Add notes to:

Or add a reference to download later

Note added.

0.0 EXECUTIVE SUMMARY

Photo E-1. Summer storm in the Arctic
Photo E-1. Summer storm in the Arctic

0.1 Program Objectives and Participants

The International Association of Oil and Gas Producers (OGP) in support of the Arctic Oil Spill Response Technology – Joint Industry Programme (JIP) funded this review of available information on the environmental impacts from oil spills in the Arctic and impacts that may be associated with the application of specific treatment technologies that may be applied during an oil spill response.  The objective of the review was to compile significant findings of prior investigations and suggest priority areas of work needed to improve assessment of the consequences of the various treatment strategies prior to application under Arctic conditions. The primary outcome is development of a process to integrate ecological consequence assessments within net environmental benefit analyses (NEBA) to better evaluate the environmental effects to valuable ecosystem components (VECs) within key environmental compartments (ECs) c that would result from using different response actions in the Arctic.  Specifically, development of Arctic response consequence analysis table matrices (ARCAT) and a semi-quantitative analytical tool will optimize decision-making and lessen environmental impact related to arctic oil spills in the Arctic.  The process is proactive rather than responsive to a spill event improving the ability to reduce environmental damage by chosing mitigating measures with the minimized effect on the environment.  As a result of this proactive approach there will also be broader consensus and acceptance of decisions among regulators and stakeholders.

To increase confidence in evaluations and environmental impact statement (EIS) of oil spills and oil spill response (OSR) technologies proposed for the Arctic, a comprehensive review of the existing literature was conducted to identify priority areas for future research efforts.  Because of the breadth of topic areas supporting consequence analysis approaches, this effort called for a multidisciplinary team with an understanding of food webs throughout the Arctic; behavior of oil in surface waters, at depth, and in ice; the effectiveness of OSR countermeasures in cold-water surface and subsurface environments; the toxicity of petroleum treatment residues on Arctic species; and models used to predict individual and population effects in Arctic ecosystems based on the concept of resilience.  This literature review followed a pan-arctic approach that recognizes regional similarities and differences based on peer-reviewed literature and technical reports from government and research institutions representing circumpolar interests.  The review is based on more than 650 literature citations and the personal experience of the work group participants.  To best utilize the expertise assembled on this team, representatives from different regions and disciplines worked in subgroups and all were able to review the entire document since many of the investigators have multi-disciplinary research backgrounds. The team was divided into nine technical working groups, each with a focused assessment goal and each with the assignment to identify areas of priority research:

  • The Physical Environment – Summarize available literature to describe the different environmental compartments present in the Arctic and the physical characteristics of those compartments that could potentially affect the fate and effects of oil, as well as defining what response measures are an option for those environment compartments.  Seasonality in Arctic eco-regions is addressed (ice or ice-free conditions) as well as unique habitats and use of those compartments by VECs (e.g. areas with significant riverine input).  
  • Arctic Ecosystems and Valuable Resources – Identify environmental compartments, food web connections within those environmental compartments, key Valuable Ecosystem Components (VEC species/taxa) within each compartment that connects those food webs and variations among VEC species/taxa between sub-regions within the Arctic.
  • Transport and Fate of Oil in the Arctic – The various treatment strategies may enhance or reduce the amount of oil that is transported away from sites of surface or subsea released oil.  This section discusses the mechanisms behind short and long term transport of spilled oil in the Arctic.
  • Oil Spill Response and Related Effects – Identify the implications of various response actions to increase or decrease the exposure to various treatment residuals within environmental compartments. Identify alternative response options for surface and subsurface spills of oil under Arctic seasonal conditions and describe the environmental effects of those response actions.
  • Biodegradation – Identify the measurement tools that can quantify the use of oil by Arctic microbes including direct uptake and use of carbon containing petroleum compounds, mineralization of petroleum hydrocarbons to CO2 and chemical changes of parent petroleum compounds to other compounds under different treatment options and how those changes are modified in different ECs
  • Ecotoxicology of Oil and Treated Oil in the Arctic – The four lines of ecotoxicity research include acute and chronic response assessments and biomarker and body burden assessments of exposure.  Evaluate the literature to establish the environmental relevance of these four lines of evidence.  Identify gaps in experimental programs to address the ecotoxicology of oil in separate environmental compartments.
  • Population Effects Modeling – Petroleum treatment residuals are environmental stressors that have the potential to produce adverse impacts on individuals or populations of VECs.  Determine which parameters are necessary components of a population impact model for VECs; identify models useful in predicting the effects of treatment options on these species/taxa.  Parameters of interest include changes in the toxicity of biodegraded oil, transfer of impact related effects to other trophic level VECs, or consideration of the resiliency of populations to recover from a stress.
  • Ecosystem Recovery – The ecotoxicology and biodegradation of oil spill residuals is better understood than the ability of a compartment or VEC to recover from the impacts of a spill.  Environmental compartments have different abilities to recover from oil impacts.  Application of treatment methods need to account for and minimize the movement of oil to locations or species that have a low resiliency to respond to oil.  This section examines the duration of continued impact of an oil release and expected recovery of different environmental compartments and VECs. Response actions need to minimize encroachment on less resilient ECs and VECs.
  • Net Environmental Benefit Analyses for Oil Spill Response Options in the Arctic – The environmental consequences of an oil spill response strategy will influence the overall impact of an oil spill (surface or subsurface).  In typical applications of net environmental benefit analysis (NEBA) framework, the consequences of a treatment option are generally associated with the near-term impacts and less so with consideration of the longer term consequences to recovery resulting from impacts to less resilient environmental compartments or VECs.  As an example, far-field impacts can result from not selecting a dispersant option which result in stranding of untreated oil on cobble shorelines where the effects can be observed decades after a spill.  Alternatively, selection of the dispersant option will result in greater impacts to the more resilient zooplankton community while reducing the effects on less resilient seabird and marine mammals that might contact oil on the sea surface.  The goal of this workgroup was to develop a preliminary consequence analysis of treatment decisions and demonstrate through an adaptive NEBA process the interrelationships between treatment decisions and shortening or lengthening of recovery from all forms of environmental effects, not restricted to the immediate near-field impacts.

       

Recent literature was reviewed for each topic area, and the information was compiled into a Microsoft Access® database.  Approximately 650 documents have been cited within the body of this report, and at this time there are over 1300 documents represented in the Access database that has been produced for recent JIP programs.  The database is searchable and can generate a list of references related to Arctic ecology, food webs, deep water food webs, petroleum-related fate and effects, biodegradation, and ecosystem-level case studies; aquatic toxicity and chemistry data can be queried via pivot graphs to dynamically represent the data compiled to date.   Based on results of these summary investigations, recommended new environmental studies are identified and prioritized. The recommended research will help reduce uncertainties related to each topic area and strengthen the NEBA approach in the Arctic.  The organizations and team members participating in this synergistic effort are presented in Figure E-1.

Figure E-1. Organizational matrix (Workgroup chairmen and co-chairmen are noted in bold text)
Figure E-1. Organizational matrix (Workgroup chairmen and co-chairmen are noted in bold text)

0.1.1 The Pan-Arctic Region: Highlights of the Literature Review

Currently five of the eight countries bordering the polar region are pursuing exploration and/or development of oil and gas resources in the Arctic [Canada, Greenland (Denmark), Norway, Russia, and the United States]. The changing environmental conditions in the Arctic may provide increased opportunity for development of these resources that were less accessible due to presence of ice in past decades and the improved technological advances for extracting petroleum resources. Activities of the petroleum industry are based on promulgated regulations set by each sovereign nation but there has been a move toward international cooperation and sharing of knowledge related to the technological development required to ensure safe drilling operations as well as spill response preparedness. International, federal, and local agencies from North America, Northern Europe, and Russia are in the process of developing baseline ecosystem and biodiversity assessments and research programs in order to better understand and protect the Arctic marine ecosystem and the communities that rely on these resources. For example, the Arctic Monitoring and Assessment Program (AMAP) has completed a significant effort in publishing comprehensive baseline information on Arctic geo-political activities and regulations, available drilling technologies, spill response initiatives, and potential environmental impacts (AMAP 2010). Additional cooperative research sponsored by joint industry programs (JIP) have augmented the knowledge base associated with the oil and gas industry activities (Sørstrøm et al. 2010; NewFields 2012). These efforts and convened workshops have integrated contributions from the scientific community, governmental agencies, public interest groups, and indigenous people of the Arctic.

Additionally, extensive field and laboratory studies have been conducted to examine the behavior and fate of oil and its potential effects on Arctic resources under the disparate seasonal conditions. Many recent studies have concentrated on understanding the influence of these harsh environmental conditions on the relative sensitivity of Arctic species to additional stressors, the success and rates of microbial degradation of oil compounds, and more recently the resilience of Arctic populations to recover from responses to those stressors. Similar to other parts of the world these investigations have increased our understanding of the basic behavior and movement of oil, its potential effects on VECs and the ultimate fate of released oil in various environmental compartments. The Arctic environment has added complexities resulting from seasonal patterns of ice and light that need to be considered to provide the foundation for development of strategic spill response strategies and evaluation of the environmental consequences of released oil that are very relevant today. Excellent comprehensive reviews have been published in recent years (Potter et al. 2012, Lee et al. 2012, SL Ross et al. 2010; USGS 2011).

The fundamental role of comparing the adverse biological effects of different response options in NEBA requires an information base that identifies VECs within multiple environmental compartments. The potential adverse effects and resiliency of these VEC organisms within each of these various compartments are then compared as a consequence of the OSR actions. These comparisons should examine the acute and long-term effects of spilled oil resulting from the impacts of various response options such as natural attenuation, surface-applied or subsea-injected dispersants, in-situ burning, and mechanical or naturally occurring containment methods followed up by recovery of spilled petroleum in Arctic ecosystems. Review and tabulation of published data, such as toxicity effect concentrations and population recovery times, is a key component of this review. However, the overall objective related to exploration and production projects in the Arctic is not only to tabulate this information or determine the most sensitive end-points that might be considered but to demonstrate the relative differences in the magnitude and duration of effects that might be observed at various ecosystem compartments associated with various response actions. As has been attempted for other regions where new exploration and production activities have been implemented, demonstrating these relative differences will require additional effort aimed at bridging the gap between relatively straightforward measurements of toxicity to arctic species and more complex investigations to assess ecosystem-level or population level impacts and recovery dynamics. This report considers the similarities and differences in species sensitivity between arctic and non-arctic species. To effectively monitor habitat recovery and identify ecologically relevant endpoints for remediation operations in Arctic regions there is a need to increase our knowledge on natural variability among populations and how that variability relates to vulnerability to petroleum exposure.

Prior to the granting of approvals for exploration and production activities, the public seeks increased assurances that industry and the various governmental entities have the capability to ensure safe exploration and extraction of oil as well as the capability to respond to oil spills. To meet these challenges a number of current and emerging oil spill countermeasure technologies have been identified for use in the Arctic. While use of different OSR methods can potentially reduce the impact of spills within the Arctic under various environmental conditions, not all options have been readily accepted by the public and regulators. Lack of endorsement of some OSR options is related to the perceived change in impacts and biodegradation rates that the use of these options in the Arctic marine ecosystem may bring. Several studies have been undertaken to address such concerns for different response measures; however, conflicting interpretations and conclusions impact stakeholders’ confidence. For example, the assumptions that dispersant treated oils are more toxic than undispersed oil, dispersants are more toxic than oil, dispersants reduce the ability of microbes to degrade oil, and Arctic species are more sensitive to oil than non-arctic species are incorrect although all of these assumptions may be proposed as facts by multiple stakeholders.

The purpose of the following section is to describe those key areas that the workgroup recommended for further evaluation based on their critical review of available information. The subjects for further consideration are grouped into major subject headings for this executive summary. Details of the recommendations will be found in each of the sections of the reports.

0.1.1.1 Behavior and Fate of Oil in the Arctic

Arctic conditions influence the behavior and fate of untreated surface oil due to the low temperatures and the presence of different types of ice. Petroleum is generally immiscible in seawater and more so under colder temperatures. Surface oils with lower specific gravity accumulate on the sea surface and spread horizontally with the more volatile or soluble components quickly released into the air or into the water, respectively (NRC 1989; EPPR 2011). The surface oil also encounters disturbance by wind and wave action increasing the exposed surface area of the oil to the vertical transport processes of volatilization and solubilization. The wind and waves also adds water to oil creating oil/water emulsions that become more stable with time. The presence of broken ice with wind and waves enhances the disruption of the surface oil. These physical processes produce small globules of oil that may undergo further physical, chemical and biological weathering, creating aggregations of the heavier residual compounds that remain, eventually producing tar balls. In general, slicks formed in cold water are thicker with less exposed surface areas with reduced spatial coverage than the same oil released under temperate conditions. Laboratory trials have provided Information on key parameters that influence oil spreading under solid ice such as under-ice currents and ice roughness (Potter et al. 2012). In situations that the pour point of oil is above ambient temperature, the physical characteristics of the oil change dramatically with the wax components of the oil precipitating and forming a gel-like semi solid that is resistant to flow and spreading and which also restricts diffusion of volatiles through the slick, effectively reducing evaporation.

Untreated surface oils form slicks that are transported laterally by winds and currents and largely remain on the water surface and in the upper water column. The oil continues to spread on the surface of the ocean forming ribbons of slicks that rapidly spread to approximately 1 mm in thickness and are patchily distributed. Wind and waves break up and drive some of this surface oil into the water column in relatively large globules (>100 µm). The oil slick contains all of the chemical components of the spilled oil less those components that volatilize into the atmosphere or solubilize into the water column under the slick. The maximum concentration of oil is therefore contained in the immediate slick area, potentially exposing organisms that use the water near the air-water interface to the highest petroleum hydrocarbon concentrations that immediately follow a spill event.

The behavior and fate of released petroleum is an important consideration in understanding the potential effects of released petroleum and in evaluating the potential OSR options in the arctic. The spreading and weathering of petroleum in the arctic is complex, influenced by factors such as water temperature, local currents and wind conditions, the presence and absence of seasonal and multi-year ice, effects of pressure in deep water environments and seasonal changes in salinity during the Arctic spring. The presence of ice has been shown to slow the rate of spreading and weathering of surface oil, as well as affecting predictions of spill locations and trajectories. Migration of petroleum into brine channels or fissures in the ice can not only alter fate, but also the species that are potentially exposed and the exposure point concentration. At depth, petroleum is affected by increased pressures and decreased temperatures, resulting in phase shifts and changes in solubility as well as the dynamics of deep water currents and bathymetry. Key considerations included in the review were as follows:

  • Cold water temperatures and the presence of ice can dramatically affect the weathering and natural attenuation of oil in the Arctic, but the ice can also trap oil so that OSR options and time available to implement the necessary response can be extended.
  • Changes in behavior and fate of petroleum in deep sea environments associated with seeps and well blow-outs or leakage alter the bioavailability of oil components by allowing more volatile components to diffuse into the water where they may form clathrates altering the biodegradation potential and toxicity of those structures.
  • Adhesion of oil to particulate matter and how this may affect the potential for uptake into tissues;
  • The change in globule size and bioavailability of physically and chemically/OMA dispersed oil under Arctic conditions and,
  • The behavior of oil in the absence and presence of ice and how it influences the selection of OSR options.

