online]. Available:
http://www.ensec.org/index.php?option=com_content&view=article&id=311:the-
prospects-and-the-perils-of-beaufort-sea-oil-how-canada-is-dealing-with-its-high-
north&catid=116:content0411&Itemid=375. [2011, June 27].
McFarlin, K. M., M. B. Leigh and R. Perkins. 2011a. Indigenous Microorganisms Degrade Oil
in Arctic Seawater (Poster). In the 2011 International Oil Spill Conference, Portland, OR,
USA. 1pp.
McFarlin, K. M., R. A. Perkins, W. W. Gardiner and J. D. Word. 2011b. Evaluating the
biodegradability and effects of dispersed oil using arctic test species and conditions:
Phase 2 activities, Evaluating the biodegradability and effects of dispersed oil using arctic
test species and conditions: Phase 2 activities, Calgary, Canada.
240

McFarlin, K. M., R. A. Perkins, W. W. Gardiner and J. D. Word. 2011c. Evaluating the
biodegradability and effects of dispersed oil using arctic test species and conditions:
Phase 2 activities. In Proceedings of the 34th Arctic and Marine Oilspill Program
(AMOP) Technical Seminar on Environmental Contamination and Response. Calgary,
AB. Environment Canada, Ottawa, Ontario.
McFarlin, K. M., R. A. Perkins, W. W. Gardiner, J. D. Word and J. Q. Word. 2011d. Toxicity of
physically and chemically dispersed oil to selected Arctic species. In Proceedings of the
2011 International Oil Spill Conference, Portland, OR, USA. 7 pp.
McInerney, M. J., C. G. Struchtemeyer, J. Sieber, H. Mouttaki, A. J. M. Stams, B. Schink, L.
Rohlin and R. P. Gunsalus. 2008. Physiology, ecology, phylogeny, and genomics of
microorganisms capable of syntrophic metabolism. In Wiegel, J., Maier, R. J. and
Adams, M. W. W. (eds.), Incredible Anaerobes: from Physiology to Genomics to Fuels.
Wiley-Blackwell. Vol. 1125, pp. 58-72.
McMinn, T. J. 1972. Crude Oil Behaviour on Arctic Winter Ice: Final Report. Office of
Research and Development, United States Coast Guard, Project 734108, Wash. D.C.,
NTIS Publication No. AP-754, 261 pp.
McMinn, T. J. and P. C. Golden. 1973. Behavioural characteristics and cleanup techniques of
north slope crude oil in an Arctic environment. In Proceedings of the 1973 Joint
Conference on Prevention and Control of Oil Spills, American Petroleum Institute,
Washington, DC, USA: 263-276.
Means, J. C., C. S. Milan and D. J. McMillin. 1990. Hydrocarbon and trace metal concentrations
in produced water effluents and proximate sediments. In St. P, K.M. (ed.), An
Assessment of Produced Water Impacts to Low-Energy, Brackish Water Systems in
Southeast Louisiana. Report to Louisiana Dept. of Environmental Quality, Water
Pollution Control Division, Lockport, LA, pp. 94-199.
Meeres, L. S. 1987. Meteorological Operations at Cape Hatt in support of the Baffin Island Oil
Spill Project. Arctic 40: 42-50.
Meneley, R. 2008. The significance of oil in the Sverdrup Basin. Back to Exploration
Convention, May 12-15, 2008, CSPG CSEG CWLS, Abstract Archive. pp. 579-582.
[Online]. Available:
http://www.cseg.ca/conventions/abstracts/2008/2008abstracts/015.pdf. [2011, June 28].
Michel, C., T. G. Nielsen, C. Nozais and M.Gosselin. 2002. Significance of sedimentation and
grazing by ice micro- and meiofauna for carbon cycling in annual sea ice (northern Baffin
Bay). Aquatic Microbial Ecology 30: 57-68.
Michel, J., Z. Nixon, M. O. Hayes, G. V. Irvine and J. W. Short. 2011. The distribution of
lingering subsurface oil from the Exxon Valdez Oil Spill. In Proceedings of the 2011
International Oil Spill Conference, Portland, OR, USA. 14 pp.
Miller, D. S., D. B. Peakall, D. J. Hallett, J. R. Bend and G. L. Foureman. 1982. Toxicity of
Prudhoe Bay crude oil and its aromatic fractions to nestling herring gulls. Environmental
Research 27: 206-215.
Mohn, W. W. and G. R. Stewart. 2000. Limiting factors for hydrocarbon biodegradation at low
temperature in Arctic soils. Soil Biology and Biochemistry 32: 1161-1172.
Moles, A. and T. L. Wade. 2001. Parasitism and phagocytic function among sand lance
Ammodytes hexapterus pallas exposed to crude oil-laden sediments. Bulletin of
Environmental Contamination and Toxicology 66: 528-535.
241

Monson, D. H., D. F. Doak, B. E. Ballachey, A. Johnson and J. L. Bodkin. 2000. Long term
impacts of the Exxon Valdez oil spill on sea otters, assessed through age-dependent
mortality patterns. Proceedings of the National Academy of Sciences 97: 6562-6567.
Moore, S. E. and J. T. Clarke. 2002. Potential impact of offshore human activities on gray
whales (Eschrichtius robustus). Journal of Cetacean Research and Management 4: 19-
25.
Moore, S. E. and H. P. Huntington. 2008. Arctic marine mammals and climate change: Impacts
and resilience. Ecological Applications 18: S157-S165.
Morita, R. Y. 1975. Psychrophilic bacteria. Bacteriological Reviews 39: 144-167.
Mullin, J. V. 2011. Bureau of Ocean Management, Regulation, and Enforcement Arctic Oil Spill
Response Research and Development Program: Two Decades of Achievement. In
Proceedings of the 2011 International Oil Spill Conference, Portland, OR, USA. 15 pp.
Muschenheim, D. K. and K. Lee. 2002. Removal of oil from the sea surface through particulate
interactions: Review and prospectus. Spill Science & Technology Bulletin 8: 9-18.
Nahrgang, J., L. Camus, F. Broms, J. S. Christiansen and H. Hop. 2010a. Seasonal baseline
levels of physiological and biochemical parameters in polar cod (Boreogadus saida):
Implications for environmental monitoring. Marine Pollution Bulletin 60: 1336-1345.
Nahrgang, J., L. Camus, M. G. Carls, P. Gonzalez, M. Jnsson, I. C. Taban, R. K. Bechmann, J.
S. Christiansen and H. Hop. 2010b. Biomarker responses in polar cod (Boreogadus
saida) exposed to the water soluble fraction of crude oil. Aquatic Toxicology 97: 234-
242.
Nahrgang, J., L. Camus, P. Gonzalez, A. Goksyr, J. S. Christiansen and H. Hop. 2009. PAH
biomarker responses in polar cod (Boreogadus saida) exposed to benzo(a)pyrene.
Aquatic Toxicology 94: 309-319.
Nahrgang, J., L. Camus, P. Gonzalez, M. Jnsson, J. S. Christiansen and H. Hop. 2010c.
Biomarker responses in polar cod (Boreogadus saida) exposed to dietary crude oil.
Aquatic Toxicology 96: 77-83.
Nahrgang, J., M. Jnsson and L. Camus. 2010d. EROD activity in liver and gills of polar cod
(Boreogadus saida) exposed to waterborne and dietary crude oil. Marine Environmental
Research 70: 120-123.
NEB, Canada-Newfoundland Offshore Petroleum Board, Canada-Nova Scotia and Offshore
Petroleum Board. 2002. Offshore Waste Treatment Guidelines. [Online]. Available:
http://www.cnsopb.ns.ca/Regframework/regulatory.html. [2011, May 4].
Nedwed, T., G. P. Canevari, R. Belore, J. Clark, T. Coolbaugh and A. Tidwell. 2011a. New
Dispersant Gel Effective on Cold, Viscous Oils (poster). In 2011 International Oil Spill
Conference, Portland, OR.
Nedwed, T., G. P. Canevari and E. Febbo. 2011b. Spreading agents provide a new oil spill
response option In Proceedings of the 2011 International Oil Spill Conference Portland,
OR. 8 pp.
Nedwed, T., W. Spring, R. Belore and D. Blanchet. 2007. Basin-scale testing of ASD icebreaker
enhanced chemical dispersion of oil spills. In Proceedings of the 30th Arctic and Marine
Oilspill Program (AMOP) Technical Seminar, Edmonton, AB, 5-7 June. Environment
Canada, Ottawa ON: 151-160.
Nedwed, T. J., J. P. Smith and M. G. Brandsma. 2004. Verification of the OOC mud and
produced water discharge model using lab-scale plume behavior experiments.
Environmental Modelling & Software 19: 655-670.
242