The microbial degradation potential of oil in the Arctic has been demonstrated and is as effective as this process occurring in lower latitudes when natural communities of Arctic microbes respond to the presence of oil. Microbial response to oil in the Arctic and deep, cold and dark waters are emerging areas of research. While microbial degradation in temperate waters has long been recognized, recent laboratory and field studies have documented microbial degradation of petroleum and dispersed petroleum in these extreme environments. Current research is using analytical chemistry, respirometry, genomics, transcriptomics, and proteomics assays to not only show the presence of oil-degrading species, but measure the response and results of microbial activity upon being exposed to oil. Key considerations included in the review are as follows:

  • The presence and effectiveness of microbial communities to degrade oil in the Arctic (in open waters, in the presence of ice, along shorelines and in subtidal sediments);
  • The characterization of the microbial community responses and gene expression associated with exposure to oil and aerobic and anaerobic biodegradation and use of hydrocarbons and organosulfur compounds associated with the unresolved complex mixture; and,
  • The effects of biodegradation on the toxicity and availability of metabolites created during biological use of oil compounds and the changes that occur in toxicity and further biodegradation resulting from the more recalcitrant residual compounds of oil

There has been substantial research regarding the fate and effects of oil in the Arctic over the past 40 years; studies have been published in a number of different forums including peer-reviewed literature, technical reports, government studies and professional symposia. Additional data exists in a number of different languages, since research has been conducted throughout the North American, European, and Russian Arctic. Finally, important sources of data include emerging datasets from current research and older datasets that may not be as readily found in electronic search engines but nonetheless contain valuable information on environmental conditions and ecological resources. Consideration must also be given to the quality of data available for use. However, the most important aspect of making environmental consequence comparisons for OSR options is the appropriate framing of questions so that the consequences of response actions can be compared appropriately among all environmental compartments.

0.1.1.2 VECs and Ecotoxicity

The physiological, morphological, and behavioral adaptations of Arctic species may alter their sensitivity to petroleum and treated petroleum. To address this concern there have been a number of recent efforts to characterize the sensitivity of Arctic species to treated and untreated petroleum. Evaluations have included pelagic and benthic species, as well as those in close association with the ice. Endpoints that have been evaluated include survival, growth, reproduction, and behavioral effects, as well as molecular, cellular, physiological responses. Custom experimental facilities have been developed for working with chemically and mechanically dispersed oil and water soluble fractions (WSF) of differently weathered oil. Methodologies have been developed by project team members to capture and maintain Arctic species of interest for controlled laboratory studies. The VEC species that have been evaluated to this point have been found to have sensitivities similar to non-Arctic species for oil exposure. Both field and laboratory data have also been integrated with population models to provide estimates of population-level effects from oil exposures (e.g. SYMBIOSES and fishery population analysis). Key considerations included in the review are as follows:

  • Recent, historic and ongoing field and laboratory studies evaluating toxicity of petroleum and treated petroleum provided in Species Sensitivity Distributions to compare sensitivity of tested species.
  • Different exposure scenarios facilitate different types of evaluations and can dramatically affect comparability of data. Spiked exposures followed by reducing concentrations of oil represent the exposures of stationary species or those present in the water column when oil is undergoing the initial spreading and dilution following the spill event or after application of dispersants or OMA. Exposures to constant concentrations of oil represent zones of concentrated oil observed with neuston associated species and life stages and marine mammals and seabirds that move in and out of the air/water interface. These more constant higher exposure concentrations can also occur when oil is concentrated at edges such as shorelines, convergence zones, and water/ice edges.    
  • Endpoints found in the literature review range from body burden assessment to biomarker responses as well as mortality, growth, reproductive, developmental and behavioral responses. The diversity of potential end point assessments range from exposure assessment to end points that have a direct influence on estimating population level response to the oil components. For the purpose of this review, those responses that are better predictors of effects at the individual and population levels are the central focus. Mortality, growth and reproductive endpoints are those most closely associated with population level effects. Reviews on exposure markers will concentrate on demonstrating the relationship of the exposure marker to mortality, growth or reproductive endpoints.
  • All toxicity assessments are surrogate measures used to predict the potential effects of oil spills on living resources. As such, data obtained using sub-arctic and temperate species representing different groups of organisms or different environmental compartments may also be useful in augmenting datasets with Arctic species. Recent comparisons of the relative sensitivity of Arctic and non-arctic species suggest that non-arctic species have similar sensitivities warranting a broader evaluation of much larger data sets.
  • Additional testing of species that are long-lived, unique to selected habitat types and low in reproductive capacity that have not been evaluated in other regions have been identified during the reviews.

The toxicity of a mixture is characterized based on the analytical approach used to characterize the exposure. Variable conclusions regarding the relative toxicity of oil and water often can be tracked to the test waters being produced by different processes. The water accommodated fraction (WAF) is designed to only introduce the more soluble components into the water column while retaining the less soluble components on the surface of the water. The breaking wave water accommodated fraction (BWWAF) introduces additional physical disturbance, introducing more oil into the water as droplets with increased surface area exposure than occurs with the WAF allowing more of the soluble components to diffuse into the water from the oil droplets. The chemically enhanced water accommodated fraction (CEWAF) reduces the surface tension of the oil and produces much smaller droplets with much larger available surface area for diffusion of the soluble components of the oil. One of the objectives of the ecotoxicology section was to evaluate alternative methods of characterizing exposure.

0.1.2 Role of Ecosystem Consequence Analyses in NEBA Applications for the Arctic

A NEBA evaluation of OSR strategies for use in the Arctic must consider ecosystem-level consequences of the selected response. First, the effectiveness of the proposed solution(s) under the appropriate conditions to determine how much of the oil can be treated by the proposed action. Second, the consequences to various compartment VECs resulting from exposure to the untreated oil and the treated oil. Such comparisons must be made for resources in environmental compartments to determine the relative environmental benefits or risks of different response options. Third, the resilience of the populations of organisms that are being exposed as a result of no action or a response treatment needs to be addressed in order to determine the long term consequences of the decision.

Due to both logistical and environmental constraints, responses to oil spills rely on combinations of remote sensing and monitoring. The OSR options include 1) natural attenuation, 2) containment followed by recovery, 3) in-situ burning, and 4) dispersion using chemicals or oil mineral aggregates. With the exception of lower temperatures, oil spill response (OSR) options during the ice-free season are generally similar to other parts of the world. The presence of ice results in additional challenges as well as opportunities not encountered in regions without ice. The manner in which ice affects OSR effectiveness is determined in part by the ice characteristics which can differ regionally and seasonally. Recent field and laboratory studies that have evaluated the behavior, fate and effects of chemically dispersed oil, in-situ burning, OMA, and natural attenuation. Team members have evaluated spill response options in the presence and absence of ice and in both surface and deeper water environments. Both COOGER and SINTEF have led field releases of oil and studied field applications of OSRs on arctic shoreline/intertidal environments and in cold water and harsh environments including ice.

The objective of consequence analysis applied to oil spills is to provide spill responders a choice of response option(s) in terms of the lowest overall negative impact on the environment. It is likely that multiple response options will be selected and utilized for various stages of the response to reduce the exposure of VEC species. This process recognizes that once oil is spilled, some level of environmental impacts will occur, independent of the spill response options chosen. The goal of an effective response is to apply the combination of response techniques that will be effective in minimizing overall short and long term impacts. For the Arctic and other environments, this approach helps to focus technical discussions on the potential for short term and long term impacts on key ecosystem components and those resources of greatest cultural value to indigenous peoples. Assessments include comprehensive discussions of acute and chronic toxicity, food web bioaccumulation issues and reproductive and developmental impacts to exposed species. However, the discussions should focus on overall assessments at the population and community level of ecological organization, and ultimately promote a response strategy that allows for the fastest recovery of important ecosystem components. This approach has been used by governments and industry around the world to establish environmental protection priorities and spill response preparedness that will provide the greatest degree of overall environmental protection. IPIECA, International Maritime Organization (IMO), and OGP have long-supported this approach fostering development of a rational basis for setting oil spill response guidance and regulations (www.world-petroleum.org; www.imo.org; www.ogp.org).

These consequence evaluations have recently been incorporated into consensus building exercises that include all stakeholders. The process guides technical discussions and social prioritization among response planners, environmental agencies and local citizens as they compare ecological consequences of specific response options. The communication process is complex with many different opinions and levels of understanding of the effects of various response actions. The NEBA process has been particularly useful when considering use or non-use of dispersants, in-situ burning, containment and recovery of oil, as all of these present challenges with regard to potential environmental impacts. The process recognizes there will be damage during a spill but focuses on ecological consequences of different responses and compares “trade-offs” or cross-resource comparisons. Through a facilitated and structured analytical approach, participants find “common ground” for evaluating impacts and to develop defensible logic to support their conclusions. Discussion often can get stalled when there is a focus on localized and transient impacts of spills and response actions, without stepping back and trying to incorporate a longer term view of population and community recovery. Technical advisors and facilitators help guide the group to reaching their consensus among the diverse stakeholders by using a series of analytical tools specifically developed for use in a group environment. Knowledge regarding oil spill response capabilities and strategies gained by participants in the consensus-building process facilitates real-time decision-making in the event of actual spill incidences.

0.1.2.1 Arctic Population Resiliency and Potential for Recovery

Resiliency of VEC communities is a critical component to evaluating the consequences of OSR that needs to be further addressed in non-Arctic as well as Arctic environments. The direct, toxic effects of oil on individuals among the VECs are better understood than the resiliency of the populations and communities of these valuable ecosystem resources within various environmental compartments. Each environmental compartment is at a different level of risk resulting from response actions. Their resiliency is related to biological, physical and chemical attributes of the compartment and the species living in that compartment. The arctic environment is variable and harsh, featuring water temperatures that can range from -2 °C to greater than 5 °C, a light regime ranging from total darkness to total light, and regions that are covered in ice year-round and areas that cycle from being ice-covered to being ice-free. These widely varying conditions require behavioral, physiological, and morphological adaptations that may affect the sensitivity of some species to released petroleum as well as the dynamics of the population during recovery. The Arctic is considered to have relatively short food webs, with the higher trophic levels dominated by mammals and birds. Many of these species are dependent on rich populations of plankton that bloom heavily in spring in close association with ice break-up or upwelling zones. Key considerations included in the review were based on the following observations:

  • The projected damages to VEC populations are based on application of each OSR option. Use of OSRs changes the fate of oil; whatever OSR (including no response) is selected will alter the communities of organisms exposed to the oil at concentrations that can result in adverse effects. Damages include acute and chronic toxicity responses to oil. Additionally, the potential for recovery from oil contamination is also influenced by physical/chemical attributes that control the distribution and availability of the oil in each of the compartment as well as the availability of the oil to microbial degradation. Environmental compartment attributes that influence oil availability and ongoing biodegradation vary and need to be well quantified when establishing the long term consequences of OSR options.
  • The focus of the section on recovery potential is to document the available knowledge and identify uncertainties in our understanding for VECs in different environmental compartments. Species found in arctic waters have a number of unique physiological and morphological adaptations to allow them to tolerate the cold water temperatures and the extensive periods of ice cover and absence of sunlight and associated food resources. The data reviews in the ecotoxicology section shows that these factors may influence the time period for demonstration of effects after exposure. However, the responses of Arctic VECs are similar to non-Arctic test species.
  • The reproductive potential of Arctic VEC species is the key factor in assessing the ability of their populations to recover from stress/damage. In terms of recovery times, this resilience will vary from very short periods of hours in the case of microbial populations to many decades in the case of marine mammals and seabirds. This resilience and recovery potential for VECs in the Arctic is similar to what has been characterized in other regions of the world.
  • Communities and food webs change dramatically during periods of open water or iced over water resulting in seasonal modifications of the available environmental compartments. During the winter the annual ice environmental compartments increase while during the spring and summer the melting ice adds more open water pelagic compartments. Seasonal OSR evaluations need to consider the change in the availability of these compartments, as well as the associated seasonal changes in effectiveness of the response options.
  • There are specialized and unique species that live under and within ice, including larval forms of important water-column species. While the communities under the multi-year ice are becoming better known the ecological importance of the annual undersea ice is less understood.
  • There are also deep-water Arctic communities that feature unique species and communities, including deep water corals and sponges. The extent of these communities under multi-year ice is not well characterized but the presence of ridge topography determined by geophysical means indicate there is a potential for these species to be more broadly distributed than is known at the present time.
  • There are also regional and seasonal differences across the arctic which is principally associated with some of the higher trophic level species. The rationale for selection of VEC species is to emphasize those which are pan-arctic species that support these higher trophic levels. Unique species are also examined to determine associations with specific environmental compartments.
  • In general, there is less knowledge of ecological processes occurring during the Arctic winter season.

Current oil spill contingency and response models representing transport, fate, and limited effect-based components have been used to support ecosystem evaluations. A key area of additional research recommended by the work groups is to augment Arctic NEBA assessments with an environmental compartment approach in order to evaluate the short and long term consequences of oil potentially impacting the resources in those compartments (Figure ES-2). Resiliency of the inhabitants using different environmental compartments will govern the recovery from oiling and is the key focus for the recommended new work.

0.2 Priority Recommendations to Enhance NEBA Applications in the Arctic

The review of Arctic literature found many high quality assessments at the laboratory, mesocosm, and field scales on the efficacy of the treatment options (natural recovery; containment and product recovery; use of chemical dispersants or OMA; and in-situ burning). These studies provide information for efficacy estimates and for application to transport models providing good estimates of the movement of residuals from one EC to another within the Arctic. There are also high quality toxicological effect and biodegradation studies conducted under Arctic conditions that indicate the relative acute sensitivity of Arctic species are equivalent to sub-Arctic, temperate and tropical species and that biodegradation of oil by indigenous Arctic microbes is efficient and more rapid than anticipated based on the cold inclement conditions in the Arctic. The priority recommendations of each workgroup are included at the end of each section. A list of the high priority work elements within each section is appended to the summary of recommendations. Several of the recommendations apply to local operations and should be incorporated in baseline monitoring programs, Environmental Impact Assessments and operational oil spill response plans. These recommendations have not been prioritized here, but should be considered by the individual operators.

Four overarching areas of research were identified and were considered by the scientific panel of experts to be of the highest priority to advancing NEBA applications in the Arctic:

  1. Increase availability of the vast amount of data on the impacts of oil spill response techniques and on the resilience of VECs reviewed in the current study by developing and populating matrices, or Arctic Response Consequence Analysis Tables (ARCATs), in support of Arctic NEBA processes.
  2. Determine influence of oil on unique Arctic communities within EC interface habitats as well as corresponding response consequences (resiliency, sensitivity, and exposure potential)
  3. Further investigate in-situ biodegradation of oil and oil residues
  4. Further explore consequences of acute and chronic toxicity from oil exposure through use of population modeling

0.2.1 Development of ARCAT Matrices

All response actions, including natural recovery, result in adverse effects on some portion of the environment and to some species. The consequences of those actions need to be compared holistically among all ECs and VECs. Direct ecotoxicological responses to the oil spill residuals is an important aspect of these considerations but the long term recovery potential of the ECs and VECs is even more important. The relationships between EC resilience and the sensitivity and resilience of VECs within those ECs is an important area that needs to be addressed in OSR consequence analysis. A way to consistently summarize and present available data on the impacts of oil spill response techniques on selected VECs within ECs would be through the development of information matrices that collate the relevant available data to support Arctic NEBA processes. These ARCATs (Arctic response consequence analysis tables) summarize the relevant information from literature to focus on the initial and long term consequences of oil spill response actions so that impacts to ECs and VECs that are less resilient and which take the longest time to recover are avoided. To accomplish that goal a number of tasks are suggested.