Neff, J. M. 1987. Biological effects of drilling fluids, drill cuttings and produced waters. In D.F.
Boesch and N.N. Rabalais (eds) Long-Term Effects of Offshore Oil and Gas
Development. Elsevier Applied Science Publishers, London. pp. 469-538.
Neff, J. M. 1990. Composition and fate of petroleum and spill-treating agents in the marine
environment. In J.R. Geraci and D.J. St Aubin (eds.), Sea Mammals and Oil: Confronting
the Risks. Academic Press, San Diego, CA, pp. 1-33.
Neff, J. M. 2002. Bioaccumulation in Marine Organisms. Effects of Contaminants from Oil Well
Produced Water. Elsevier Science Publishers, Amsterdam.
Neff, J. M. 2005. Composition, Environmental Fates, and Biological Effect of Water Based
Drilling Muds and Cuttings Discharged to the Marine Environment: A Synthesis and
Annotated Bibliography. Report prepared for the Petroleum Environment Research
Forum (PERF). Available from American Petroleum Institute, Washington, DC. 73 pp.
Neff, J. M., A. E. Bence, K. R. Parker, D. S. Page, J. S. Brown and P. D. Boehm. 2006a.
Bioavailability of polycyclic aromatic hydrocarbons from buried shoreline oil residues
thirteen years after the Exxon Valdez oil spill: A multispecies assessment. Environmental
Toxicology and Chemistry 25: 947-961.
Neff, J. M. and W. A. Burns. 1996. Estimation of polycyclic aromatic hydrocarbon
concentrations in the water column based on tissue residues in mussels and salmon: An
equilibrium partitioning approach. Environmental Toxicology and Chemistry 15: 2240-
2253.
Neff, J. M., R. E. Hillman, R. S. Carr, R. L. Buhl and J. I. Lahey. 1987. Histopathologic and
biochemical responses in Arctic marine bivalve molluscs exposed to experimentally
spilled oil. Arctic 40: 220-229.
Neff, J. M., S. Johnsen, T. K. Frost, T. I. R. Utvik and G. S. Durell. 2006b. Oil well produced
water discharges to the North Sea. Part II: Comparison of deployed mussels (Mytilus
edulis) and the DREAM model to predict ecological risk. Marine Environmental
Research 62: 224-246.
Neff, J. M., E. H. Owens, S. W. Stoker and D. M. McCormick. 1995. Shoreline oiling conditions
in Prince William Sound following the Exxon Valdez oil spill. In Peter G. Wells, James
N. Butler, Jane S. Hughes (eds.), Exxon Valdez Oil Spill: Fate and Effects in Alaskan
Waters. American Society for Testing and Materials, Philadekphia.
Neff, J. M., D. S. Page and P. D. Boehm. 2011. Exposure of sea otters and harlequin ducks in
Prince William Sound, Alaska, USA, to shoreline oil residues 20 years after the Exxon
Valdez oil spill. Environmental Toxicology and Chemistry 30: 659-672.
Neff, J. M., T. C. Sauer and N. Maciolek. 1989. Fate and Effects of Produced Water Discharges
in Nearshore Marine Waters. API Publication No. 4472. American Petroleum Institute,
Washington, DC, 300 pp.
Nelson, W. G. and A. A. Allen. 1982. The physical interaction and cleanup of crude oil with
slush and solid first year ice. In Proceedings of the 5th Arctic Marine Oil Spill Program
(AMOP) Technical Seminar, Edmonton, Canada. Environment Canada, Ottawa, Ontario,
pp.37-59.
Nichols, W. J. and A. D. Venosa. 2008. Summary of the literature on the use of commercial
bioremediation agents for cleanup of oil-contaminated environments In in Proceedings of
the 2008 International Oil Spill Conference, Savannah, GA, USA: 1275-1280.
243

Nikolopoulou, M. and N. Kalogerakis. 2008. Enhanced bioremediation of crude oil utilizing
lipophilic fertilizers combined with biosurfactants and molasses. Marine Pollution
Bulletin 56: 1855-1861.
NOAA. 2010. Environmental Impact Statement on Effects of Oil and Gas Activities (Seismic and
Exploratory Drilling) in the Arctic Ocean. National Oceanic and Atmospheric
Administration (NOAA), National Marine Fisheries Service, Office of Protected
Resources, Silver Spring, MD, USA.
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, B.C. 201 pp.
NRC. 1983. National Research Council: Drilling Discharges in the Marine Environment.
National Academy Press. Washington, DC. 195pp.
NRC. 1989. National Research Council: Using Oil Spill Dispersant on the Sea, National
Research Council: Using Oil Spill Dispersant on the Sea. National Academy Press,
Washington D.C., pp. 78-80.
NRC. 2003. National Research Council: Oil in the Sea III: Inputs, Fates and Effects. National
Academies Press, Washington, DC. 395 pp.
NRC. 2005. National Research Council: Understanding Oil Spill Dispersants: Efficacy and
Effects. National Academies Press, Washington, D.C.
NRT Science & Technology Committee. 2000. NRT Fact Sheet: Bioremediation in Oil Spill
Response. Environmental Protection Agency.
OHara, T. M. and T. J. OShea. 2005. Assessing impacts of environmental contaminants. In J.E.
Reynolds III, W.F. Perrin, R.R. Reeves, S. Montgomery and T.J. Ragen (eds.), Marine
Mammal Research Conservation Beyond Crisis. Johns Hopkins University Press,
Baltimore, MD, pp. 63-83.
Obermajer, M., K. Dewing and M. G. Fowler. 2010. Geochemistry of crude oil from Bent Horn
field (Canadian Arctic Archipelago) and its possible Paleozoic origin. Organic
Geochemistry 41: 986-996.
OGP. 2002. Aromatics in Produced Water: Occurance, Fate and Effects and Treatments.
International Association of Oil and Gas Producers Publications. Report No. 1.20/324,
24 pp.
OGP. 2004. Environmental Performance in the E&P Industry. International Association of Oil
and Gas Producers 2003 Data Report No. 359. OGP, London, UK. 32 pp.
OGP. 2005. Fate and Effects of Naturally Occurring Substances in Produced Water on the
Marine Environment. International Association of Oil and Gas Producers, Produced
Water Fate & Effect Tast Force, Report No. 364, 42 pp.
Ohtsuka, N., K. Ogiwara, S. Takahashi, K. Maida, Y. Watanabe and H. Saeki. 2001.
Experimental study on spreading of oil under ice floes. In Proceedings of the 16th
International Port and Oceans Engineering Under Arctic Conditions. Vol 3, pp 1333-
1342.
Ohtsuka, N., K. Ohshima, K. Maida, N. Usami and H. Saeki. 1999. Weathering of oil spills in
icy seawater and characteristics of crude oil which trapped into ice floe. In Proceedings
of the 15th International Port and Oceans Engineering Under Arctic Conditions, pp 828-
837.
244

Olsen, G., M. Carroll, P. Renaud, W. Ambrose, R. Olssn and J. Carroll. 2007. Benthic
community response to petroleum-associated components in Arctic versus temperate
marine sediments. Marine Biology 151: 2167-2176.
Olsen, G. H., J. Carroll, E. Sva and L. Camus. 2008. Cellular energy allocation in the Arctic sea
ice amphipod Gammarus wilkitzkii exposed to the water soluble fractions of oil. Marine
Environmental Research 66: 213-214.
Omotoso, O. E., V. A. Munoz and R. J. Mikula. 2002. Mechanisms of crude oil-mineral
interactions. Spill Science & Technology Bulletin 8: 45-54.
Ritsland, N. A., F. R. Engelhardt, F. A. Juck, R. J. Hurst and P. D. Watts. 1981. Effect of Crude
Oil on Polar Bears. Indian and Northern Affairs Canada, Northern Environmental
Protection Branch, Ottawa, Ontario.
Ritsland, N. A. and R. A. Y. Schweinsburg. 1983. Polar bear hunt strategies evaluated by a
Leslie matrix population model. Polar Research 1: 241-248.
Owens, E. H. 1994. Natural cleaning of oiled coarse sediment shorelines in Arctic and Atlantic
Canada. Spill Science & Technology Bulletin 1: 37-52.
Owens, E. H. 1999. The interaction of fine particles with stranded oil. Pure and Applied
Chemistry 71: 83-93.
Owens, E. H., R. A. Davis Jr., J. Michel and K. Stritzke. 1995. Beach cleaning and the role of
technical support in the 1993 Tampa Bay spill. In Proceedings of the 1995 International
Oil Spill Conference, Long Beach, CA, USA: 627-634.
Owens, E. H., J. R. Harper, W. Robson and P. D. Boehm. 1987a. Fate and persistence of crude
oil stranded on a sheltered beach. Arctic 40: 109-123.
Owens, E. H. and K. Lee. 2003. Interaction of oil and mineral fines on shorelines: review and
assessment. Marine Pollution Bulletin 47: 397-405.
Owens, E. H. and W. Robson. 1987. Experimental design and the retention of oil on Arctic test
beaches. Arctic 40: 230-243.
Owens, E. H., W. Robson and C. R. Foget. 1987b. A field evaluation of selected beach-cleaning
techniques. Arctic 40: 244-257.
Owens, E. H., G. A. Sergy, C. C. Guenette, R. C. Prince and K. Lee. 2003. The reduction of
stranded oil by in-situ shoreline treatment options. Spill Science and Technology Bulletin
8: 257-272.
Owens, E. H., G. A. Sergy and R. C. Prince. 2002. The fate of stranded oil at the BIOS site
twenty year after the experiment. In Proceedings of the 25th Arctic Marine Oilspill
Program (AMOP) Technical Seminar, Environment Canada, Ottawa ON, pp. 111.
Page, D. S., P. D. Boehm, G. S. Douglas, A. E. Bence, W. A. Burns and P. J. Mankiewicz. 1996.
The natural petroleum hydrocarbon background in subtidal sediments of Prince William
Sound, Alaska, USA. Environmental Toxicology and Chemistry 15: 1266-1281.
Page, D. S., P. D. Boehm and J. M. Neff. 2010. Comment on "Unlike PAHs from Exxon Valdez
crude oil, PAHs from Gulf of Alaska coals are not readily bioavailable". Environmental
Science & Technology 44: 2210-2211.
Page, D. S., P. D. Boehm, W. A. Stubblefield, K. R. Parker, E. S. Gilfillan, J. M. Neff and A. W.
Maki. 2002. Hydrocarbon composition and toxicity of sediments following the Exxon
Valdez oil spill in Prince William Sound, Alaska, USA. Environmental Toxicology and
Chemistry 21: 1438-1450.
Page, D. S., J. S. Brown, P. D. Boehm, A. E. Bence and J. M. Neff. 2006. A hierarchical
approach measures the aerial extent and concentration levels of PAH-contaminated
245

shoreline sediments at historic industrial sites in Prince William Sound, Alaska. Marine
Pollution Bulletin 52: 367-379.
Paine, R. T., J. L. Ruesink, A. Sun, E. L.Soulanille, M. J. Wonham, C. D. G. Harley, D. R.
Brumbaugh and D. L. Secord. 1996. Trouble on oiled waters: Lessons from the Exxon
Valdez oil spill. Annual Review of Ecology and Systematics 27: 197-235.
Parker, K. R. and A. W. Maki. 2003. Defining recovery and detecting when it occurs. In
Proceedings of the 2003 International Oil Spill Conference, Vancouver, BC, Canada.
Parker, K. R. and J. A. Wiens. 2005. Assessing recovery following environmental accidents:
Environmental variation, ecological assumptions, and strategies. Ecological Applications
15: 2037-2051.
Parsons, J., J. Spry and T. Austin. 1980. Preliminary observations on the effect of Bunker C fuel
oil on seals on the Scotian shelf In Proceedings: Scientific Studies During the Kurdistan
Tanker Incident - A Workshop at Bedford Institute of Oceanography, Dartmouth, NS,
Canada: 193-202.
Patin, S. 1999. Environmental Impact of the Offshore Oil and Gas Industry. Ecomonitor
Publishing, New York, NY.
Payne, J., C. Andrews, S. Whiteway and K. Lee. 2001a. Definition of Sediment Toxicity Zones
Around Oil Development Sites: Dose Response Relationships for the Monitoring
Surrogates Microtox and Amphipods, Exposed to Hibernia Source Cuttings Containing
a Synthetic Base Oil. Canadian Manuscript Report on Fisheries and Aquatic Sciences No.
2577, 10p.
Payne, J., L. Fancey, C. Andrews, J. Meade, F. Power, K. Lee, G. Veinott and A. Cook. 2001b.
Laboratory Exposures of Invertebrate and Vertebrate Species to Concentrations of IA-35
(Petro-Canada) Drill Mud Fluid, Production Water and Hibernia Drill Mud Cuttings.
Canadian Data Repert on Fish and Aquatic Sciences No. 2560. 27p.
Payne, J. F., A. Mathieu and T. K. Collier. 2003a. Ecotoxicological studies focusing on marine
and freshwater fish. In P.E.T. Douben (ed.), Polycyclic Aromatic Hydrocarbons: an
Ecotoxicological Perspective John Whiley and Sons, pp. 192-224.
Payne, J. R., W. B. Driskell, J. W. Short and M. L. Larsen. 2008. Long term monitoring for oil in
the Exxon Valdez spill region. Marine Pollution Bulletin 56: 2067-2081.
Payne, J. R., J. G.D. McNabb and J. J.R. Clayton. 1991. Oil weathering behaviour in Arctic
environments. Polar Research 10: 631-662.
Payne, J. R., J. J.R. Clayton and B. E. Kirstein. 2003b. Oil/suspended particulate material
interaction and sedimentation. Spill Science & Technology Bulletin 8: 201-221.
Pegau, W. S. 2008. Biological Effects of Oil-in-ice in the Arctic. Oil Spill Recovery Institute. 9
pp.
Pegau, W. S. 2011. Biological effects of oil-in-ice in the Arctic. Oil Spill Recovery Institute,
Cordova, AK. [Online]. Available: http://www.pws-
osri.org/business/0802advboard/board_packet.shtml. [2011, June 3], Biological effects of
oil-in-ice in the Arctic. Oil Spill Recovery Institute, Cordova, AK. [Online]. Available:
http://www.pws-osri.org/business/0802advboard/board_packet.shtml. [2011, June 3].
Peishi, Q., S. Zhiguo and L. Yunzhi. 2011. Mathematical simulation on the oil slick spreading
and dispersion in nonuniform flow fields. International Journal of Environmental
Science & Technology 8: 339-350.
246