  1. Develop a format for ARCAT matrices that will address the potential effects of oil spill response treatment residuals on ECs and VECs
  2. Describe the OSR actions, ECs and VECs that will be used to develop the ARCAT matrices.
  3. Map the EC locations that are relatively permanent and indicate the areas for seasonal positions of other ECs.
  4. Document the use of ECs by VECs on a seasonal basis with the assumption that EC usage by VECs occurs during those seasons whether they are observed or not during a spill event.
  5. Provide biological attributes that influence exposure potential (e.g. skim feeding, haul-out locations), reproductive capacity (e.g. time to adult, progeny production rates that may be influenced by age, daily natural mortality rates of populations), and immigration potential that augment recovery potential.
  6. Document physiological and behavioral characteristics of VECs that increase or decrease exposure potential especially for mobile species (e.g. large fish marine mammals and seabirds).
  7. Document assumptions of the population sizes and age classes of VEC components.
  8. Document those characteristics that will be used to compare the resiliency of EC and VEC response to oil spill residuals.
  9. Summarize the responses of species to OSR that are less studied (e.g. ISB, OMA, and chemical herding).
  10. Demonstrate food web complexity in the Arctic using stable isotope ratios to demonstrate structure or unstructured aspects of key food webs

Figure E-2. Arctic ecosystem (surface microlayer, ice, nearshore continental shelf, pelagic and deep water communities) [Most literature is representative of the pelagic community, outlined in red]
Figure E-2. Arctic ecosystem (surface microlayer, ice, nearshore continental shelf, pelagic and deep water communities) [Most literature is representative of the pelagic community, outlined in red]

0.2.2 Influence of Oil on Unique Arctic Communities

An understanding of the transport, fate, and exposure potential of oil spill treatment residuals is an ongoing need for convergence zone communities. Convergence zones include shorelines, air/water, sediment/water, ice/water, and water/water interfaces. The VECs that occupy these convergence zones or that move in and out of these zones undergo exposures that are not similar to the better understood open pelagic water exposures. The areas of additional work suggested by the work group include a better understanding of toxicology and biodegradation for two convergence zones air/water and ice/water:

  1. Describe the ecological relevance of interface zones, specifically the surface of the ocean and the ice water interface edges.
  2. Describe the seasonal use of these two interface zones by VECs.
  3. Develop testing protocol for exposure evaluations at these two interfaces. Consider Early Life History of fish, Arctic algal/kelp, sediment, deep water hard substrate, and fouling exposures, bioaccumulation and lipid characterization.
  4. Develop or provide analytical chemistry methods for OSR residuals and degradation products in water, sediment and tissues.
  5. Describe the differences in exposure of VECs to OSR residuals and the modes of toxic action encountered in the surface microlayer (neuston) and water/ice interface environments.
  6. Initial responses by pelagic organisms to oil spills are relatively well understood and predictable. However, the long term consequences of an oil spill and subsequent effects on the recovery potential of sites, species or communities is less well developed. A better understanding of the factors that control the physical, chemical and biological attributes that control VEC and EC resiliency, which ultimately controls the amount of time for recovery needs to be developed.
    1. Define and establish measurements of resilience and develop methods to combine resiliency metrics into compartment specific short and long term recovery assessments.
    2. Compare EC population sizes for VEC in EC so that the relative importance of the air/water and ice/water interface environments can be assessed.
    3. Toxicity responses to exposure from accommodated fractions of fresh oil and chemically dispersed oil have predictable results based on the measured concentrations of oil in the liquid phase of these preparations. The responses are relatively consistent from experiments performed on tropical, temperate and Arctic species. Less is known about the toxicity of other OSR residuals (ISB, OMA, chemical herders) and weathered OSR residuals.

0.2.3 Biodegradation in Unique Communities

Biodegradation of fresh and chemically dispersed oil using indigenous organisms from pelagic Arctic environments has been demonstrated to occur at rates similar to those observed in deep and surface waters throughout the marine environment. Biodegradation of oil and OSR residuals when they are concentrated at interfaces (e.g. air/water and ice/water) is less well understood but assumed to be slower based on the reduction in surface area of the oil when it is re-concentrated in these convergence zones. The recommended areas of research for these areas are:

  1. Establish a method to evaluate biodegradation occurring at interface environments (air/water and ice/water).
  2. Compare biodegradation success for OSR residuals using this method.
  3. Evaluate –omics procedures for application to OSR residual assessments.
  4. Consider the implications of the storage of OSR residuals within convergence zones. This should not only include the air/water and ice/water but also shoreline stranding and the influence of shoreline sediment type on short and long term storage of OSR residuals.
  5. Produce a GIS or EC database that addresses recovery potential of those ECs impacted by OSR residuals.
  6. Evaluate food web structure within areas of oil/gas seeps and demonstrate the presence or absence of food webs that are oil/gas based and whether those food webs are supplementary to detrital based food webs in the same area.
  7. Demonstrate whether these oil/gas based food webs provide communities that are pre-adapted and more resilient to exposure of OSR residuals.

0.2.4 Modeling of Acute and Chronic Population Effects of Exposure to OSRs

Key attributes to determining the importance of acute and chronic responses of organisms to OSR residuals are measures of the natural resiliency of individuals, populations of species, and structure of food webs. Modeling of the transport, fate and effects of oil spills is relatively well known and predictable. Toxicity at interface environments is much higher due to the concentration of contaminants at these locations but the population influence of that concentrated contaminant exposure is less well understood. The areas of research suggested to provide improvement in the understanding of population impacts to VEC from exposure to OSR, especially at these interface habitats is provide in the following:

  1. Establish the natural structural, functional and population dynamics measures for VECs within key ECs, including comparing the interface environments of air/water and ice/water interfaces.
    1. Determine whether the ‘sustainable’ reduction of 20% used in resiliency assessments for fishery harvests is applicable and whether that level of effect can be determined within the natural population fluctuations of VECs.
    2. Evaluate whether a 20% reduction of a VEC population would influence success of food web structure.
  2. Obtain population dynamics, age-related characteristics of populations of key components leading to VECs (daily mortality rates, age/size population structure, variation in population size/standing crop, seasonal sequestration of significant portions of populations, emigration/immigration potential of individuals, etc.).
    1. Refine those characteristics for populations of these species living in different ECs.
  3. Establish food web connections based on literature and alternative measures (e.g., stable isotope ratios).
  4. Provide a predictive modeling diagram that incorporates knowledge of spilled materials, transport and fate of OSR residuals, influence of biological transport of contaminants within the water column, toxicity assessment and consequence analysis of the application of alternative OSR actions.
  5. Develop factors to address transport and fate of OSR residuals along ice edges, under ice and near sea-surface.
  6. Modify models based on new input parameters obtained from these research areas.

Each review team arrived at similar overall conclusions. They observed that there were high quality assessments on the efficacy of various response options. Those assessments were conducted under different physical conditions representing the Arctic at laboratory, mesocosm, and field scales. It was also generally recognized that strong transport models exist and are applicable to some Arctic conditions, including broken ice but some necessary adjustments to include information on currents, and transport under multi-year and annual ice conditions may be required. However, past estimations of the effects of various response actions have been generally isolated to a single environmental compartment or species based on acute toxicity responses using pelagic species or age classes. These response data have been used to predict ecosystem level effects in different habitats by modeling the concentrations of oil as it moves through the environment and interacts with new species or age classes. Although the research on the acute effects is generally well done, the long term consequences to different environmental compartments and species has not been well developed. This includes chronic toxicity responses but more importantly the overarching physical and chemical resiliency of environmental compartments and the biological recovery potential for species or age classes living in those compartments that are exposed to weathered and treated or untreated oil. Resiliency of environmental compartments and population to oil exposure is less well known not only in the Arctic but also non-Arctic environments. These longer term factors in recovery are key components needed to provide a better balance to evaluations of the consequences of oil spills and use of alternative treatment options.

0.3 Further Information

References

AMAP

2010

Assessment 2007: Oil and Gas Activities in the Arctic - Effects and Potential Effects - Volume 2

Arctic Monitoring and Assessment Programme; Oslo, Norway

EPPR

2011

Behavior of Oil and Other Hazardous and Noxious Substances Spilled in Arctic Waters.  Arctic Council, Emergency Prevention, Preparedness and Response working group

http://eppr.arctic-council.org/). 121 p.

National Research Council (NRC)

1989

Using Oil Spill Dispersants on the Sea

National Academy Press

NewFields

2012

Joint Industry Program to Evaluate Biodegradation and the Effects of Dispersed Oil on Cold Water Environments of the Beaufort and Chukchi Seas.  Phase I and II Report

NewFields LLC, Port Gamble Washington.

USGS

2011

An Evaluation of the Science Needs to Inform Decisions on Outer Continental Shelf Energy Development in the Chukchi and Beaufort Seas, Alaska

USGS Circular 1370.

Potter S, I Buist, K Trudel, D Dickins, E Owens

2012

Spill Response in the Arctic Offshore

Report prepared for American Petroleum Institute and Joint Industry Programme on Oil Spill Recovery in Ice

Sørstrøm SE, Brandvik PJ, Buist I, Daling P, Dickins D, Faksness L-G, Potter S, Fritt Rasmussen J, Singsaas I

2010

Joint industry program on oil spill contingency for Arctic and ice-covered waters

Summary report; SINTEF report A14181. SINTEF. Trondheim, Norway. Www.sintef.no/Projectweb/JIP-Oil-In-Ice/Publications/.

SL Ross

2010b

Literature review of chemical oil spill dispersants and herders in fresh and brackish waters

SL Ross Environmental Research; Ottawa, ON.  66 pp.

1.0 THE PHYSICAL ENVIRONMENT

Executive Summary

Photo 1-1. Polynya (MODIS)
Photo 1-1. Polynya (MODIS)

The Arctic maritime region is comprised of the Arctic Ocean, six marginal seas, and six deep water basins. The Arctic Ocean is a beta ocean, i.e. it is dominated by a strong halocline which acts to retard mixing and vertical flux of biotic and abiotic materials. The main influx of water is derived from the Atlantic Ocean; waters entering via the Chukchi Sea are minor in comparison due to the shallow sill depths in the areas of the Bering Strait and Chukchi Sea. Most oil and gas production, major shipping routes, and population centers are located near the continental shelf regions, often coincident with the specialized environments of polynyas (free waters within ice) and freshwater discharge from several major river systems. Continental runoff is a major source of freshwater, terrigenous materials and nutrient loads to the Arctic seas. The Arctic Ocean surface layer (ASL) is generally insulated from the warmer and higher salinity bottom water (Atlantic water layer; AWL) by a well-developed halocline (cold halocline layer, CHL). The halocline is a strong barrier to upward mixing by turbulence, and consequently most of the ocean heat flux in the central basins of the Arctic is generated by solar heating through open leads and thin ice during the summer months. Sea ice nomenclature describes age (thickness as it ages), forms (e.g. pancake or brash ice, floes, icebergs), and concentration (i.e. relative ice/sea coverage) [Tables 1-1 and 1-2]. Average multiyear sea ice distribution and thickness changes over time; estimates of ice cover for February/March 2004 - 2008 and February/March 2012 are compared in Figure 1-1. Ice provides habitat for ice algal communities and enhances food resources, whereas solid ice fields may sufficiently retard solar irradiance to reduce primary production. The processes of ice production, salt flux, and heat transfer from leads and polynyas are important contributors to biological productivity and the larger scale climate events in the Arctic, in addition to the overarching seasonal changes in degree of solar radiation. Biological productivity hinges upon solar irradiance and nutrient availability; the availability of these two resources are regulated by the albedo of ice and snow surfaces and the salinity stratification of the Arctic waters.

Table 1-1. Sea ice Formation (Environment Canada 2005)

DescriptionThickness

New Ice

Frazil Ice

Grease Ice

Slush

< 10 cm

Fine spicules or plates of ice

Ice crystals have coagulated

Snow, saturated and mixed with water

Nilas, Ice rind

< 10 cm

Young Ice

Grey Ice

Grey-white ice

10 - 30 cm

10 - 15 cm

15 - 30 cm

First-year ice

Thin first-year ice

Thick first-year ice

≥ 30 cm

30 - 70 cm

> 120 cm

Old Ice

Second-year Ice

Multi-year Ice

 

Variable to 5 m

The changing physical environment in the Arctic consists of extremes in temperature, amount of light, and weather conditions that act upon the pelagic open waters and convergence zones (shorelines, sea surface layers, ice water interfaces, sediment water interfaces, and water convergences). Convergence zones form an environmental compartment (EC) of the Arctic where valuable ecosystem components (VECs) congregate (see Section 2) providing opportunities for alternate exposure pathways and responses resulting from different modes of toxic action (See Section 6). The physical environment also influences the fate of oil (see Sections 3 and 5) and provides opportunities for alternative oil spill response (OSR) actions (see Section 4). The physical environment needs to be well understood because it sets the stage for minimizing short and long term consequences of applying alternative OSR options (see Sections 7-9). For these reasons determining the location of more stable convergence zones (shorelines and sediment water interface environments) and understanding the variations in the location of more mobile convergence zones (ice water interfaces and current convergence zones) are key components in the selection of OSR options to reduce the consequences of an oil spill.

Some description
Figure 1-1. Changing distribution of ice cover (NSID) (Colors indicate ice thickness in meters; blue = 1m, red=5 m)

Table 1-2. Relative ice concentrations [Adapted from Environment Canada 2005] 

Some description

<1/10 Open water

Some description

8/10 Close pack/drift ice

Some description

3/10 Very open drift ice

Some description

9+/10 Very close pack

Some description

6/10 Open drift ice

Some description

Compact/Consolidated Ice

1.1 Introduction

In contrast to Antarctica, the physical environment in the Arctic is a consequence of recent glaciation and the relatively short time span for ecosystem and faunal diversification.  Polar ecosystems are characterized by extreme environmental conditions induced by cold temperatures, extensive snow and ice cover with abbreviated periods of solar radiation and primary productivity.  In general the productivity in Arctic freshwater and marine systems is concentrated over short periods of time centered around ice breakup when nutrients become available and light becomes more prevalent and water temperatures warm as the ice cover is reduced.  These environmental constraints concentrate recolonization during a few months of the year resulting in low species diversity due to extremely fast population growth of key zooplankton species responding to the release of nutrients which then contribute food to slower growing and longer-lived species. The objective of this chapter is to present an overview of the physical environment of the Arctic and indicate important characteristics that determine presence of Arctic ecosystem components.

1.1.1 The Arctic Ocean, Marginal Seas, and Basins

The Arctic Ocean and associated waters comprise one of the most unique marine ecosystems in the world.  Two sources of productivity are instrumental in actively replenishing nutrients over short time periods:  ice algae, and riverine input.  It has been estimated that ice algae contributes 10-70% of annual productivity (AMAP 1998).  River discharges to the Arctic shelf regions augment nutrients, organic materials, and sedimentation on a seasonal basis.  The marginal seas are either influenced by the Atlantic or Pacific Oceans (Nordic, Barents, Northern Labrador Sea and Bering, Chukchi Seas, respectively), or are relatively isolated and border the Asian or North American continents (Kara, Laprev, East Siberian, and Beaufort, respectively).  The Chukchi, Bering and Barents Seas are among the most seasonally productive ecosystems.  The seas bordering the continental landmasses are influenced by freshwater runoff from the river systems and prior to the onset of recent increased warming trends had landfast ice associated with the shorelines for most of the year.  Freshwater discharge from rivers leads to earlier open water in the nearshore zone.  Maximum productivity is limited to open coastal waters during spring/summer months or to polynyas between landfast ice and the polar pack ice (occurring near the continental shelf edges).  Organic material not cycled through organisms or advected to the central Arctic basin is incorporated into sediments, producing localized areas of high organic enrichment near the mouths of major rivers and to locally deposited sections of tundra breaking off from shore locations.  Nearshore river systems are characterized by estuarine conditions, i.e. higher water temperatures, lower salinity, increased turbidity, as well as increased productivity due to the organic enrichment.  In most other areas the benthic standing crop decreases with increasing depth; ridges such as the Lomonosov Ridge have higher benthic standing crops than adjacent basins (Kröncke 1994; Kröncke et al. 1994).  The age of the organisms creating these standing crops on ridge environments may be quite old and reflect development over many decades but this has been a difficult environment to investigate so little is well known.