Pella, J. J. and J. Maselko. 2007. Probability Sampling and Estimation of the Oil Remaining in
2001 from the Exxon Valdez Oil Spill in Prince William Sound, Alaska. U.S. Department
of Commerse, Springfield, VA. 71 pp.
Percy, R., B. Smiley and T. Mullen. 1985. Fishes, Invertebrates and Marine Plants: The
Beaufort Sea and the Search for Oil. Beaufort Sea Project, R.J. Childerhose (ed).
Department of Fisheries and Oceans, Sidney, BC, Canada.
Petersen, D. and I. Dahllf. 2007. Combined effects of pyrene and UV-light on algae and
bacteria in an arctic sediment. Ecotoxicology 16: 371-377.
Peterson, C. H. 2001. The Exxon Valdez oil spill in Alaska: acute, indirect, and chronic effects
on the ecosystem. Advances in Marine Biology 39: 1-103.
Peterson, C. H., S. D. Rice, J. W. Short, D. Esler, J. L. Bodkin, B. E. Ballachey and D. B. Irons.
2003. Long-term ecosystem response to the Exxon Valdez oil spill. Science 302: 2082-
2086.
Pilisi, N., M. Maes and D. B. Lewis. 2011. Deepwater drilling for Arctic oil and gas resources
development: a conceptual study in the Beaufort Sea. In Proceedings of the 2011 Arctic
Technology Conference, Houston TX Feb 7-9, 12 pp.
Pope, G., K. Gordon and J. Bragg. 2011. Using fundamental practices to explain field
observations twenty-one years after the Exxon Valdez Oil Spill. In Proceedings of the
2011 International Oil Spill Conference, Portland, OR, USA. 13 pp.
Potter, S. and I. Buist. 2008. In-Situ Burning for Oil Spills in Arctic Waters: State-of-the-Art and
Future Research Needs. In-Situ Burning for Oil Spills in Arctic Waters: State-of-the-Art
and Future Research Needs, pp. 23-39.
Prince, R. C. 1993. Petroleum spill bioremediation in marine environments. Critical Reviews in
Microbiology 19: 217-242.
Prince, R. C. 2010b. Eukaryotic Hydrocarbon Degraders. Eukaryotic Hydrocarbon Degraders,
pp. 2065-2078.
Prince, R. C. 2010c. Bioremediation of Marine Oil Spills. Bioremediation of Marine Oil Spills,
pp. 2617-2630.
Prince, R. C., R. E. Bare, R. M. Garrett, M. J. Grossman, C. E. Haith, L. G. Keim, G. J. Holtom,
G. A. Sergy, P. Lambert, K. Lee, E. H. Owens and C. C. Gunette. 1999. In-situ
Treatment of Oiled Sediment Shorelines, Volume 3 - Bioremediation of a Marine Oil Spill
in the Arctic. Environment Canada, Edmonton, AB. 57 pp.
Prince, R. C., R. E. Bare, R. M. Garrett, M. J. Grossman, C. E. Haith, L. G. Keim, K. Lee, G. J.
Holtom, P. Lambert, G. A. Sergy, E. H. Owens and C. C. Gunette. 2003a.
Bioremediation of stranded oil on an arctic shoreline. Spill Science and Technology
Bulletin 8: 303-312.
Prince, R. C. and J. R. Clark. 2004. Bioremediation of Marine Oil Spills. Vol.1. Studies in
Surface Science and Catalysis. pp. 495-512.
Prince, R. C., D. L. Elmendorf, J. R. Lute, C. S. Hsu, C. E. Haith, J. D. Senius and G. J. Dechert.
1994. 17alpha(H),21beta(H)-Hopane as a conserved internal marker for estimating the
biodegradation of crude oil. Environmental Science & Technology 28: 142-145.
Prince, R. C., R. M. Garrett, R. E. Bare, M. J. Grossman, T. Townsend, J. M. Suflita, K. Lee, E.
H. Owens, G. A. Sergy, J. F. Braddock, J. E. Lindstrom and R. R. Lessard. 2003b. The
roles of photooxidation and biodegradation in long-term weathering of crude and heavy
fuel oils. Spill Science and Technology Bulletin 8: 145-156.
247

Prince, R. C., A. Gramain and T. J. McGenity. 2010a. Prokaryotic Hydrocarbon Degraders.
Prokaryotic Hydrocarbon Degraders, pp. 1669-1692.
Prince, R. C., E. H. Owens and G. A. Sergy. 2002. Weathering of an Arctic oil spill over 20
years: the BIOS experiment revisited. Marine Pollution Bulletin 44: 1236-1242.
Purves, F. E. E. 1978. The Interaction of Crude Oil and Natural Gas with Laboratory-grown
Saline Ice. EPS4EC789, Environment Canada, Ottawa, ON. 18 pp.
Querbach, K., G. Maillet, P. J. Cranford, C. Taggart, K. Lee and J. Grant. 2005. Potential effects
of produced water discharges on the early life stages of three resource species. In S.
Armsworthy, P.J. Cranford and K. Lee (eds) Offshore Oil and Gas Environmental Effects
Monitoring (Approaches and Technologies). Battelle Press, Columbus, OH. pp. 343-372.
Rabalais, N. N., B. A. McKee, D. J. Reed and J. C. Means. 1991. Fate and Effects of Nearshore
Discharges of OCS Produced Waters. Vol. 1: Executive Summary. Vol. 2: Technical
Report. Vol. 3. Appendices. OCS Studies MMS 91-004, MMS 91-005, and MMS 91-006.
U.S. Dept. of the Interior, Minerals Management Service, Gulf of Mexico OCS Regional
Office, New Orleans, LA.
Ramseier, R. O. and O. Rene. 1973. Possible Fate of Oil in the Arctic Basin In in First World
Congress on Water Resources, The International Water Resources Association, Chicago,
Illinois: 58-70.
Rand, K. M. and E. A. Logerwell. 2011. The first demersal trawl survey of benthic fish and
invertebrates in the Beaufort Sea since the late 1970s. Polar Biology 34: 475-488.
Raymond, J. A. and C. A. Knight. 2003. Ice binding, recrystallization inhibition, and
cryoprotective properties of ice-active substances associated with Antarctic sea ice
diatoms. Cryobiology 46: 174-181.
Reed, M. and O. M. Aamo. 1994. Real time spill forecasting during MIZ93. In Proceedings of
the 17th Arctic and Marine Oilspill Program (AMOP) Technical Seminar, Environment
Canada, Ottawa, pp 785-797.
Reed, M., . Johansen, P. J. Brandvik, P.Daling, A. Lewis, R. Fiocco, D. Mackay and R.
Prentki. 1999. Oil spill modeling towards the close of the 20th century: overview of the
state of the art. Spill Science and Technology Bulletin 5: 3-16.
Reynolds, J. H. and N. Braman. 2009. Using tolerance intervals to assess recovery of mussel
beds impacted by the Exxon Valdez oil spill. Marine Pollution Bulletin 58: 1496-1504.
Ricca, M. A., A. Keith Miles, B. E. Ballachey, J. L. Bodkin, D. Esler and K. A. Trust. 2010. PCB
exposure in sea otters and harlequin ducks in relation to history of contamination by the
Exxon Valdez oil spill. Marine Pollution Bulletin 60: 861-872.
Rice, S. D., R. B. Spies, D. A. Wolfe and B. A. W. (eds). 1996. Proceedings of the Exxon Valdez
Oil Spill Symposium. American Fisheries Society, Bethesda, Maryland.
Rice, S. D., R. E. Thomas, R. A. Heintz, A. C. Heintz, A. C. Wertheim, M. L. Murphy, M. G.
Carls, J. W. Short, A. Moles, R. B. Spies, D. A. Wolfe and B. A. W. (eds). 2001.
Synthesis of Long-term Impacts to Pink Salmon Following the Exxon Valdez Oil Spill:
Persistence, Toxicity, Sensitivity, and Controversy. Exxon Valdez Trustee Council. 76 pp.
Rivkin, R. B., R. Tian, M. R. Anderson and J. F. Payne. 2000. Ecosystem level effects of
offshore platform discharges: Identification, assessment and modelling. In Proceedings of
the 27th Annual Aquatic Toxicity Workshop, St John's, NL, Canada, Canadian Technical
Report of Fisheries and Aquatic Sciences. 2331:3-12.
248

Roberts, A. P., J. T. Oris and W. A. Stubblefield. 2006. Gene expression in caged juvenile Coho
Salmon (Oncorhynchys kisutch) exposed to the waters of Prince William Sound, Alaska.
Marine Pollution Bulletin 52: 1527-1532.
Roberts, P., C. B. Henry, A. Fukuyama and G. Shigenaka. 1999. Weathered petroleum
''bioavailability'' to intertidal bivalves after the T/V Exxon Valdez incident. In Proceedings
of the 1999 International Oil Spill Conference. American Petroleum Institute, Seattle,
WA, USA: 4.
Re Utvik, T. I., G. S. Durell and S. Johnsen. 1999. Determining produced water originating
polycyclic aromatic hydrocarbons in North Sea waters: Comparison of sampling
techniques. Marine Pollution Bulletin 38: 977-989.
Rosenberg, E., A. Perry, D. T. Gibson and D. L. Gutnick. 1979. Emulsifier of Arthrobacter
RAG-1: Specificity of hydrocarbon substrate. Applied and Environmental Microbiology
37: 409-413.
Rytkonen, J., S. Liukkonen and T. Riipi. 1998. Laboratory tests of oil spreading under the ice
cover. In Proceedings of the International Conference on Oil and Hydrocarbon Spills:
Modelling, Analysis and Control. Southampton, UK, pp 155-164.
Saeki, H., M. Sasaki, K. Komatsu, A. Miura and H. Matsuda. 2009. Oil spill remediation by
using the remediation agent JE1058BS that contains a biosurfactant produced by
Gordonia sp strain JE-1058. Bioresource Technology 100: 572-577.
Sayed, M. and S. Lset. 1993. Laboratory experiments of oil spreading in brash ice. In
Proceedings of the 3rd International Offshore and Polar Engineering Conference,
Golden, CO, USA. pp. 224-231.
Schmidt, K. A. G., S. E. Quinones-Cisneros and B. Kvamme. 2005. Density and viscosity
behavior of a North Sea Crude Oil, natural gas liquid, and their mixtures. Energy & Fuels
19: 1303-1313.
Sekerak, A. D. 1982. Young-of-the-year cod (Boreogadus) in Lancaster Sound and Western
Baffin Bay. Arctic 35: 75-87.
Semples, J.-M. 1987. The coastal morphology and sedimentology of Cape Hatt Peninsula. Ambio
40: 10-19.
Sendstad, E. 1980. Accelerated biodegradation of crude oil on Arctic shorelines. In Proceedings
of the Third Arctic and Marine Oilspill Program (AMOP) Technical Seminar.
Environment Canada, Ottawa, Canada. p. 402416.
Sendstad, E., P. Sveum, L. J. Endal, Y. Brattbakk and O. Ronning. 1984. Studies on a seven year
old seashore crude oil spill on Spitsbergen. In Proceedings of the Seventh Arctic Marine
Oilspill Program (AMOP) Technical Seminar. Environment Canada, Ottawa, Canada. p.
6074.
Sergy, G. A. 1999. Key conclusions of the Svalbard shoreline experiment. In Proceedings of the
22nd Arctic and Marine Oilspill Program (AMOP) Technical Seminar, Environment
Canada, Ottawa, pp 845.
Sergy, G. A. and P. J. Blackall. 1987. Design and conclusions of the Baffin Island Oil Spill
Project. Arctic 40: 1-9.
Sergy, G. A., C. C. Guenette and E. H. Owens. 1998. The Svalbard shoreline oilspill field trials.
In Proceedings of the 21st Arctic and Marine Oilspill Program (AMOP) Technical
Seminar, Environment Canada, Ottawa, pp. 873-889.
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.
249