The two main deep basins, the Eurasian and the Canadian are separated by the transpolar Lomonosov Ridge (Figures 1-2 and 1-3).  The Canadian Basin which is < 3500 m in depth is transected by the Alpha Cordillera ridge into the Makarov and Canada Basins.  The Eurasian Basin is deeper, reaching depths of 4000m, and is divided by the Nansen Cordillera into Amundsen and Nansen basins.  While the continental shelf generally extends 50 to 100 km offshore, the shelf is broad north of Siberia, extending up to 900 km offshore. 

Figure 1-2. Basins and ridges of the Arctic Basin (Mike Norton)
Figure 1-2.   Basins and ridges of the Arctic Basin  (Mike Norton)

Figure 1-3. Main water bodies of the Arctic (Source:  AMAP 1998)
Figure 1-3.  Main water bodies of the Arctic (Source: AMAP 1998)

1.2 Knowledge Status

1.2.1 The Circumpolar Margins

Most oil and gas development activities, shipping routes, as well as major fishing grounds occur along the margins of the Arctic Ocean at the interface of land and sea edges.  Figure 1-4 highlights the main shipping routes transiting Arctic waters and areas of active oil and gas exploration.  Increased periods of ice-free conditions along shipping routes will result in increased vessel and fishing activities.   It should be noted that increased commercial fishery pressure itself may lead to changes in fishery stocks and biodiversity in the Arctic.

Continental runoff is a major source of freshwater, terrigenous materials and nutrient loads to the Arctic Ocean.  Information gathered from the Regional Arctic Hydrographic Network data set (R-ArcticNET) indicates that the overall annual discharge is ~3,300 km3 y-1.  This buoyant freshwater contributes to a low saline layer (upper 200 m) of the Arctic Ocean which is isolated from the warmer, saltier Atlantic layer by a strong halocline (Fichot et al. 2013).  Rivers account for 2% of the influx of water to the Arctic region resulting in highly productive areas, particularly during the peak flow occurring from April to July (~60%).  The rivers in Siberia (Ob, Yenisey, Lena, and Kolyma) and in North America (Yukon, and Mackenzie) provide the majority of continental fresh water to the Arctic Ocean.  The Arctic Great Rivers Observatory (Arctic-GRO) monitors the discharge of these six river systems.

River systems generally increase productivity due to the availability of increased organic nutrients in the areas of riverine discharges.  Figure 1-5 shows the locations of the major rivers in the Arctic.   The regions along the continental shelf may be the most likely to encounter oil spill incidents, either due to oil and gas production activities or vessel mishaps. In determining environmental impacts of released oil, areas of special importance are the interfaces where oil components may re-concentrate (e.g. current convergence zones, pycnoclines, upwelling or downwelling of water masses, shoreline stranding or concentration at air-water, ice-water, or sediment-water interfaces).

Some description
Figure 1-4. Circumpolar regions of activities (AMAP)

Figure 1-5. Major river systems in the Arctic
Figure 1-5. Major river systems in the Arctic

1.2.2 Arctic Hydrography

Ocean temperatures vary widely depending on latitude and proximity to warm Atlantic or Pacific Ocean waters.  For the Arctic Ocean, temperature variation between winter and summer is small (remains close to freezing year round) and salinities vary between 30 and 33.  In the coastal shelf areas, surface water temperatures range from -1 °C to 4-5 °C, winter to summer, respectively while salinities may be <30, especially in areas receiving freshwater from rivers and ice melt.  In areas where oceanic mixing occurs, the temperature remains higher than 0 °C throughout the year.  In the Kara Sea and Siberian shelf, salinity is <20 throughout the year, and may drop to 10 during the summer (Figure 1-5; AMAP 1998).  Arctic seas are primarily ‘beta oceans’, i.e. a salinity profile is the most important permanent stratification feature (Carmack and Wassmann 2006).

Figure 1-6. Temperature, salinity, and density profiles (AMAP 1998)
Figure 1-6. Temperature, salinity, and density profiles (AMAP 1998)

Warm ocean currents flow northward from the Atlantic and Pacific Oceans and cold Arctic countercurrents flow southward (Figure 1-6).  Arctic Ocean water mainly flows from the Atlantic Ocean (79%), while inflow from the Bering Strait accounts for only 19%.  Concern with thinning of Arctic ice cover over recent decades has generated numerous investigations of the ocean-ice heat flux and changes in the Arctic Ocean surface layer (ASL).  The ASL is generally insulated from the warmer and higher salinity bottom water (Atlantic water layer; AWL) by a well-developed halocline (cold halocline layer, CHL).  The halocline establishes as a strong barrier to upward mixing by turbulence, and consequently most of the ocean heat flux in the central basins of the Arctic is generated by solar heating through open leads and thin ice during the summer months.

Figure 1-7. Arctic currents (AMAP 1998)
Figure 1-7. Arctic currents (AMAP 1998)

Researchers have found that although changes in surface velocity and surface stress in the open ocean reflects large scale atmospheric pressure fields (the Arctic oscillation, AO), the properties of sea-ice concentrate energy into relatively narrow zones of intense shear which can raise the pycnocline.  Such events may greatly enhance ocean-to-ice heat transfer (McPhee et al. 2005).  Examination of hydrographic records indicated that the CHL dissipated during a low pressure system from 1988 to 1997 in the Nansen and Amundsen Basins, resulting in an increased rate of melting sea-ice (15-25 cm/y).  The lateral Ekman currents moved in a counter-clockwise direction, resulting in upwelling of warmer, saltier water from the AWL and raising a weakened halocline layer about 50 m.  Average seasonal changes in sea-ice conditions (concentration and movement), sea level pressure, Ekman transport vectors, and upwelling patterns are presented in Figure 1-7 (Yang 2008).  

Figure 1-8. Seasonal sea-ice and ocean movement averaged over a 28 year period (Yang 2008)
Figure 1-8. Seasonal sea-ice and ocean movement averaged over a 28 year period (Yang 2008)

Closer to shore, the seasonal cycle of Ekman transport is highly influenced by high sea level pressure in fall and winter.  For example, strong offshore transport along the Beaufort, Chukchi and East Siberian coasts intensifies coastal upwelling and downwelling in the interior Canadian Basin.  The offshore transport also pushes low-salinity warm shelf water toward the deeper basins, reducing salinity and increasing temperature.  

The continental shelf zones comprise about 50% of the Arctic Ocean surface and there are marked regional differences.  Narrow shelves are characteristic of the North American continent, and wide shelves with very steep slopes are typical of Eurasian continents (Table 1-1). Carmack and Wassmann (2006) group shelf areas according to function:  1) Inflow shelves (Bering Strait/Chukchi Sea, Barents Sea); 2) 2) Interior shelves (Beaufort and Kara/Laptev/East Siberian Seas); and 3) Outflow shelves (Canadian Archipelago and East-Greenland shelf).   The pan-Arctic shelves are estuarine in nature in that waters originating from the Atlantic and/or the Pacific Oceans mix with inflowing river waters.  Although there are regional differences in source water characteristics, all shelf areas are generally associated with the nutrient enrichment, productivity, and biodiversity.  Figure 1-8 is an excellent illustration of the dynamic physical, chemical, and biological interactions occurring in the Arctic shelf regions in the vicinity of receiving waters of major river systems.  In general, the nearshore ecosystems are characterized by relatively few trophic links and low biodiversity, and would be more sensitive to increased climatic warming than temperate regions (Carmack and Wassmann 2006).

Arctic SeaFunctionMean depth (m)Area  (103 km2)% Total shelf area

Barents Sea

Inflow

200

1597

27

Kara Sea

Interior

56

926

15

Laptev Sea

Interior

131

498

8

East Siberian Sea

Interior

48

987

16

Chukchi Sea

Inflow

58

620

10

Beaufort Sea

Interior

80

178

3

Canadian Arctic Archipelago (CAA)

Outflow

124

1032

18

Northern CAA

Outflow

310

210

3

Total

 

140

6048

100%

Table 1-3. Characteristics of Arctic shelves (after Carmack and Wassmann 2006)

Figure 1-9. Dynamic shelf processes augmented by nutrient loading from riverine discharge (Source: AMAP 1998)
Figure 1-9. Dynamic shelf processes augmented by nutrient loading from riverine discharge (Source: AMAP 1998)

1.2.3 Ice And Ice-Edges

Sea ice forms when the temperature of the ocean falls below the freezing point, effectively converting the seawater to ice (-1.8 °C at 33 ‰ salinity).  Ice is characterized by where it is located:  landfast ice lines the shoreline and first year ice is intermediate between the multi-year ice and open water.  The annual changes in pack ice are determined by temperature, winds, and ocean currents (such as the Alaska, Labrador, East Greenland currents and the warm West Spitsbergen and North Cape currents).  Multi-year ice averages 2.5 to 4 m in thickness and may be intersected by leads such as shown in the photo above (~1% open in winter to 10-20% open in summer; Gow and Tucker 1990).  The two main ice circulation systems in the Arctic are the clockwise Beaufort Gyre in the Amerasian Arctic and the Transpolar Drift in the Eurasian Arctic (migrating east to west, exiting via the Fram Strait).  The North Cap current keeps the southern Barents Sea ice-free during the winter.

Photo 1-2. Ice lead (Lionel Camus)
Photo 1-2. Ice lead (Lionel Camus)

Photo 1-3.  Polynya
Photo 1-3.  Polynya

Ice leads and open-water lenses (polynyas) form in nearshore or ocean areas and are characterized by cold, highly saline water (Photos 1-2 and 1-3).  The processes of ice production, salt flux, and heat transfer from leads and polynyas are important contributors to the larger scale climate events in the Arctic.  Polynyas occur for the most part during winter and harsh weather conditions and may be formed by two processes: 1) Mechanically forced (wind-driven; strong winds force ice cover away from the coast; 2) Convectively forced (ice subsidence; subsurface heat is transferred to the surface water by the upwelling of warmer, deeper water); refer to Figure 1-10.   Improved satellite technologies (e.g. synthetic aperture radar, SAR) and mathematical algorithms (e.g. polynya signature simulation method, PSSM) have made it possible to define the size and shape of polynyas as well as differentiate open water, new ice, and young ice (Dokken et al. 2002).  The heat exchange from the ocean to the atmosphere can be orders of magnitude larger in polynyas compared to the surrounding ice pack, and polynyas can contribute ~50% of the seasonal mean of annual ice production in some areas (Winsor and Bork 2000).  Winsor and Bjork (2000) estimated the mean ice production and corresponding salt flux from Arctic polynyas and found that the salt flux represents about 30% of that necessary to maintain the CHL.    The marginal seas contributing the most to the CHL are the Barents, Kara, Chukchi, and Bering; the Chukchi Sea is the only sea contributing actively to deep water formation.  The North Water Polynya is located between Greenland and Canada in northern Baffin Bay.  It is 85,000 sq. km in area and creates a warm microclimate that is one of the most biologically productive marine areas in the Arctic Ocean.  Polynyas are of critical importance to significant populations of marine birds, mammals, and fish.

Some description
Figure 1-10. Polynyas and Sensible-Heat Exchange [Based on image modified from Ocean Circulation, 2nd Edition by Open University, Butterworth-Heinemann Publishers, page 219; Source: National Sea Ice Data Center (NSIDC)]

The term seasonal ice zone (SIZ) describes the presence of transitional ice (i.e. freezes and melts annually) whereas marginal ice zone (MIZ) is used to describe the areas where open water meets ice cover during any season.  About half of the Arctic sea-ice freezes and melts annually.  During summer months, narrow MIZ and wide SIZ bands circumscribe the polar region.  It is along this band of expanding and shrinking open water linkages that most of the Arctic Ocean productivity takes place and where climatic changes would be most evident (Carmack and Wassmann 2006).  In recent years, the ice-free season in the Arctic increased at a steady rate of 1.1 days/year.  A record low in sea ice was recorded in 2007, with 168 ice-free days (Rodrigues 2009).  Increasing seasonal open water periods is coincident with anticipated increases in shipping traffic via the polar routes (Northwest Passage and Northern Sea Route).

1.2.4 Seasonality: Productivity and the Carbon Cycle in the Arctic

Biological productivity hinges upon solar irradiance and nutrient availability.  These two resources are regulated by salinity stratification of the Arctic waters.  In general stratification is expressed as buoyancy frequency relating vertical temperature and salinity gradients with consideration of gravity, thermal expansion, and haline contraction coefficients (Carmack and Wassmann 2006).

The annual productivity/carbon cycle in the Arctic is illustrated in Figure 1-10.  The sun appears over the horizon in April, supplying light to support the growth of ice algae and phytoplankton. Production peaks in June when the sun is at its highest; zooplankton thrive on this superabundance of food. The production gradually declines during the season as the phytoplankton use up the nutrients in the water and as the sun sinks below the horizon, the plankton hibernate until the next growing season (http://www.arcticsystem.no). For areas in the Canadian high Arctic, Welch et al. (1992) estimated the relative importance of contributors to primary productivity as: phytoplankton (90%) > ice algae (10%) > benthic algae (1%).  This relationship likely varies in different locations.

Figure 1-11. Productivity/carbon cycle in the Arctic [Original illustration: Alexander Keck & Paul Wassmann (1993), modified by Frøydis Strand, NFH, University of Tromso; with adaptation from Forest et al. 2013)]
Figure 1-11. Productivity/carbon cycle in the Arctic [Original illustration: Alexander Keck & Paul Wassmann (1993), modified by Frøydis Strand, NFH, University of Tromso; with adaptation from Forest et al. 2013)]

Physical processes inherent in ice field characteristics, water column stratification, vertical mixing, light and nutrients in combination with phytoplankton determine the balance of productivity and biomass build-up in the euphotic zone and its transfer to the aphotic zone.  The fate of suspended and sinking biogenic matter is based on the magnitude of export production and biological activities such as grazing in the upper water column.  The transfer of carbon from the surface layer to the deep ocean is accomplished via the passive sinking (exported) or the active transport of organic material (harvested, fecal production).   The greatest congregation of zooplankton often occurs just below the euphotic zone (Olli et al. 2006); excess detrital matter ultimately sinks into the aphotic zone.  In some regions such as the Barents Sea, the pelagic food web is positively correlated with benthic standing stocks (Carmack and Wassmann 2008).   There is considerable variability in sediment community structures and biomass, depending on carbon development in the upper layers.   Ice cover and other physical factors have been observed to limit primary production and benthic biomass (Grebmeier et al. 1995). 

The highly seasonal attributes of the Arctic influence species abundance and distribution patterns.  The lack of sunlight during the Arctic winter in combination with ice-covered waters limits primary and secondary production when solar irradiance is unavailable.  The periods of low production are punctuated by extremely high production during breakup and seasonal open water.  The combination of open water, long days, and high nutrient concentrations during break up and early summer creates intense periods of primary production, grazing by primary consumers, and predation by the higher trophic levels.  Whereas the aquatic communities are limited during the winter, the diversity and abundance of species increases dramatically in the summer. Key Arctic species have adapted their life cycles to take advantage of this period of high production, and larval or juvenile life stages are adapted to capitalize on the short period of enriched food resources.

The initiation and duration of summer production follows a latitudinal gradient from the southern Barents Sea in March – April to August – September in the Fram Strait and Arctic Ocean (Figure 2-5; Falk-Peterson et al. 2007) with substantially shorter periods of high production in the higher latitudes.  This suggests not only a strong temporal component, but a strong regional component. The largest seasonal effects are seen in the surface and nearshore waters, however there are seasonal patterns in the deeper waters as well, with Arctic cod (adults, YOY and larval cod) and copepods moving to deeper waters during the winter months. 

Seasonal migrations and changes in abundance, particularly of the younger life stages are an important consideration for environmental consequence analysis (ECA) and the relative vulnerability of different VECs.  Based on information provided in this review, there are some general pan-Arctic patterns of seasonal sensitivity.  However, for the purposes of assessing the spill response options, a regional approach is recommended.  Surface waters along ice-edges and in leads, particularly during the breakup and early ice-free period are important concentration points for copepods and juvenile cod, as well as migratory birds and mammals.  During the ice-free period, the ice edge continues to be an important feeding ground for large marine mammals as well as ice-seal pups.  Nearshore waters, are also a concentration point for water column resources, particularly juvenile forms and migratory species (e.g. toothed whales, sea birds).  As in more temperate waters, estuarine lagoons and rocky intertidal habitat during the ice-free season are important nursery grounds and act as a convergence point for aquatic, avian, and terrestrial species (Dunton et al. 2006).  During the open water season, nearshore waters and shallow shelf waters are also important feeding and breeding grounds for walrus.  Portions of the nearshore and shallow shelf are staging and feeding grounds for nesting sea birds and their chicks (Robertson 2013).  The midwater pelagic region is a concentration point for shoals of arctic cod and feeding migratory whales.  In the winter months, Arctic cod and Calanoid copepods move to deeper waters.  Leads and polynyas are concentration points for VEC species during the winter, in particular toothed whales.

For further detailed information on Arctic species, communities, and trophic linkages, please refer to Section 2.

 

1.3 Future Research Considerations

The review of the physical environment described by the authors in this section led to suggestions of further research which can reduce remaining uncertainties; these recommendations are summarized below while recommendations that are particularly important for improving Arctic NEBA are listed separately. 

  1. Seasonal and interannual fluctuations. The prevailing environmental characteristics such as extreme cold temperatures, variable solar irradiation and ice coverage affect the marine ecosystem and its populations. Seasonal productivity ties in with the degree of solar irradiation and ice coverage and directly affects species population recruitment and migration patterns.
    1. A better understanding of ice, and the corresponding variance of faunal distributions and life-history patterns on a spatial (regionally and throughout the pan-artic region) and temporal (annual and interannual) scale would be especially useful to protecting areas of densely populated VECs. 
  2. Interface habitats. The physical structure of the Arctic provides additional interface as well as open water environments that are occupied by VECs.
    1. Interface environments that are stable (e.g. shorelines, sediment/water interface, edges of multi-year ice, convergence of water masses near subsurface topographic features) have generally been mapped throughout most of the Arctic. Where they have not been mapped they need to be. Seasonal use patterns by VECs need to be summarized by these interface ECs (see Section 2). 
    2. Interface environments that are variable (annual ice edges including polynyas and under-ice brine channels, fresh water releases and their transport, barrier island beaches/shorelines should be characterized by indicating most likely locations in the Arctic for these less permanent features. Understanding of patterns of use by VECs will be strengthened under work proposed in Section 2. 
  3. Alterations in chemical processes. We need to better understand how cold temperatures and extreme weather interacts with broken and more permanent ice to affect the chemical processes associated with the fate of oil compounds and natural biodegradation rates (Sections 3 and 5). 

1.3.1 Priority Recommendations to Enhance NEBA Applications in the Arctic

The recommendations presented below indicate where increased knowledge of physical environment would result in reducing existing uncertainties in NEBA assessments. No prioritization has been made to the list; for some of the recommendations, surrogate data may be already available. 

  1. Arctic ecosystem consequence analyses need to include impact assessments of open pelagic waters as well as shorelines, river discharges, sediment/water interfaces, ice/water interfaces, convergence zones and surface layers near the air/water interface. These ECs may receive concentrated oil exposure depending on the type of OSR actions employed, and should receive further study.
    1. The location of primary ECs used by VECs could be added to regional sensitivity maps already available. 
    2. The locations where less permanent ECs can be prevalent are also important to VEC distribution and need to be located generally so that site specific assessments of their presence can be added to environmental sensitivity maps for reference during response actions. The environmental features that control the locations of these less permanent features and means to determine their presence during a response action are an area of additional study.

1.4 Further Information

Authors Jack Q. Word and Lucinda S. Word (ENVIRON) 

References

Blenkinsopp SA, G Sergy, K Doe, G Wohlgeschaffen, K Li, M Fingas

1996

Toxicity of weathered crude oil used at the Newfoundland offshore burn experiment (NOBE) and the resultant burn residue

Spill Sci Tech Bull 3(4):277-380

Buist I, S Potter, SE Sorstrom

2010

Barents Sea Field Test of Herder to Thicken Oil for In situ Burning in Drift Ice 

Proceedings, 33rd AMOP Technical Seminar, Halifax NS

Buist I, S Potter, T Nedwed, J Mullin

2011

Herding surfactants to contract and thicken oil spills in pack ice for in situ burning

Cold Regions Sci Tech 67(1-2):3-23

Carmack E, P Wassmann

2008

Food webs and physical-biological coupling on pan-Arctic shelves: unifying concepts and comprehensive perspectives.

Prog Oceanogr 71:446-477

Dokken ST, P Winsor, T Markus, J Askne, G Bjork

2002

ERS SAR characterization of coastal polynyas in the Arctic and comparison with SSM/I and numerical model investigations

Remote Sensing of Environ 80:321-335

Dunton KH, T Weingartner, EC Carmack

2006

The nearshore western Beaufort Sea ecosystem: Circulation and importance of terrestrial carbon in arctic coastal

Prog Oceanogr 71:362-378.

Environment Canada

2005

MANICE:Manual of Standard Procedures for Observing and Report Ice Conditions

Environment Canada, Canadian Ice Service, Ottawa, Ontario CAN

Fichot CG, K Kaiser, SB Hooker, RMW Amon, M Babin, S Belanger, SA Walker, R Benner

2013

Pan-Arctic distributions of continental runoff in the Arctic Ocean

Nature Scientific Reports No.1053. pp. 1-6

Fingas M

2011

An overview of in-situ burning

Oil Spill Science and Technology. Chapter 23. Elsevier Inc.

Forest A, C Lalande, J Hwang, M Sampei, J Berge

2013

Bio-mooring arrays and long-term sediment traps: key tools to detect change in the biogeochemical and ecological functioning of Arctic marine ecosystems

Presentation to Arctic Observing Summit, 2013; Laval University et al. pp. 1-20.

Fritt-Rasmussen J, PJ Brandvik A Villumsen, EH Stenby

2012

Comparing ignitability for in situ burning of oil spills for an asphaltenic, a waxy and a light crude oil as a function of weathering conditions under arctic conditions

Cold Regions Sci Tech 72:1-6

Gow AJ and WB Tucker III

1990

Sea ice in the polar regions, p 47-122.

In: Polar oceanography, Part A. WO Smith, Jr (ed.)

Grebmeier JM, WO Grebmeier, RJ Conover

1995

Biological processes on Arctic continental shelves: Ice-ocean-biotic interactions.

In Arctic oceanography: Marginal Ice Zones and Continental Shelves. Smith WO and JM Grebmeier (eds). Am Geophys Union, Was

Guenette CC ad P Sveum

1995

In situ burning of emulsions R&D in Norway

Spill Sci Tech Bull 2(1):75-77

Jason NH (ed)

1994

In situ burning oil spill workshop proceedings

MS, Dept of the Interior. NIST Special publication 867. 101 p

Kroncke I Tan TL, Stein R

1994

High benthic bacteria standing stock in deep Arctic basins

Polar Biology 14:423-428.

Kroncke, I

1994

Macrobenthos composition, abundance and biomass in the Arctic Ocean.

Polar Biology 14(8): 519-529.

Luers J and J Bareiss

2010

The effect of misleading surface temperature estimations on the sensible heat fluxes at a high Arctic site – the Arctic Turbulence Experiment 2006 on Svalbard (ARCTEX-2006).

Atmos Chem Phys 10:157-168.

McPhee MG, R Robins, M Coon

2005

Upwelling of Arctic pycnocline associated with shear motion of sea ice.

Geophys Res Letters 32:L10616; 4 p.

Mullin JV and MA Champ

2003

Introduction/overview to in situ burning of oil spills

Spill Sci Tech Bull 8(4):323-330

Norton M

Arctic Ocean bathymetric features, on-line graphic

http://upload.wikimedia.org/wikipedia/commons/d/d4/Arctic_Ocean_bathymetric_features.png

Olli K, P Wassmann, M Reigstad, TN Ratkova, E Arashkevich, A Pasternak, P Matrai, J Knulst

2006

Suspended concentration and vertical flux of organic particles in the upper 200 m during a 3 week ice drift at 88°N.

Progr Oceanogr; j.pocean.2006.8.002.

Petrich C, J Karlsson, H Eicken

2013

Porosity of growing sea ice and potential for oil entrainment

Cold Regions Sci Tech 87:27-32

Pistruzak WM

1981

Dome Petroleum's oil spill research and development program for the Arctic

International Oil Spill Conference, IOSC. March 1981. p 173-181

Robertson M

2013

Coastal and marine bird usage of the Beaufort Sea

BREA Results Forum; Inuvik, NWT February 21, 2013

Rodrigues J

2009

The increase in the length of the ice-free season in the Arctic

Cold Regions Sci Tech 59(1):78-101

Scholz D, SR Warren Jr, AH Walker, J Michel

2004

In situ burning: The fate of burned oil

API , American Petroleum Institute; Washington DC. 40 p.

Stirling I

1980

The biological importance of polynyas in the Canadian Arctic.

Arctic 33(2):303-315.

Walsh JJ

1995

DOC storage in Arctic seas: the role of continental shelves.

In Arctic oceanography: Marginal ice zones and continental shelves, Smith WO and JM Grebmeier (eds). Am Geophys Union, Wash

Welch HE, MA Bergmann, TD Siferd, KA Martin, MF Curtis, RE Crawford, RJ Conover, H Hop

1992

Energy flow through the marine ecosystem of the Lancaster Sound Region, Arctic Canada.

Arctic 45:343-357.

Winsor P and G Bjork

2000

Polynya activity in the Arctic Ocean from 1958 to 1997

Jnl Geophys Res 105(C4):8789-8803.

Yang J

2008

Interannual variability of the Arctic ocean Ekman transport and upwelling.

Poster presentation; Woods Hole Oceanographic Institution, Department of Physical Oceanography

Carmack E, Wassman P

2006

Food webs and physical-biological coupling on pan-Arctic shelves: Unifying concepts and comprehensive perspectives

Progress in Oceanography 71:446-447

Falk-Peterson S, Pavlov V, Timofeev S, Sargent JR

2007

Climate variability and possible effects on arctic food chains: The role of Calanus

Arctic Alpine Ecosystems and People in a Changing Environment. Springer, 434pp.

AMAP

1998

AMAP Assessment report: Arctic Pollution Issues

Arctic Monitoring and Assessment Programme (AMAP), Oslo, Norway. Xii+859pp

Buist I, Dickins D, L Majors, K Linderman, J Mullin, C Owens

2003

Tests to determine the limits to in situ burning of thin oil slicks in brash amd frazil ice

Proceedings, AMOP Technical Seminar No 26, Vol 2, pp. 629-648.

Dickins D

2011

Behavior of oil spills in ice and implications for Arctic spill response

Proceedings, Arctic Technology Conference, Houston Texas, 7-9 February6 2011. 15 p.

SL Ross, DF Dickins, Envision Planning Solutions

2010

Beaufort Sea oil spills state of knowledge | Review and identification of key issues

Environmental Research Studies Funds, Report No. 177

Dickins DF

2004

Advancing oil spill response in ice covered areas

Report submitted to the Prince William Sound Oil Spill Response Institute and US Arctic Research Commission, Cordova Alaska and

Gerdes B, R Brinkmeyer, G Dieckmann, E Helmke

2005

Influence of crude oil on changes of bacterial communities in Arctic sea-ice

FEMS Microbiology Ecology 53:129-139.

2.0 ARCTIC ECOSYSTEMS AND VALUABLE RESOURCES

Executive Summary

Photo 2-1. Arctic field study (Jack D Word)

Photo 2-2. Arctic field study (Jack D Word)

Photo 2-3. Arctic field study (Jack D Word)

Photos 2-1, 2-2, 2-3 Arctic field study (Jack D Word)

In order to minimize the potential impacts of an oil spill, valuable ecosystem components (VECs) that potentially become impacted should be indentified. Each compartment where oil might end up contains its own set of VECs with their own sensitivity and resilience to oil. Apart from the identification of VECs, their distributional patterns by life stage within environmental compartments (ECs) both in time and space are of importance. Except for the multi-year ice environments or species that are fixed to demersal or shoreline environments, the organisms and their young undergo seasonal migratory patterns with many species occupying Arctic ECs on a temporary basis. While these distributional patterns for migratory species and resident species are becoming better known, their locations during an oil spill needs to be characterized so that the optimum spill response options with the least amount of environmental consequences to VECs can be identified. Arctic specific VECs tend to congregate at interface habitats like the surface layers at or near the air/water interface (SML), ice edges and under ice environments, polinyas, sediment/water in demersal nearshore and offshore locations, shorelines, and convergence zones for different water masses. Key topics for further study would focus on the importance of these interface habitats for the diversity and long and short term functioning of Arctic ecosystems. Once the main VECs have been identified, research should establish the seasonal distribution patterns of the life stages and population levels of VECs within each EC, especially within the interface environments. Resilience of VECs determines to a large extent the long term population level effects that would occur after an oil spill. For identified VECs a proper and generic resilience metric should be developed so that relevant information on VECs can be applied in net environmental benefit analysis (NEBA) decision making.

2.1 Introduction

The objective of this section is to compile existing information on the dominant organisms that comprise the communities associated with the Arctic marine waters and to identify the valuable ecosystem components (VECs).  For the purposes of this review, a valuable ecosystem component (VEC) is a species of the marine ecosystem that is identified as having scientific, social, cultural, or economic importance. VECs may be determined on the basis of any or all of these important concerns.  The VECs defined here are based on five qualities that include their importance in supporting the Arctic marine food webs, as well as ecosystem services.  In addition, consideration is given to those species that may act as indicators of potential effects of oil spill response measures on the different ecosystem compartments. 