Shata, A. A. 2010. Recovery of Oil Spills by Dispersants in Marine Arctic Regions. Mathematics
and Natural Sciences, Faculty of Science and Technology. University of Stavange,
Stavanger, Norway. 89 pp.
Shaw, D. G., J. W. Farrington, J. S. Conner, B. W. Trippm and J. R. Schubel. 1999. Potential
Environmental Consequences of Petroleum Exploration and Development on Grand
Banks. New England Aquarium Aquatic Forum Series Report 00-3, Boston. 64pp.
Shigenaka, G. 2005. Perspectives on the Exxon Valdez In Proceedings of the 2005 International
Oil Spill Conference, Miami Beach, FL, USA.
Short, J. W., G. V. Irvine, D. H. Mann, J. M. Maselko, J. J. Pella, M. R. Lindeberg, J. R. Payne,
W. B. Driskell and S. D. Rice. 2007. Slightly weathered Exxon Valdez oil persists in Gulf
of Alaska beach sediments after 16 years. Environmental Science & Technology 41:
1245-1250.
Short, J. W., M. R. Lindeberg, P. M. Harris, J. M. Maselko, J. J. Pella and S. D. Rice. 2004.
Estimate of oil persisting on the beaches of Prince William Sound 12 years after the
Exxon Valdez oil spill. Environmental Science & Technology 38: 19-25.
Short, J. W., J. M. Maselko, M. R. Lindeberg, P. M. Harris and S. D. Rice. 2006. Vertical
distribution and probability of encountering intertidal Exxon Valdez oil on shorelines of
three embayments within Prince William Sound, Alaska. Environmental Science &
Technology 40: 3723-3729.
Short, J. W., K. R. Springman, M. R. Lindeberg, L. G. Holland, M. L. Larsen, C. A. Sloan, C.
Khan, P. V. Hodson and S. D. Rice. 2008. Semipermeable membrane devices link site-
specific contaminants to effects: PART II - A comparison of lingering Exxon Valdez oil
with other potential sources of CYP1A inducers in Prince William Sound, Alaska.
Marine Environmental Research 66: 487-498.
Singsaas, I., P. J. Brandvik, P. S. Daling, M. Reed and A. Lewis. 1994. Fate and Behaviour of
Oil spilled in the Rpresence 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.
Siron, R., E. Pelletier and C. Brochu. 1995. Environmental factors influencing the
biodegradation of petroleum hydrocarbons in cold seawater. Archives of Environmental
Contamination and Toxicology 28: 406-416.
Siron, R., E. Pelletier, D. Delille and S. Roy. 1993. Fate and effects of dispersed crude oil under
icy conditions simulated in mesocosms. Marine Environmental Research 35: 273-302.
Skalski, J. R., D. A. Coats and A. K. Fukuyama. 2001. Criteria for oil spill recovery: A case
study of the intertidal community of Prince William Sound, Alaska, following the Exxon
Valdez oil spill. Environmental Management 28: 9-18.
SL Ross Environmental Research, L. 2010. Literature Review of Chemical Oil Spill Dispersants
and Herders in Fresh and Brackish Waters. U.S. Department of the Interior Minerals
Management Service, Herndon, VA. 66 pp.
SL Ross Environmental Research Ltd. 2002. Extending Temporary Storage Capacity Offshore
with Emulsion Breakers. Technology Assessment and Research Division, Minerals
Management Service and Rescue; Safety and Environmental Response, Canadian Coast
Guard, Ottawa, ON. 96 pp.
SL Ross Environmental Research Ltd., DF Dickins Associates LCC. and Envision Planning
Solutions Inc. 2010. Beaufort Sea Oil Spills State of Knowledge Review and Identifcation
250

of Key Issues. Environmental Studies Research Funds Report No.177, Calgary, AB. 126
pp.
SL Ross Environmental Research Ltd. and DF Dickins Associates LLC. 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. 118
pp.
Smith, J. P., M. G. Brandsma and T. J. Nedwed. 2004. Field verification of the Offshore
Operators Committee (OOC) mud and produced water discharge model. Environmental
Modelling & Software 19: 739-749.
Snow, N. B., W. E. Cross, R. H. Green and J. N. Bunch. 1987. The biological setting of the
BIOS site at Cape Hatt, N.W.T., including the sampling design, methodology and
baseline results for macrobenthos. Arctic 40: 80-99.
So, C. M., C. D. Phelps and L. Y. Young. 2003. Anaerobic transformation of alkanes to fatty
acids by a sulfate reducing bacterium strain Hxd3. Applied and Environmental
Microbiology 69: 3892-3900.
Solbakken, J. E., S. Tilseth and K. H. Palmork. 1984. Uptake and elimination of aromatic
hydrocarbons and a chlorinated biphenyl in eggs and larvae of cod Gadus morhua.
Marine Ecology Progress Series 16: 297-301.
Solsberg, L. B., N. W. Glover and M. T. Bronson. 2002. Best available technology for oil in ice.
In Proceedings of the 25th Arctic and Marine Oil Spill Program (AMOP) Technical
Seminar. Environment Canada, Ottawa, Ontario: 25: 1207-1224.
Srstrm, S. E., P. J. Brandvik, I. Buist, P. Daling, D. Dickins, L.-G. Faksness, S. Potter, J. F.
Rasmussen and I. Singsaas. 2010. Joint Industry Program on Oil Spill Contingency for
Arctic and Ice-Covered Waters: Summary Report. Oil In Ice JIP Report no. 32, SINTEF
Materials and Chemistry, Trondheim, Norway. 40 pp.
Southam, G., M. Whitney and C. Knickerbocker. 2001. Structural characterization of the
hydrocarbon degrading bacteria-oil interface: implications for bioremediation.
International Biodeterioration & Biodegradation 47: 197-201.
Spaulding, M. L. 1988. A state-of-the-art review of oil spill trajectory and fate modeling. Oil and
Chemical Pollution 4: 39-55.
Spies, R. B., S. D. Rice, D. A. Wolfe and B. A. Wright. 1996. The effects of the Exxon Valdez
Oil Spill on the Alaskan coastal environment. In Proceedings of the 1996 Exxon Valdez
Oil Spill Symposium. American Fisheries Society Symposium 18: 1-161.
Spring, W., T. Nedwed and R. Belore. 2006. Icebreaker enhanced chemical dispersion of oil
spills. In Proceedings of the 29th Arctic and Marine Oil Spill Program (AMOP)
Technical Seminar, Vancouver, BC, 6-8 June. Environment Canada, Ottawa, ON: 711-
727.
Springman, K. R., J. W. Short, M. Lindeberg and S. D. Rice. 2008. Evaluation of bioavailable
hydrocarbon sources and their induction potential in Prince William Sound, Alaska.
Marine Environmental Research 66: 218-220.
Srinivasan, R., Q. Lu, G. A. Sorial, A. D. Venosa and J. Mullin. 2007. Dispersant effectiveness
of heavy fuel oils using baffled flask test. Environmental Engineering Science 24: 1307-
1320.
St. Aubin, D. J. and J. R. Geraci. 1994. Chapter 21: Summary and Conclusions. In Thomas R.
Loughlin (ed.), Marine Mammals and the Exxon Valdez. Chapter 21: Summary and
251

Conclusions. In Thomas R. Loughlin (ed.), Marine Mammals and the Exxon Valdez.
Academic Press, San Diego, CA, pp. 395.
Stange, K. and J. Klungsr. 1997. Organochlorine contaminants in fish and polycyclic
aromatichydrocarbons in sediments fromthe Barents Sea. ICES Journal of Marine
Science 54: 318-332.
Stapleton, R. D. and G. S. Sayler. 2000. Changes in subsurface catabolic gene frequencies during
natural attenuation of petroleum hydrocarbons. Environmental Science & Technology 34:
1991-1999.
Stepaniyan, O. 2008. Effects of crude oil on major functional characteristics of macroalgae of the
Barents Sea. Russian Journal of Marine Biology 34: 131-134.
Stephenson, M. T., R. C. Ayers, L. J. Bickford, D. D. Caudle, J. T. Cline, G. Cranmer, A. Duff,
E. Garland, T. A. Herenius, R. P. W. M. Jacobs, C. Inglesfield, G. Norris, J. D. Petersen
and A. D. Read. 1994. North Sea Produced Water: Fate and Effects in the Marine
Environment. Report No. 2.62/204. E&P Forum, London, England. 48 pp.
Stige, L. C., G. Ottersen, D. . Hjermann, P. Dalpadado, L. K. Jensen and N. C. Stenseth. 2011.
Environmental toxicology: Population modeling of cod larvae shows high sensitivity to
loss of zooplankton prey. Marine Pollution Bulletin 62: 395-398.
Stochmal, W. and H. Gurgul. 1992. The Crude Oil-Water Emulsion in the Polar Conditions. In
Combatting Marine Oil Spills in Ice and Cold Climates, HELLCOM Seminar, Helsinki,
Finland.
Stoffyn-Egli, P. and K. Lee. 2002. Formation and characterization of oil-mineral aggregates.
Spill Science & Technology Bulletin 8: 31-44.
Stout, S. A. and Z. Wang. 2008. Chemical fingerprinting of spilled or discharged petroleum -
methods and factors affecting petroleum fingerprints in the environment. In Wang, Z. and
Stout, S. A. (eds.), Oil Spill Environmental Forensics: Fingerprinting and Source
Identification. Elsevier, Burlington, MA, 1-72.
Strmgren, T., S. E. Srstrm, L. Schou, I. Kaarstad, T. Aunaas, O. G. Brakstad and .
Johansen. 1995. Acute toxic effects of produced water in relation to chemical
composition and dispersion. Marine Environmental Research 40: 147-169.
Stubblefield, W. A., G. A. Hancock, H. H. Prince and R. K. Ringer. 1995. Effects of naturally
weathered Exxon Valdez crude oil on mallard reproduction. Environmental Toxicology
and Chemistry 14: 1951-1960.
Sveum, P. and A. Ladousse. 1989. Biodegradation of oil in the Arctic: enhancement by oil-
soluble fertilizer application. In Proceedings of the 1989 International Oil Spill
Conference. American Petroleum Institute, Washington, D.C.: 439446.
Swannell, R. P., K. Lee and M. McDonagh. 1996. Field evaluations of marine oil spill
bioremediation. Microbiological Reviews 60: 342-365.
Swannell, R. P. J. and F. Daniel. 1999. Effect of dispersants on oil biodegradation under
simulated marine conditions. In Proceedings of the 1999 International Oil Spill
Conference, Seattle, WA, USA: 11.
Taylor, E. and D. Reimer. 2008. Oil persistence on beaches in Prince William Sound - A review
of SCAT surveys conducted from 1989 to 2002. Marine Pollution Bulletin 56: 458-474.
Terrens, G. W. and R. D. Tait. 1993. Effects on the Marine Environment of Produced Formation
Water Discharges from Esso/BHPP's Bass Strait Platforms. Esso Australia Ltd.,
Melbourne, Australia. 25 pp.
252

Terrens, G. W. and R. D. Tait. 1996. Monitoring ocean concentrations of aromatic hydrocarbons
from produced formation water discharges to Bass Strait, Australia In Proceedings of the
1996 International Conference on Health, Safety & Environment. Society of Petroleum
Engineers, Richardson, TX: 739-747.
Thomas, J., Y. JR, A. JA and A. M. 1986. Rates of dissolution and biodegradation of water-
insoluble organic compounds. Applied and Environmental Microbiology 52: 290-296.
Thomas, R. E., M. Lindeberg, P. M. Harris and S. D. Rice. 2007. Induction of DNA strand
breaks in the mussel (Mytilus trossulus) and clam (Protothaca staminea) following
chronic field exposure to polycyclic aromatic hydrocarbons from the Exxon Valdez spill.
Marine Pollution Bulletin 54: 726-732.
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 Science 65: 44-50.
Townsend, G. T., R. C. Prince and J. M. Suita. 2003. Anaerobic oxidation of crude oil
hydrocarbons by the resident microorganism of a contaminated anoxic aquifer.
Environmental Science & Technology 37: 5213-5218.
Trudel, B. K., R. C. Belore, A. Guarino, A. Lewis and J. Mullin. 2005. Determining the viscosity
limits for effective chemical dispersion: Relating OHMSETT results to those from tests
at-sea. In Proceedings of the 2005 International Oil Spill Conference, Miami Beach, FL,
USA: 10687-10692.
Trudel, K., R. C. Belore, J. V. Mullin and A. Guarino. 2010. Oil viscosity limitation on
dispersibility of crude oil under simulated at-sea conditions in a large wave tank. Marine
Pollution Bulletin 60: 1606-1614.
Uzuner, M. S., F. B. Weiskopf, J. C. Cox and L. A. Schultz. 1979. Transport of Oil Under
Smooth Ice. US Environmental Protection Agency, Report number EPA-600/3-79-041.
Valentine, D. L., J. D. Kessler, M. C. Redmond, S. D. Mendes, M. B. Heintz, C. Farwell, L. Hu,
F. S. Kinnaman, S. Yvon-Lewis, M. Du, E. W. Chan, F. G. Tigreros and C. J. Villanueva.
2010. Propane respiration jump-starts microbial response to a deep oil spill. Science 330:
208-211.
Van den Heuvel Greve, M. and M. Koopmans. 2007. Safety at Sea: Development of a Simple
Classifcation Scheme for the Toxicity of Crude Oils. Report No A3 Revision No 6. 29 pp.
Vefsnmo, S. and B. O. Johannessen. 1994. Experimental oil spill in the Barents Sea - Drift and
spread of oil in broken ice. In Proceedings of the 17th Arctic and Marine Oilspill
Program (AMOP) Technical Seminar. Environment Canada. Ottawa, Ontario: 1331-
1343.
Veil, J. A. 2006. Comparison of two international approaches to controlling risk from produced
water discharges. Paper presented at the 70th PERF Meeting, Paris, France, March 21-
22, 2006.
Veil, J. A., C. J. Burke and D. O. Moses. 1995. Synthetic Drilling Fluids - A Pollution
Prevention Opportunity for the Oil and Gas Industry. Argonne National Laboratory,
Washington, DC. 11 pp.
Velez, P., H. G. Johnsen, A. Steen and Y. Osikilo. 2011. Advancing oil spill preparedness and
response techniques for Arctic conditions. In Proceedings of the 2011 International Oil
Spill Conference, Portland, OR, USA. 9 pp.
Venkatesh, S., H. El-Tahan, G. Comfort and R. Abdelnour. 1990. Modeling the behavior of oil
spills in Ice-infested waters. Atmosphere-Ocean 28: 303-329.
253