  • Taxa that are important to the function of Arctic food webs: Certain taxa play a critical role in maintaining and supporting the ecosystems and other trophic levels, as well as supporting the ecosystem resources and function.   For example, many species in each of the ECs of the Arctic rely on copepods and Arctic cod as a primary food source.  The removal or reduction of populations of copepods, krill or cod would have a significant impact on ecosystem function in the Arctic, as well as impacting upper trophic levels that may represent the ecosystem services for that realm (e.g. traditional fisheries for Bowhead whales are supported by copepod and krill food resources).
  • Taxa that are representative of pelagic, benthic, deep-sea, and sea-ice realms: Taxa and age classes important to marine food webs in each of the primary ecosystem compartments were included as potential VECs. 
  • Taxa that are relatively abundant: While abundance is not a core characteristic of keystone species, the removal or significant reduction in population levels of species that are relatively abundant or have a greater number of food-web linkages are more likely to impact ecosystem function than species that are less common or form fewer food web linkages.
  • Taxa that may have cultural or commercial importance: Ecosystem services are incorporated in VEC determinations.  This can include both economic values (e.g. commercial fisheries) and social values.  Arctic communities are tightly linked to ecosystem services with goods and services extending beyond typical economic values of natural resources.  Most of the goods and services for Arctic communities have cultural and social values and are dependent on other food web components.  While some of the traditional ecosystem services are captured here, often they are locally defined and may include species not listed here.
  • Taxa that are well suited to impacts analysis:    Species that are well suited to experimental approaches for evaluating the effects of OSR alternatives were included in this list of VECs.  This includes species for which there are toxicity testing methods that can or have been used to evaluate effects at the individual or population level.   Species characteristics for toxicity evaluations include sensitivity, availability, and the ability to withstand laboratory handling and stress.  Copepods and Arctic cod are common test models for the Arctic.

Recently, there have been several notable international efforts to consolidate data to provide a more pan-Arctic understanding of the biological communities of the Arctic.  As part of the International Polar Year (IPY; 2007-2008), specimen collections and data sets from numerous research institutes and government agencies were collated, reviewed and entered into a centralized database.  As part of the CENSUS, the Arctic Organism Database (ArcOD) has allowed for pan-Arctic reviews of species distributions throughout the Arctic.  The IPY was also the incentive for a number of research programs to address data gaps.  The RUSALCA program is an international effort to better understand the environment of the Beaufort, Chukchi, Bering Strait, and Siberian system.  This has included studies on circulation, benthic substrates, benthic and demersal communities, fish, zooplankton and birds and mammals.  Similar efforts have been conducted in the Atlantic Sector, in the High Canadian Arctic and the Eurasian Arctic.  The current review includes the findings of recent pan-Arctic studies and reviews, recent research programs, and peer-reviewed literature. 

2.2 Knowledge Status

2.2.1 Habitats of the Arctic

The Arctic marine waters generally include those waters above the Arctic Circle; however, the southern boundaries of the Arctic are variously defined by physical, biological, and political boundaries.  For the purposes of this review, the Arctic will be defined in a manner consistent with the Arctic Ocean Diversity program and the Arctic Register of Marine Species (ARMS) which is based on biologically relevant physical criteria (areas within the seasonally average 2 °C surface isotherm or the median maximum sea-ice extent; Sirenko et al. 2011).  Based on this definition, the Arctic includes the Arctic Ocean and associated coastal seas and bays bounded by the North American and Eurasian continental landmasses, including the northern portions of the Bering Sea, the Bering Strait, the Norwegian Sea, the Labrador Sea, Disko Bay, the Lincoln Sea, and Hudson Bay (Figure 2-1).  The Atlantic portion of the Arctic is sometimes referred to as the Atlantic sector and includes the waters surrounding Greenland and Iceland, as well as the Norwegian Sea and into the Barents Sea, those areas most heavily affected by the advancement of Atlantic waters.

Figure 2-1. The Arctic Region and Major Water Bodies 9 (Map created by Brad Cole http://geology.com/world/arctic-ocean-map.shtml)
Figure 2-1. The Arctic Region and Major Water Bodies 9 (Map created by Brad Cole http://geology.com/world/arctic-ocean-map.shtml)

Figure 2-2. Bathymetric Features of the Arctic Ocean (Base map is from IBCAO http://www.ngdc.noaa.gov/mgg/bathymetry/arctic/)
Figure 2-2. Bathymetric Features of the Arctic Ocean (Base map is from IBCAO http://www.ngdc.noaa.gov/mgg/bathymetry/arctic/)

The Arctic Ocean is a central deep ocean divided into four abyssal plains by prominent ridges surrounded by shallower continental shelves.  The only deep water connection to the world’s oceans is through Fram Strait to the Norwegian Deep and Atlantic Ocean (Figure 2-2).  A secondary connection to the Atlantic is through the Baffin Bay and the Labrador Sea and the Bering Strait, linking the Arctic Ocean to the Pacific.  The shelves comprise nearly 50% of the area in the Arctic and are a dominant feature of the East Siberian, Laptev, Kara, and Barents Seas; the shelves are relatively narrow in the Beaufort and Chukchi Seas.  Sandy and soft-bottom substrate dominates the Arctic, with finer silts and clays found in the outer shelves and deep basins.  Coarser sand and gravel are predominant substrate on the inner shelf and nearshore areas.  Although less common, there are notable and important harder substrate boulder fields and rocky intertidal areas, particularly in the vicinity of northern Greenland, Svalbard, and other island groups in eastern Canada.

Figure 2-3. Sea Ice Extent (National Snow and Ice Data Center, Boulder, CO)
Figure 2-3. Sea Ice Extent (National Snow and Ice Data Center, Boulder, CO)

The predominant environmental influences on aquatic food webs in the Arctic are sea ice and light.  A substantial portion of the Arctic Ocean is covered in ice throughout the year, while the adjoining waters are seasonally covered in ice with varying open-water periods coinciding with periods of 24h bright daylight.  Approximately 14 to 15 million km2 of the Arctic region are covered in ice during the winter months (Figure 2-3).  In the summer, the northernmost portions of the Arctic remain ice-covered (approximately 4 to 7 million km2).  The lack of sunlight during the Arctic winter in combination with ice-covered waters limits primary production in Arctic waters, resulting in a period of low secondary productivity.  The periods of low production are punctuated by periods of extremely high production during breakup and periods of open water in spring and summer.  The combination of open water and long, bright days creates intense periods of primary production, grazing by primary consumers, and predation in the higher trophic levels.  Whereas the aquatic communities are limited during the winter, the diversity and abundance of species increases dramatically in the summer.  Thus the nearshore and open water communities vary considerably during these seasonal periods.

A secondary effect of the breakup and snow melt periods is a shift in temperature and salinity, particularly in the nearshore areas.  The Arctic waters are stratified throughout the year; however, the freshwater input from extensive river systems in the Russian and Canadian Arctic result in coastal corridors of estuarine to freshwater conditions during breakup.   Nearshore species assemblages change throughout the open water season in response to the transition in salinity and water temperature.

In many respects the Arctic seas are shared waters and as such have similar components in their aquatic food webs.  However, regional differences and geographic isolation have also created some notable differences in the food web components. As in other oceanic basins, the Arctic region includes nearshore, pelagic, and deep-sea food webs.  In addition, the Arctic includes communities associated with the annual and multi-year ice.  An understanding of food webs and the key species (valuable ecosystem components) within those food webs allows is an important component of the environmental consequences analysis.

2.2.2 Arctic Food Webs

Marine communities of the Arctic can be divided into three compartments or realms:  the pelagic, benthic and sea-ice.  While each of these compartments have a closely associated assemblage of organisms, they are linked to each other with many overlapping species.

Our understanding of food webs and trophic linkages within the food web are built upon several different types of data.  Direct observations of feeding behavior and stomach contents analysis are the most common types of information used to determine prey items and linkages.  However, the ability to collect this type of data in certain habitats is difficult, particularly in Arctic and deep water environments.  More recent methods have been developed that evaluate stable isotope ratios in tissues and in lipids.  In the Arctic, the ratio of N13 and N15 has been an effective method for evaluating trophic levels.  The presence of long-chained C20:1(n-9) and C22:1(n-11) fatty acids and alcohols has also been used as an indication of whether Calanoid copepods are a component of the diet.

In this section we briefly summarize the food webs for each of these different realms and identify VECs for the Arctic marine ecosystems (Figure 2-4).  Following this section, the components of each of these different food webs is discussed in more detail.

2.2.2.1 Pelagic Communities

The pelagic food web is controlled by light and ice cover, altering the growth conditions for phytoplankton and the ability of surface-oriented predators to access prey.  In early spring, increasing light and ice melt result in the release of ice algae into the water column and a dramatic increase in phytoplankton growth (Falk-Peterson et al. 2005).  Some of the highest rates of primary production occur in these marginal ice zones (MIZ).  While phytoplankton blooms are initiated by an increase in light, the magnitude and duration of the blooms are controlled by the nutrient concentrations (Tremblay et al. 2012). Nutrients increase in surface waters during the low production of the polar night and allow for the high production (approximately 50% of the total annual production) in the MIZ.  The duration of the phytoplankton blooms is limited by nutrient availability and the decrease in light availability during the fall results in only one significant bloom event in the Arctic Ocean.  The increase in open water and phytoplankton blooms begins in the boreal waters and progresses northwards towards higher latitudes over the spring and summer (Falk-Peterson et al. 2005). The predominant groups of phytoplankton include prasinophytes, diatoms, haptophytes, green flagellates, dinoflagellates, and chrysophytes, with blue-green algae (cyanobacteria) in the southern regions (Hsiao 1978, Sakshaug 2004; Li et al. 2009, Fujiwara et al. 2014).

The zooplankton community in the Arctic is dominated by copepods, particularly the Calanoid copepods Calanus glacialis and C. hyperboreus.  Both species are endemic to Arctic waters, with all life stages found in the Arctic (Sakshaug 2004). Calanus finmarchicus is a subarctic species that is found in the Atlantic domain, but does not reproduce in Arctic waters (Falk-Petersen et al. 2007).  The diatom –Calanus food chain is considered to be critical to the overall production in the Arctic.  Calanus spp. take advantage of early ice-algae blooms, continue feeding through the planktonic diatom blooms, converting low energy sugars into a high energy lipid reserves (Niehoff 2007).  They create lipid stores that are rich in longer-chained fatty acids and alcohols; a characteristic that allows them to over-winter in a non-feeding state.  The combination of rich lipid reserves and their large size make C. glacialis and C. hyperboreus a key prey item for higher level consumers throughout the Arctic.  While less lipid rich and smaller in size, C. finmarchicus is a valuable food resource in the Atlantic sector, particularly the Barents Sea.  Other pan-Arctic copepods include Oithona similis and Metridia longa.  The subarctic species Neocalanus spp., Eucalanus bungii, Pseudocalanus spp. and M. pacifica are found in the Pacific domain (Sakshaug 2004; Griffiths and Thomson 2002).

Figure 2-4. Arctic Food Webs
Figure 2-4. Arctic Food Webs

Euphausiids (krill) are a subarctic herbivorous species that is abundant in portions of the Atlantic domain (Thysanoessa inermis) and in the Bering Strait-Chukchi-Beaufort region (T. longicauda and T. raschii; Suydam and Moore 2004; Letessier et al. 2009).  While not as common as in the Antarctic, euphausiids are a key prey resource for higher-level consumers, in particularly the Bowhead whale (Brinton 1962).  Other pelagic invertebrates that act as secondary consumers include amphipods, squid, and jellyfish.  Hyperiid amphipods are large, free-swimming amphipods that feed on both smaller zooplankton and Calanoid copepods (Auel et al. 2002).  The species Themisto libellula is a pan-arctic species found associated with sea-ice and shelf waters, whereas the species T. abyssorum and Cyclocaris guilelmi are more closely associated with outer shelf and deep waters of the Arctic (Auel et al. 2002; Kraft 2012).  Second to copepods, hyperiid amphipods are a common food item for fish, seals, and birds.  The squid,Gonatus fabricii is abundant in the Arctic and subarctic waters of the North Atlantic, with the squid Berryteuthis magister more commonly found in the Pacific Siberian-Chukchi waters (Gardiner and Dick 2010; Roper and Young 1975).  Squid are agressive predators and can move easily from the surface to deeper waters of the Arctic, being an important vertical integrator of marine food webs (Navarro et al.2013).  Arctic squid are an important prey item for narwhals (Monodon monoceros), White whales (Delphinapterus leucas), porpoise, and some seals.  Jellyfish are common in Arctic waters and can occur in abundance, representing a important consumer of zooplankton (Gardiner and Dick 2010).

Throughout the Arctic, Arctic cod (Boreogadus saida) and Polar cod (Arctogadus glacialis) represent a critical link between the zooplankton community and higher trophic levels (e.g. seals, toothed whales).  Both species are truly pan-Arctic occurring in all marine waters of the Arctic and are widely distributed throughout the Arctic, occupying nearshore, pelagic, and sea-ice habitats, residing both at depth and near the surface waters, depending upon age and season (Breines et al. 2008; Madsen et al. 2009).  Both B. saida and A. glacialis can be found in small numbers or in large, densely packed schools.  The primary prey for the Arctic gadids is Calanoid copepods and hyperiid amphipods (Sufke et al. 1998; Lonne and Gulleksen 1989; Bradstreet and Cross 1982; Frost and Lowry 1984).  Capelin (Mallotus villosus) are also an important secondary consumer throughout the Arctic, particularly in the Barents Sea where they are the primary link between C. finmarchicus and Atlantic cod (Gadus morhua; Blanchard et al. 2002; Hamre 1994; Mehl and Yaragina 1992; Titov et al. 2006).  Capelin are energy dense fish that move throughout the more estuarine nearshore waters and the Arctic waters.  Arctic and polar cod, capelin, and herring are important food resources for higher trophic levels including larger fish (e.g. Atlantic cod), marine mammals, and birds.

Anadromous fish are primarily found in the sub-adult and adult stages in the nearshore brackish waters of the Arctic.  Arctic char are circumpolar and can be numerous in waters of the Russian Arctic, where lower salinity waters extend across much of the shelf (Craig 1984; Mecklenburg et al. 2011; Sherman and Hempel 2008).  Cisco (Coregonus spp.) and salmonids will also use the nearshore zone during periods of high river flow and ice melt.  Anadromous fish feed on a variety of nearshore resources including benthic copepods and amphipods, herring and capelin (Dunton et al. 2012).  They represent an important prey resource for marine mammals and humans while in their freshwater habitats.

Pelagic fishes of the Arctic deep water regions are not as well understood as nearshore and shelf species.  Studies that have focused on deep waters have found that Arctic cod are numerically important throughout the year, with some stratification by water properties and by age group (Geoffrey et al. 2013; Parker-Stettner et al. 2011). Other midwater and deep water fishes common in other oceanic basins have been observed in the Arctic, including Myctophids and Gonostomidae (Reist and Majewski 2013; Dolgov et al. 2009; Jorgensen et al. 2005).  The diel vertical migration patterns observed in other systems may not occur throughout the year in the Arctic due to the polar day and night.

2.2.2.2 Benthic and Demersal Communities

Benthic communities in the Arctic are influenced by substrate type, presence and interaction with seasonal or permanent ice, salinity and temperature.  Intertidal and nearshore subtidal communities are often limited by direct contact with ice, seasonal freezing of soft substrates, and scouring and scraping behavior of melting ice.  Additional limitations in community diversity and abundance in the nearshore environments are highly variable salinities in areas influenced by large river systems in the Beaufort, Barents, Kara, and Laptev seas.  Soft substrates dominate the Arctic seas, with silts and clays occupying the deep water basins and the outer shelf; fine sands and silts are common across the shelf, while coarser sands and cobbles limited to isolated portions of the shelf and the nearshore zone.