Venosa, A. D., P. Campo and M. T. Suidan. 2010. Biodegradability of lingering crude oil 19
years after the Exxon Valdez oil spill. Environmental Science & Technology 44: 7613-
7621.
Venosa, A. D., K. Lee, M. T. Suidan, S. Garcia-Blanco, S. Cobanli, M. Moteleb, J. R. Haines, G.
Tremblay and M. Hazelwood. 2002. Bioremediation and biorestoration of a crude oil
contaminated freshwater wetland on the St. Lawrence River. Bioremediation Journal 6:
261 - 281.
Venosa, A. D., M. T. Suidan, B. A. Wrenn, K. L. Strohmeier, J. R. Haines, B. L. Eberhart, D.
King and E. Holder. 1996. Bioremediation of an experimental oil spill on the shoreline of
Delaware Bay. Environmental Science & Technology 30: 1764-1775.
Venosa, A. D. and X. Zhu. 2003. Biodegradation of crude oil contaminating marine shorelines
and freshwater wetlands. Spill Science and Technology Bulletin 8: 163-178.
von Ziegesar, O. G., E. Miller, E. Dahlheim and M. E. Loughlin. 1994. Impacts on humpback
whales in Prince William Sound. pp. 173-192. In Loughlin, T.R. (ed.), Marine Mammals
and the Exxon Valdez. Academic Press, San Diego, CA. xx + 395 pp.
Wang, K. G., M. Lepparanta, M. Gastgifvars, J. Vainio and C. X. Wang. 2008. The drift and
spreading of the Runner 4 oil spill and the ice conditions in the Gulf of Finland, winter
2006. Estonian Journal of Earth Sciences 57: 181-191.
Wang, Z. and M. Fingas. 1995. Differentiation of the source of spilled oil and monitoring of the
oil weathering process using gas chromatography-mass spectrometry. Journal of
Chromatography A 712: 321-343.
Weerasuriya, S. and P. Yapa. 1993. Unidirectional spreading of oil under ice. Canadian Journal
of Civil Engineering 20: 50-56.
Weis, J. S. and P. Weis. 1989. Effects of environmental pollutants on early fish development.
Review in Aquatic Sciences 1: 45-73.
Wells, P. G., C. Blaise and K. Lee. 1998. Microscale Aquatic Toxicology: Advances, Techniques
and Practice. P.G. Wells, K. Lee and C. Blaise (eds.), CRC Press, Incorporated. 679 pp.
Wells, P. G., J. N. Butler and J. S. H. (eds.). 1995. Exxon Valdez Oil Spill: Fate and effects in
Alaskan Waters. American Society for Testing and Materials, Philadelphia, USA.
Wertheimer, A. C., R. A. Heintz, J. F. Thedinga, J. M. Maselko and S. D. Rice. 2000. Straying of
adult pink salmon from their natal stream following embryonic exposure to weathered
Exxon Valdez crude oil. Transactions of the American Fisheries Society 129: 989 - 1004.
Weslawski, J. M., M. Zajaczkowski, S. Kwasniewski, J. Jezierski and W. Moskal. 1988.
Seasonality in an Arctic fjord ecosystem: Hornsund, Spitsbergen. Polar Research 6: 185-
189.
Whyte, L. G., L. Bourbonniere and C. W. Greer. 1997. Biodegradation of petroleum
hydrocarbons by psychrotrophic Pseudomonas strains possessing both alkane (alk) and
naphthalene (nah) catabolic pathways. Applied and Environmental Microbiology 63:
3719-3723.
Whyte, L. G., B. Goalen, J. Hawari, D. Labbe, C. W. Greer and M. Nahir. 2001. Bioremediation
treatability assessment of hydrocarbon-contaminated soils from Eureka, Nunayut. Cold
Regions Science and Technology 32: 121-132.
Whyte, L. G., A. Schultz, J. B. van Beilen, A. P. Luz, V. Pellizari, D. Labbe and C. W. Greer.
2002. Prevalence of alkane monooxygenase genes in Arctic and Antarctic hydrocarbon-
contaminated and pristine soils. FEMS Microbiology Ecology 41: 141-150.
254

Whyte, L. G., S. J. Slagman, F. Pietrantonio, L. Bourbonniere, S. F. Koval, J. R. Lawrence, W.
E. Inniss and C. W. Greer. 1999. Physiological adaptations involved in alkane
assimilation at a low temperature by Rhodococcus sp strain Q15. Applied and
Environmental Microbiology 65: 2961-2968.
Wick, L. Y., P. Wattiau and H. Harms. 2002. Influence of the growth substrate on the mycolic
acid profiles of mycobacteria. Environmental Microbiology 4: 612-616.
Widdel, F. and R. Rabus. 2001. Anaerobic biodegradation of saturated and aromatic
hydrocarbons. Current Opinion in Biotechnology 12: 259-276.
Wiens, J. A. 2007. Applying ecological risk assessment to environmental accidents: Harlequin
ducks and the Exxon Valdez Oil Spill. BioScience 57: 769-777.
Wilkinson, J. P., B. Wadhams and N. E. Hughes. 2007a. A new technique to determine the
spread of oil spill under fast ice. In Proceedings of the 19th International Port and
Oceans Engineering Under Arctic Conditions. Vol. 2, Dalian University of Technology.
Dalian, China: 855-867.
Wilkinson, J. P., P. Wadhams and N. E. Hughes. 2007b. Modelling the spread of oil under fast
sea ice using three-dimensional multibeam sonar data. Geophysical Research Letters 34:
L22506.
Willette, M. 1996. Impacts of the Exxon Valdez oil spill on the migration, growth, and survival
of juvenile pink salmon in Prince William Sound. In Proceedings from the 1996 Exxon
Valdez Oil Spill Symposium. American Fisheries Society, Bethesda, MD, USA: 931.
Wohlgeschaffen, G., K. H. Mann, D. V. S. Rao and R. Pocklington. 1992. Dynamics of the
phycotoxin domoic acid: accumulation and excretion in two commercially important
bivalves. Journal of Applied Phycology 4: 297-310.
Wong, D., M. Lyons, L. Burridge and K. Lee. 2011. Fertilisation and hatching success of
Atlantic cod (Gadus morhua) eggs when exposed to various concentrations of produced
water. In Fletcher, T., Holdway, D., Simmons, D., Dutton, M., Burridge, L.E. (eds)
Proceedings of the 37th Annual Aquatic Toxicity Workshop, October 3 to 6. Canadian
Technical Report of Fisheries and Aquatic Sciences, Toronto, ON. (in press).
Word, J. Q. and R. A. Perkins. 2011. Toxicology and Biodegradation of Crude and Dispersed
Oil in the Arctic Marine Environment. Joint Industry Program Research, 14pp.
Wrenn, B., K. L. Sarnecki, E. S. Kohar, K. Lee and A. D. Venosa. 2006. Effects of nutrient
source and supply on crude oil biodegradation, in continuous-flow beach microcosms.
Journal of Environmental Engineering 132: 75-84.
Wrenn, B. A., J. R. Haines, A. D. Venosa, M. Kadkhodayan and M. T. Suidan. 1994. Effects of
nitrogen source on crude oil biodegradation. Journal of Industrial Microbiology 13: 279-
286.
Wrenn, B. A., M. T. Suidan, K. L. Strohmeier, B. L. Eberhart, G. J. Wilson and A. D. Venosa.
1997a. Nutrient transport during bioremediation of contaminated beaches: Evaluation
with lithium as a conservative tracer. Water Research 31: 515-524.
Wrenn, B. A., M. T. Suidan, K. L. Strohmeier, B. L. Eberhart, G. J. Wilson, A. D. Venosa, J. R.
Haines and E. Holder. 1997b. Influence of tide and waves on washout of dissolved
nutrients from the bioremediation zone of a coarse-sand beach: Application in oil-spill
bioremediation. Spill Science & Technology Bulletin 4: 99-106.
WWF. 2007. Oil Spill Response Challenges in Arctic Waters, Oil Spill Response Challenges in
Arctic Waters. World Wildlife Fund International Arctic Programme, Oslo, Norway.
255