Benthic macrofauna in the Arctic, like most oceanic basins, are dominated by polychaetes, bivalve mollusks, crustaceans (e.g. amphipods and isopods), and echinoderms.  In the nearshore and inner shelf communities, Arctic mollusks often define the benthic communities and are a key prey item for higher trophic levels (e.g. walrus, bearded seals).  The salinity tolerant bivalve clam Portlandia arctica is a dominant species in nearshore zones and is able to rapidly repopulate areas of disturbance.  The Greenland cockle, Serripes groenlandicus, is a common circumpolar bivalve found up to 100 m depth on a variety of substrates and is a main component of the diet for walrus and bearded seals.  Macoma calcarea is also a common component in Siberian-Chukchi-Beaufort and Barents-Kara-Laptev shelf communities, as well as in the fjords of northern Canada and Baffin Island (Grebmeier and Cooper 2012; Denisenko 2007; Filatova and Zenevich 1957).  Other dominant clams in the shelf include Astarte sp.,Mya truncata, Tellina sp. and Yoldiella solidula.  In deep water benthic habitats, bivalves are smaller and less common, with species that are common to other oceanic regions (Axinopsida, Nucula, andNuculana.

As with clams, amphipods play an amplified role in the benthic communities of the Arctic, relative to other oceanic regions.  Arctic amphipods include both infaunal and epifaunal species.  The most widely distributed species across the Arctic are Ampelisca eschrichti, Anonyx nugax, Arrhis phyllonyx, Gammarus setosus, and Byblis gaimardi (Piepenburg et al. 2011; Dunton et al. 2012).  Of particular interest are the Lysianassid amphipods; a species rich group of epibenthic omnivorous amphipods that are key scavengers in the Arctic and deep sea waters.  Many are especially adapted to scavenging with specialized mouthparts and extendable guts for food storage.  In shallower waters, Lysianassid amphipods may have a more diverse diet.  Lysianassid amphipods are domenate invetebrate macrofauna in certain environments.  On tidal flats, Onisimus litoralis constitutes up to 95% of the macrofaunal density (Weslawski et al. 2000).

Polychaetes are among the most abundant infaunal species of the shelf and deep water benthos throughout much of the Arctic (Bluhm et al. 2011).  Species that represent the Arctic shelf are species observed in temperate habitats and include Maldane sarsi, Spiochaetopterus sp., Chone sp., Lumbrineris sp., Capitella capitata, and Eteona longa (MacDonald et al. 2010; Renaud et al. 2007).  Echinoderms in the Arctic are the dominant epibenthic megafauna, with ophiuroid brittle stars occurring throughout the shelf.  Dense aggregations of brittle stars can be found in areas of organic enrichment, such as polynyas (Piepenburg et al. 1997).  The sea urchin, Strongylocentrotus droebachiensis is also found in Arctic waters and is an epibenthic omnivore, often grazing along the bottom of rocky or soft substrates.  Decapod crustaceans are less common in the Arctic and are primarily represented by the shrimp Pandalus borealis and Pandalopsis dispar and the crab Chionoecetes spp. and the Red king crab (Paralithodes camtschaticus; Bluhm et al. 2009; Iken et al. 2010; Orlov and Ivanovo 1978).

Epibenthic or demersal fish communities in the Arctic are dominated by sculpins (Cottids) and eelpouts (Zoarcids); taxa that are speciose, eurybathic, and pan-arctic (Mecklenberg et al. 2011).  Arctic cod are also an important demersal fish found associated with the benthic zone at all depth ranges (Majewski and Riest 2013).  The genera Myoxocephalus and Lycodes are genera that well represented in the Arctic (Mecklenburg et al. 2011).  The Arctic flounder (Pleuronectes glacialis) and the Greenland halibut (Reinhardtius hippoglossoides) are also important epibenthic predators.  Demersal fish generally feed on benthic infauna, as well as small or juvenile fish.  Common predators include seals, birds, and some larger fish.  As such demersal fish act as a link between benthic infauna and epifauna and higher trophic levels (Dunton et al. 2012).

2.2.2.2 Sea-ice Communities

Sympagic communities include those organisms that live in close association with sea-ice.  In some cases, species are obligate to the sea ice, but many are representatives of the pelagic or benthic community that have a portion or all of their life cycle in the ice (Melnikov 1997).  There appear to be differences between the seasonal ice communities and those that inhabit the permanent ice.  However, to a great extent the species assemblage appears to be similar with differences in abundance and biomass.  Brine channels that form in the ice-water interface create a protected environment for an algal community dominated by pennate diatoms (Melnikov 1997; Sakshaug et al. 2009).  Harpacticoid and cyclopoid copepods are the primary consumers of the ice-algae moving within the brine channels as well as along the ice bottom (Kramer 2010).  The sympagic amphipods (Gammarus wilkitzkii, Onismus spp. and Apherusa glacialis) represent a critical link between the sympagic algal and copepod communities and higher trophic levels (Hop et al. 2000; Melnikov 1997; Arndt and Swadling 2006).  The herbivorous amphipod, A. glacialis, is a primary food source for the Little auk, Alle alle.  Both G. wilkitzkii and Onismus spp. are motile predators that feed on the sympagic copepods and are important prey items for fish (e.g. cod) and sea birds.  The amphipod G. wilkitzkii has a life span of up to six years and generally prefers multi-year ice (Arndt and Swadling 2006). 

Sea ice provides a productive substrate and protective shelter.  As such there are a number of pelagic or benthic species that spend a portion of their life in close association with sea ice.  Pelagic copepods will perform diel migrations to feed at the ice-water interface, particularly during the ice melt when the ice algae and early season phytoplankton blooms can account for more than 50% of the copepods lipid reserves (Melnikov 1997; Arndt and Swadling 2006).  One and two-year-old Arctic and polar cod find shelter in the shelves and fissures of the sea-ice, feeding on sympagic amphipods and pelagic copepods (Lonne and Guilliksen 1989; Gradinger and Bluhm 2005).  Adult fish are seldom observed in close association with the sea ice (Hop et al. 2000).  Ice amphipods and cod are subject to strong predation by top carnivores including seals and sea birds.

2.2.2.4 Mammals and Birds

Arctic mammals include species associated with the sea ice and pelagic species.  Ice seals include Ringed seals (Phoca hispeda) that feed primarily on cod and hyperiid and sea-ice amphipods when young, Bearded seals (Erignatus barbatus) that feed on clams, and Hooded (Cystophora cristata) and Harp (P. groenlandica) seals that feed on cod, capelin, squid and pelagic amphipods (Bradstreet and Cross 1982; NAMMCO 2005a).  Walrus are found in the Bering-Chukchi, the Laptev Sea, and in the Barents-Greenland-High Canadian Seas (NAMMCO 2005b).  Walrus feed primarily on clams of the inner continental shelves (Outrides et al. 2003; Bluhm and Gradinger 2008).  Whales of the Arctic include both baleen and toothed whales.  Bowhead whales (Balaena mysticetus) feed on copepods and euphausiids in the open Pacific and Atlantic waters (Rice 1998; NAMMCO 2005c).  The White, narwhal, and Orca (Orcinas orca) whales are the dominant toothed whales of the Arctic.  White whales and narwhals feed primarily on cod and capelin.

Arctic seabirds are dependent on Arctic marine resources for all or most of their energy requirements while they are in the region.  Most seabirds are migratory arriving as spring blooms and breakup begins. Arctic birds that forage in the open pelagic are mostly alcids, gulls, skuas, and terns (Huettmann et al. 2011).  Other taxa tied to marine food webs are sea ducks, most notably eider ducks. 

2.2.2.5 Communities of Special Significance

Polynyas, estuarine lagoons, and rocky substrate habitats represent important but less common habitats in the Arctic.  The species that occur in these areas are not necessarily unique in Arctic or boreal waters, but they are found either in great abundance or in an assemblage of species that are well adapted to that habitat.

Polynyas are openings or leads in the sea-ice that form due to currents or water temperatures generally in nearshore areas.  They are a seasonal feature that allows light to penetrate into the water column and allows for direct access of the water surface in the absence of ice.  Narwhals remain in close association with pack ice and congregate in great abundance in polynyas using this time for most of their annual feeding.  Pelagic copepods and fish also congregate in these areas.  Sediments associated with polynyas are highly organically enriched during the polar winter, and the abundance and biomass of the more motile components of the benthic community respond with increased abundance and biomass; strong evidence for pelagic-benthic coupling.

Estuarine lagoons also represent another productive ecosystem, receiving organic input from terrigenous sources (e.g. riverine systems).  Estuarine benthos (including amphipods, polychaetes, benthic copepods, clams, and snails) in lagoons are protected from ice or ice scour, enabling them to take advantage of organically enriched sediments.  In response to the protected habitat and enriched benthic communities, epibenthic fish use the lagoons as feeding and nursery grounds.  Anadromous fish, particularly char, cisco, and salmonids will move along the brackish nearshore zone as young adult and adult fish, taking advantage of the availability of food resources.  As salinities increase in the lagoons, pelagic invertebrates and small fish will move into the protection and production of the lagoons.  Finally, the productive lagoons are an important feeding ground for sea birds, in particular Eider ducks.

Rocky substrate is less common in the Arctic, predominantly occurring in the Atlantic region of the Arctic.  The communities associated with rocky intertidal and subtidal habitats are generally similar to those that occur in the northern Atlantic, with brown macroalgae (e.g. Fucus and Laminariales), barnacles, serpulid worms, mussels, and motile scavengers such as amphipods.  Hard substrate in the deep sea occurs along the ridges, however, there is currently little data associated with the fauna on the Arctic ridges.

2.2.3 Pelagic Realm

The pelagic environment includes those waters associated with the nearshore zone, the continental shelf and the deep water basins of the Arctic.  The distribution of pelagic fish and invertebrates is defined largely by salinity, temperature, and light.  In the extensive shelf waters of North America and the Russian Arctic there are two distinctly different bodies of water, the nearshore brackish waters and the Arctic Surface Water (ASW) or Atlantic Water (AW; Craig 1984).  The occurrence of a shore-parallel band of turbid, warmer (5-10°C) brackish (10 - 25‰) water is a characteristic feature of the Arctic coastlines (Craig 1984).  The ASW includes all open waters, extending a depth of approximately 100 m.  The ASW is typically colder (-1° to 10° C) and more saline (28‰ – 34‰) than the nearshore waters.  Underlying the ASW is the Arctic Intermediate Water (AIW), layers of denser and colder waters that extend from approximately 75 to 450 m (lower).  The shallower portions of the AIW tend to have temperatures <2°C and salinities ranging from 34.7 to 34.9‰.  The deeper portions of AIW have temperatures that are 0-3°C and salinities greater than 34.9‰.  During certain portions of the year, the deeper water may be slightly warmer than the shallower waters, with pelagic species using the deeper water as refuge from colder waters.  AW on the other hand, pushes northward into the Arctic during the spring and summer months.  Temperatures in the Eurasian Arctic increase as AW pushes northward in the spring and summer months.  This push of AW is an important transport mechanism for larval dispersal and for boreal species to enter the Arctic.  The Arctic Deep Water (ADW) occupy the abyssal basins of the Arctic, with stable environmental conditions with temperatures ranging from -2°C to -10°C and a salinity of 34.9‰.

2.2.3.1 Phytoplankton

The phytoplankton population and primary production in Arctic waters are controlled primarily by light and nutrient availability.  Ice cover limits light availability in the water column throughout the winter season. Phytoplankton blooms develop as the ice opens, with propagation of the algal bloom following along a latitudinal gradient from the southern Barents Sea in March – April to August – September in the Fram Strait and Arctic Ocean (Figure 2-5).  Once the ice recedes, there is a gradual increase in phytoplankton abundance (Falk-Peterson et al. 2007). 

Figure 2-5. Timing of Plankton Blooms in the Arctic Oceans. From Falk-Petersen et al.2005</a>
Figure 2-5. Timing of Plankton Blooms in the Arctic Oceans. From Falk-Petersen et al. 2005

While light controls the timing and seasonality of primary production in the Arctic, the bloom duration and seasonal production are controlled by nutrient availability (Tremblay et al. 2012).  Nutrient limitations in the Arctic are exacerbated by a persistent stratified water column, caused by salinity and temperature differences resulting from riverine input and melting pack ice.  The nearly permanent stratification limits replenishment of nutrients from deeper, nutrient rich waters (Hsiao 1978). Chlorophyll-aestimates indicate that phytoplankton abundance is highest in the nearshore band, gradually decreasing with distance from shore.  Unlike population cycles in more temperate waters, conditions in the Arctic do not allow for a second phytoplankton bloom during the late summer.  Although nutrients in the water column slowly recover with periodic upwelling events, solar radiation begins to decline rapidly following the equinox.  With the combination of short periods of light and nutrient limitations, primary production in the high Arctic is among the lowest in the world.  The highest primary production is found in areas with energetic mixing with deep waters (Figure 2-6; Bering Strait, Fram Strait and the Atlantic sector).  Nearshore and shelf areas having variable production depending upon the extent of stratification and import of nutrients from terrestrial sources (Tremblay et al. 2012; Carmack et al. 2006). 

Phytoplankton in the Arctic includes prasinophytes, diatoms, haptophytes, green flagellates, dinoflagellates, and chrysophytes, with blue-green algae (cyanobacteria) in the southern regions (Hsiao 1978, Sakshaug 2004; Li et al. 2009, Fujiwara et al. 2014).  Recent field investigations have documented the distribution and potentially changing functional roles of picoplankton (<2µm) and nanoplankton (2-20 µm) due to changes in sea ice coverage in the Arctic Ocean (e.g. Coupel et al. 2012).  The Sea of Okhotsk and western Barents Sea are the most speciose regions, with the deep Arctic Ocean having the fewest species (Melnikov 1997; Sakshaug 2006).  Diatoms and flagellates are most abundant, followed by lower densities of dinoflagellates and chrysophytes.  Centric diatoms are the most common planktonic diatoms, withChaetoceros spp. and Thalassiosira spp. associated with spring blooms.  Pennate diatoms are less common in the phytoplankton, but are dominant in the ice-algal community.  The microplankton Phaeocystis is also common throughout the sub-Arctic coastal basins and can form blooms of several billion cells per m3 (Sakshaug 2006)  Flagellates were more predominant at offshore stations in colder, less turbid water and lower nutrient concentrations (Hsiao 1978).  Horner (1984) found that in the western Beaufort both diatoms and flagellates were found in the open waters, with distribution controlled on a vertical rather than horizontal scale.  Li et al. (2009) found that the nearshore phytoplankton community shifted from larger nanoplankton towards picoplankton and bacterioplankton with salinity and distance.  The picoplankton community was dominated by the unique pan-Arctic, cold-adapted Micromonas.

Figure 2-6. Annual Primary Production in Arctic (gC/m2/year) [Source: Carmack et al. 2006]
Figure 2-6. Annual Primary Production in Arctic (gC/m2/year) [Source: Carmack et al. 2006]

Sea-ice algae represent an important component of primary production, particularly in regions with permanent ice.  The ice algae community is discussed more fully in later sections.  However, it is important to understand that the spring blooms of ice algae precede the phytoplankton community and provide a critical food source for zooplankton.