Xia, Y., H. Li, M. C. Boufadel and Y. Sharifi. 2010. Hydrodynamic factors affecting the
persistence of the Exxon Valdez oil in a shallow bedrock beach. Water Resources
Research 46: W10528.
Xia, Y. Q. and M. C. Boufadel. 2010. Lessons from the Exxon Valdez Oil Spill disaster in
Alaska. Disaster Advances 3: 270-273.
Yakimov, M. M., K. N. Timmis and P. N. Golyshin. 2007. Obligate oil-degrading marine
bacteria. Current Opinion in Biotechnology 18: 257-266.
Yapa, P. D. and ASCE Task Committee on Modelling of Oil Spills. 1996. State-of-the-art review
of modeling transport and fate of oil spills. Journal of Hydraulic Engineering 122: 594-
609.
Yapa, P. D. and D. P. Belaskas. 1993. Radial spreading of oil under and over broken ice: an
experimental study. Canadian Journal of Civil Engineering 20: 910-922.
Yapa, P. D. and T. Chowdhury. 1990. Spreading of oil spilled under ice. Journal of Hydraulic
Engineering 116: 1468-1483.
Yapa, P. D. and L. Dasanayaka. 2006. State-of-the-art review of modelling oil transport and
spreading in ice covered waters. In Proceedings of the 29th Arctic and Marine Oilspill
Program (AMOP) Technical Seminar, Vol. 2. Environment Canada, Ottawa: 893-909.
Yergeau, E., M. Arbour, R. Brousseau, D. Juck, J. R. Lawrence, L. Masson, L. G. Whyte and C.
W. Greer. 2009. Microarray and real-time PCR analyses of the responses of high-Arctic
soil bacteria to hydrocarbon pollution and bioremediation treatments. Applied and
Environmental Microbiology 75: 6258-6267.
Yeung, C. W., K. Lee and C. W. Greer. 2010. Microbial community characterization of
produced water from the Hibernia oil production platform. In K. Lee and J. Neff (eds.),
Environmental Risks and Mitigation Technologies for Produced Water. Springer
Publishing Company (in Press). New York, NY.
Zhang, H., M. Khabiti, Y. Zheng, K. Lee, Z. Li and J. Mullin. 2009. Investigation of OMA
formation and the effect of minerals. Marine Pollution Bulletin Submitted.
Zhang, J.-j. and X. Liu. 2008. Some advances in crude oil rheology and its application. Journal
of Central South University of Technology 15: 288-292.
ZoBell, C. E. 1973. Microbial degradation of oil: present status, problems, and perspectives. In
Ahearn, D. G. and Meyers, S. P. (eds.), The Microbial Degradation of Oil Pollutants.
Center for Wetland Resources, Louisiana State University. Baton Rouge, LA. pp. 3-16.

256

257
Appendix 1

Table 13 Sources of peer-reviewed biological effects data from Camus et al. (Camus et al., 2008) with additional references added.
Taxonomic
group
Species Habitat Chemical Treatment Endpoint Reference
Mysis oculata epi benthic dispersed oil and WAF acute LC
50
Riebel & Percy 1989
Mysids
Mysis relucta epi benthic WAF crude oil acute
median lethal
concentrations
Carls & Korn 1985
Gammarus wilkitzkii sea ice WAF crude oil chronic cellular energy allocation Olsen et al. 2008
Gammarus wilkitzkii sea ice WAF crude oil chronic embryo malformation Camus et al. 2008
Gammarus zaddachi shore oil acute survival Budosh & Atlas 1977
Anonyx nugax shore/benthic WAF crude oil acute
median lethal
concentrations
Carls & Korn 1985
Onisimus litoralis shore WAF crude oil chronic cellular energy allocation Olsen et al. 2007
Onisimus nanseni sea ice WAF crude oil acute
median lethal
concentrations
Carls & Korn 1985
Gammaracanthus
loricatus
sea ice WAF crude oil acute
median lethal
concentrations
Carls & Korn 1985
Onisimus affinis shore oil sediment acute behaviour/avoidance Percy 1977
Onisimus affinis shore dispersed oil acute respiration Percy 1977
Onisimus affinis shore oil acute survival Budosh & Atlas 1977
Onisimus affinis shore dispersed oil acute locomotory activity Percy & Mullin 1977
Onisimus affinis shore WAF crude oil chronic
survival, movement and
food search success
Budosh 1981
Corophium
clarencense
shore oil sediment acute behaviour/avoidance Percy 1977
Gammarus setosus shore WAF crude oil chronic cellular energy allocation Olsen et al. 2007
Gammarus
oceanicus
shore WAF crude oil chronic
respiration,
osmoregulation
Aunaas et al. 1991
Gammarus wilkitzkii sea ice WAF crude oil chronic
respiration cellular
energy allocation
Hatlen et al. 2009
Amphipod
Gammarus wilkitzkii sea ice WAF crude oil acute cellular energy allocation Olsen et al. 2008
Cnidaria Halitholus cirratus shore dispersed oil acute locomotory activity Percy & Mullin 1977
Isopod Mesidotea sibirica shore oil sediment acute behaviour/avoidance Percy 1977


Mesidotea entomon shore oil sediment acute behaviour/avoidance Percy 1977
Calanus glacialis pelagic pyrene acute Feeding & reproduction Holst et al. 2007
Microsetella
norvegica
epi benthic pyrene acute
Survival feeding
RNA/DNA
Hjorth & Dahlff 2008
Calanus
finmarchicus
pelagic pyrene acute feeding & reproduction Holst et al. 2007
Calanus
finmarchicus
pelagic WAF crude oil acute Feeding & reproduction Jensen and Carroll 2010
Calanus
finmarchicus
pelagic WAF crude oil acute LC
50
Hansen et al. 2011
Copepod
Calanus glacialis pelagic WAF crude oil acute LC
50
Hansen et al. 2011
Mya truncata mud oil in sediment chronic
respiration/membrane
stability/oxidative stress
Camus et al. 2003
Mya truncate mud dispersed oil field
ostial closure/ siphon
retraction/ stimulus
response siphon/ mantle
retraction
Mageau et al. 1987
Chlamys islandica epi bentic benzo(a)pyrene acute
membrane stability &
oxidative stress
Camus et al. 2002
Liocyma fluctuosa mud dweller WAF crude oil chronic cellular energy allocation Olsen et al. 2007
Serripes
groenlandicus
mud dweller dispersed oil field
burial activity/ ostial
closure/ mantle
gap/stimulus response
siphon
Mageau et al. 1987
Mytilus edulis rocky shore phenanthrene chronic Lysosomal stability Camus et al. 2000
Mytilus edulis rocky shore oil dispersant chronic
freezing tolerance and
osmoregulation
Aarset & Zachariassen
1983
Bivalves
Chlamys islandica epi benthic dispersed oil acute immune response Hannam et al. 2009
Echinoderms
Strongylocentrotus
Droebachiensis
epi benthic dispersed oil field
cover/ attachment/
morphometrics
Mageau et al. 1987
Decapod Hyas araneus epi benthic
oil in sediment/ oil
injected
chronic
heart rate/ respiration/
oxidative stress
Camus et al. 2002
Boreogadus saida sea ice WAF crude oil acute
median lethal
concentrations
Carls & Korn 1985
Boreogadus saida sea ice xylene acute LC
50
Honkanen 2007
Boreogadus saida sea ice
injected radio labelled
benzo(a) pyrene
chronic
excretion/ whole body
radiography
Ingebritsen et al. 2000
Fish
Boreogadus saida sea ice injected acute DNA adducts Aass et al. 2004
258

259
Oncocottus
hexacornis
benthic WAF crude oil acute
median lethal
concentrations
Carls & Korn 1985
Anarhichas minor benthic waternorme oil sublethal EROD in gills Jonsson et al. 2003
Fragilariopsis
oceanica
water column xylene acute LC
50
Kiel Jensen 2007
Chaetoceros
septentrionalis
pelagic crude oil acute growth Hsiao 1970
Navicula
bahusiensis
pelagic crude oil acute growth Hsiao 1970
Chlamydomonas
pulsatilla
pelagic crude oil acute growth Hsiao 1970
Nitzchia
delicatissima
pelagic crude oil acute growth Hsiao 1970
Laminaria
saccharina
benthic crude oil & corexit acute growth Hsiao et al. 1970
Phyllophora truncate benthic crude oil & corexit acute growth Hsiao et al. 1970
Phytoplankton
Community
benthic pyrene+UV chronic
14C incorporation/ NH4+/
Nitrate & silicate uptake
Petersen 2007
Phytoplankton
Community
benthic pyrene+UV chronic
14C incorporation/ NH4+/
Nitrate & silicate uptake
Petersen &Dahllf 2007
Community sediment crude oil chronic respiration Olsen et al. 2007
Community sediment crude oil chronic bioturbation Olsson 2006
Algae
Community sediment crude oil chronic diversity index Gulliksen & Taasen 1982
Community benthic oil acute nitrogen fixation Knowles & Wishart
Community benthic oil chronic
nitrogen fixation/
denitrification/CO2
Production/methane
concentration
Griffiths et al. 1982 Bacteria
Community sea ice oil chronic diversity index Gerdes et al. 2005

You might also like