2.2.3.2 Zooplankton

Copepods and euphausiids represent the most important zooplankton group in terms of energy transfer to upper trophic levels (i.e. Arctic cod, birds, baleen whales).  Among the large copepods (1-5 mm adult size), the Calanus species Calanus hyperboreus and C. glacialis are dominant throughout the Arctic region.  The subarctic species Calanus finmarchicus and Oithona atlantica are also common in Atlantic Water and adjacent coastal waters (Sakshaug 2004).  While not considered to be an Arctic species, C. finmarchicus, it is widely distributed across the Arctic by the northern Atlantic current that moves northward along the Norwegian coast through Fram Strait and into the Barents Sea and onto the Arctic Ocean (Falk-Petersen et al. 2007).  Other pan-arctic copepods include O. similis and Metridia longa.  In Pacific waters Neocalanus spp., Eucalanus bungii, Pseudocalanus spp. and M. pacifica are also common in the Pacific Water (Sakshaug 2004; Griffiths and Thomson 2002).  Smaller copepods dominate the brackish shelf waters Russian Arctic including Limnocalanus macrurus and Drepanopus bungei (Peters et al. 2004).

Herbivorous krill are also an important component of the pelagic food web throughout the Arctic.  The euphausiids Thysanoessa raschii and T. inermis are common in the Pacific and Atlantic arctic, with T. rachii also numerically important in the middle and Coastal Bering shelf.  In the subarctic Atlantic, T. longicauda can be common; whereas T. longipes and Euphausia pacifica are more common in the Pacific Waters.  Both T. longipes and E. pacifica are important prey items associated with bowhead whales in the western Beaufort Sea (Brinton 1962).

The diatom – Calanus food chain is considered to be critical to the overall production in the Arctic (Falk-Petersen et al. 2007).  Calanoid copepods comprise 50% to 80% of the mesozooplankton biomass in the Arctic.  Calanus spp. take advantage of early ice-algae blooms and feed through planktonic diatom blooms, converting low energy sugars into a high energy lipid reserves.  Calanoid copepods create lipids that are rich in longer-chained fatty acids and alcohols; a characteristic not shared by other smaller copepod species (Peters et al. 2004).  The combination of rich lipid reserves and their large size makeCalanus spp. a key prey item for a number of higher level consumers, in particular Arctic cod. 

While the three dominant calanoid copepods co-occur, they differ in their life histories.  Calanus hyperboreus is the most polar species and is most common in the deep sea areas of the Arctic, including the Arctic Ocean, Fram Strait, and the Greenland Sea.  It is a large copepod (4.5-7 mm) that can overwinter at depths of 500 m to 2000 m (Falk-Peterson et al. 2007).  C. hyperboreus has a 2-yr life span during periods of high productivity which can extend to 3 to 5 years during periods of extensive ice cover.  Calanus hyperboreus reproduce in deep waters in the winter/spring prior to the phytoplankton blooms (Niehoff 2007) with the naupliar forms developing just under the ice surface, feeding on sympagic (ice-associated) algae.  Egg releases in early March appear to provide a lipid-rich food source prior to significant primary production (Darnis et al. 2012).

Calanus glacialis is a smaller copepod (3-4.6 mm) that is found primarily on shelf waters of the Arctic.  C. glacialis has a life span of 1-3 years, spawning throughout the Arctic prior to the yearly spring bloom.  Most larvae reach the C1 copepod stage (early adult form) by ice break up allowing the nauplii to feed on ice algae in the relative protection of the pack ice (Falk-Petersen et al. 2007; Niehoff 2007).  The C1 copepods are then able to take advantage of the phytoplankton blooms in early summer.  It descends to deep areas on the shelf (200-300 m) to enter diapause and over winter (Falk-Petersen et al. 2007).

Calanus finmarchicus is the smallest of the three Calanus species (2-3.2 mm) and has less lipid-rich reserves.   While nauplii and copepodites may be advected into the Arctic Ocean and other northerly waters, C. finmarchicus does not appear to reproduce in Arctic waters (Niehoff 2007) but relies more heavily on reliable ice-algal blooms for reproduction.  As with other calanoid copepods, C. finmarchicusrelease eggs prior to the phytoplankton blooms, which develop during the Arctic summer prior to over wintering at depths of 500 to 2,000 m (Falk-Petersen et al. 2007).

In late-spring/early summer, reproduction rates for other copepods and euphausiids increase following blooms of phytoplankton, microzooplankton, and small copepods (Pseudocalanus spp.).  Omnivorous zooplankton, such as Metridia longa feed on both living plankton, as well as the organic carbon spikes following blooms.  Reproduction in such species appears to be more independent of the plankton blooms.

2.2.3.3 Neuston

Neuston refers to a community of microbial, plant (phytoneuston) and animal (zooneuston) species that live at the water’s surface.  Surface communities include obligate species that are in the surface layer throughout their life cycle, as well as facultative species that occupy the surface only during larval stages.  Arctic neuston have a number of adaptations that allow them to live in the harsh surface environment, including oil droplets to shield them from ultra violet (UV) exposure and keep them at the surface, and diel migrations that help them compensate for the temperature extremes (Zaitsev 1971).

While large temperature fluctuations and extreme UV exposure limit species diversity, marine phytoneuston abundance per unit volume may be a thousand times higher than the underlying bulk water populations (Zaitsev 1971; Word et al. 1986).  Marine phytoneustonic communities are dominated by small diatoms and flagellates.  As with abundance, the production per unit volume can be many times that of the underlying water.  The zooneustonic community in Arctic waters include chaetognathans, copepods, euphausiids, hyperiids, and in smaller numbers of rotifers, pontellids, cladocerans, decapods, and fish larvae and eggs (Zaitsev 1971).  Microbial populations, including bacterioneuston, may be 10,000 times denser than bacterioplankton populations (Hardy 1982; Sieberth 1971).

The neuston community is somewhat self-contained, with zooneuston feeding on phytoneuston.  Neuston is an important component of the pelagic food web, with zooplankton, fish, birds, and marine mammals feeding at the surface, particularly along convergence zones where the surface populations are highest (Zaitsev 1971).  Bowhead whales, as well as other baleen whales, have been observed skim-feeding at the surface (Koski and Miller 2002).

2.2.3.4 Other Pelagic Invertebrates

The pelagic invertebrate community includes a wide variety of other species, including amphipods, mysids, shrimp, squid, jellyfish, pteropods, chaetognaths, and ichthyoplankton.  There are only two Arctic species of pelagic shrimp and they do not appear to be a common component of the nekton (Hopcroft et al. 2008).  This section will focus on certain elements of the pelagic invertebrate community that are important food web components.

Krill

Unlike the Antarctic, euphausiids are not abundant in Arctic waters.  However they can be difficult to accurately assess in ice-covered waters as they seek shelter in the pockets and fissures in the ice (Percy and Fife 1985).  Euphausiids are not considered to be a truly Arctic species, rather they are advected from the Bering Sea Water inflow (Suydam and Moore 2004) or from the Atlantic (Letessier et al. 2009).    The euphausiids Thysanoessa raschii and T. inermis were important prey items associated with bowhead whales in the western Beaufort Sea (Brinton 1962).  Euphausiids are found from the surface to the deep midwater pelagic areas, in depths greater than 500 m. 

Amphipods

Free-swimming amphipods are a key component of the Arctic food web, representing a link between zooplankton and higher level consumers such as fish, marine mammals, and birds.  Hyperiid and Lysianassid amphipods are among the common taxa found in Arctic waters and are a dominant component of the zooplankton abundance and biomass (Auel et al. 2002).  The hyperiid amphipod, Themisto libellula is a pan-Arctic species that can occur in high numbers in the ice free waters of the Arctic.  Based on the high amounts of C20:1 (n-9) and C22:1 (n-11) fatty acids and alcohols found in T. libellulatissues, their diet was domianted by Calanoid copepods.  Further analysis indicated a close association with ice-algal production.  This confirms field sampling data showing Themisto sp. associated wtih sympagic communities.

Themisto abyssorum is more closely associated with deepwater communties.  Themisto abyssorum is  considered to be a boreal species and is found in close association with the Atlantic Water that moves northward through Fram Strait.  Abundance generally decreases from east to west, dropping from >200 ind/m2 over the contiental slope north of Spitsbergen to <40 ind/m2 in the central Arctic (Auel et al. 2002).  It does not show a similar lipid signature seen in T. libellula, indicating that copepods are not a dominnant component of the diet.  Rather, the deepwater species is more likely an omnivore and scavenger.

Cyclocaris guilelmi, is also an epipelagic species, occurring in the deep-waters of the Arctic.  As wtih T. abyssorum, C. guilelmi appears to be Atlantic in origin but is found throughout the Arctic ocean.  Kraft (2012) found C. guilelmi to be a dominant component found in deepwater traps in Fram Strait.  The population appeared to be stable with peaks from August to February.

Cephalopods

Cephalopods are a predator of and a key prey item for many of the VECs in the Arctic.  Squid move vertically through the water column, integrating marine resources throughout the pelagic environment.  The squid Gonatus fabricii is the most commonly reported species in the Arctic, with numerous records in the Atlantic sector, the Barents Sea and the high Canadian Arctic (Gardiner and Dick 2010). Berrytheuthis magister is the predominant squid species found in the Pacific domain.  Squid exhibit all manners of vertical distribution: near-surface dwellers (to 50 m), vertical migrators that either move into the surface waters at night or just move higher in the water column, and near-bottom dwellers.  In addition, some species exhibit ontogenetic descent moving progressively deeper as they age (Roper and Young 1975).  Cephalopods are voracious predators feeding on crustaceans, fish, other squids, and zooplankton.  Based on isotope analysis, squid are occupying different trophic levels in different regions (Navarro et al. 2013) indicating that squid are able to shift their diet based on availability.  The isotope ratios for Arctic squid indicated a trophic level of 3 to 5.  Squid are an important component of marine mammal and sea bird diets, including narwhals, White whales, walrus, murres and fulmars (Gardiner and Dick 2010).

Jellyfish

Gelatinous zooplankton are poorly understood in the Arctic, largely due to the difficulty of capturing them with traditional sampling methods.  However, they are considered to be a substantial sink for primary and secondary production (Purcell et al. 2009).  In the Beaufort Sea, the scyphomedusa, Chrysaora melanaster is among the most common gelatinous zooplankton in shelf waters 25 to 75 m in depth (Purcell et al. 2009).  In shallower waters, there was a more diverse community dominated by the delicate medusa Bolinopsis infundibulum, other small cnidarians, and ctenophore species occurred immediately underneath the sea ice (Purcell et al. 2009; Raskoff et al. 2010b).  Other common species include Sminthea arctica in the midwater depths, and Atolla tenella which was found in high abundance in the deep Canadian Basin (Raskoff et al. 2010b). 

Large populations of gelatinous zooplankton have been observed throughout the Arctic, particularly in at convergences, fronts, and polynyas (Ashjian et al. 1997).  In such areas, medusa and ctenophores can have a substantial grazing impact.  In the eastern Canadian high Arctic, the ctenophore Mertensia ovum was estimated to consume up to 9% per day of the C. hyperboreus population and 3-4% of the C. glacialis population.   Other prey items included hyperiid amphipods, pteropods, and smaller copepods.  Other food resources for gelatinous zooplankton include detritus and algal cells and in some cases small or juvenile fish (e.g. larval capelin and herring; Raskoff et al. 2010b).

 

2.2.3.4.1 Krill

Unlike the Antarctic, euphausiids are not abundant in Arctic waters.  However they can be difficult to accurately assess in ice-covered waters as they seek shelter in the pockets and fissures in the ice (Percy and Fife 1985).  Euphausiids are not considered to be a truly Arctic species, rather they are advected from the Bering Sea Water inflow (Suydam and Moore 2004) or from the Atlantic (Letessier et al. 2009).    The euphausiids Thysanoessa raschii and T. inermis were important prey items associated with bowhead whales in the western Beaufort Sea (Brinton 1962).  Euphausiids are found from the surface to the deep midwater pelagic areas, in depths greater than 500 m. 

2.2.3.4.2 Amphipods

Free-swimming amphipods are a key component of the Arctic food web, representing a link between zooplankton and higher level consumers such as fish, marine mammals, and birds.  Hyperiid and Lysianassid amphipods are among the common taxa found in Arctic waters and are a dominant component of the zooplankton abundance and biomass (Auel et al. 2002).  The hyperiid amphipod, Themisto libellula is a pan-Arctic species that can occur in high numbers in the ice free waters of the Arctic.  Based on the high amounts of C20:1 (n-9) and C22:1 (n-11) fatty acids and alcohols found in T. libellulatissues, their diet was domianted by Calanoid copepods.  Further analysis indicated a close association with ice-algal production.  This confirms field sampling data showing Themisto sp. associated wtih sympagic communities.

Themisto abyssorum is more closely associated with deepwater communties.  Themisto abyssorum is  considered to be a boreal species and is found in close association with the Atlantic Water that moves northward through Fram Strait.  Abundance generally decreases from east to west, dropping from >200 ind/m2 over the contiental slope north of Spitsbergen to <40 ind/m2 in the central Arctic (Auel et al. 2002).  It does not show a similar lipid signature seen in T. libellula, indicating that copepods are not a dominnant component of the diet.  Rather, the deepwater species is more likely an omnivore and scavenger.

Cyclocaris guilelmi, is also an epipelagic species, occurring in the deep-waters of the Arctic.  As wtih T. abyssorum, C. guilelmi appears to be Atlantic in origin but is found throughout the Arctic ocean.  Kraft (2012) found C. guilelmi to be a dominant component found in deepwater traps in Fram Strait.  The population appeared to be stable with peaks from August to February.

2.2.3.4.3 Cephalopods

Cephalopods are a predator of and a key prey item for many of the VECs in the Arctic.  Squid move vertically through the water column, integrating marine resources throughout the pelagic environment.  The squid Gonatus fabricii is the most commonly reported species in the Arctic, with numerous records in the Atlantic sector, the Barents Sea and the high Canadian Arctic (Gardiner and Dick 2010). Berrytheuthis magister is the predominant squid species found in the Pacific domain.  Squid exhibit all manners of vertical distribution: near-surface dwellers (to 50 m), vertical migrators that either move into the surface waters at night or just move higher in the water column, and near-bottom dwellers.  In addition, some species exhibit ontogenetic descent moving progressively deeper as they age (Roper and Young 1975).  Cephalopods are voracious predators feeding on crustaceans, fish, other squids, and zooplankton.  Based on isotope analysis, squid are occupying different trophic levels in different regions (Navarro et al. 2013) indicating that squid are able to shift their diet based on availability.  The isotope ratios for Arctic squid indicated a trophic level of 3 to 5.  Squid are an important component of marine mammal and sea bird diets, including narwhals, White whales, walrus, murres and fulmars (Gardiner and Dick 2010).

2.2.3.4.4 Jellyfish

Gelatinous zooplankton are poorly understood in the Arctic, largely due to the difficulty of capturing them with traditional sampling methods.  However, they are considered to be a substantial sink for primary and secondary production (Purcell et al. 2009).  In the Beaufort Sea, the scyphomedusa, Chrysaora melanaster is among the most common gelatinous zooplankton in shelf waters 25 to 75 m in depth (Purcell et al. 2009).  In shallower waters, there was a more diverse community dominated by the delicate medusa Bolinopsis infundibulum, other small cnidarians, and ctenophore species occurred immediately underneath the sea ice (Purcell et al. 2009; Raskoff et al. 2010b).  Other common species include Sminthea arctica in the midwater depths, and Atolla tenella which was found in high abundance in the deep Canadian Basin (Raskoff et al. 2010b). 

Large populations of gelatinous zooplankton have been observed throughout the Arctic, particularly in at convergences, fronts, and polynyas (Ashjian et al. 1997).  In such areas, medusa and ctenophores can have a substantial grazing impact.  In the eastern Canadian high Arctic, the ctenophore Mertensia ovum was estimated to consume up to 9% per day of the C. hyperboreus population and 3-4% of the C. glacialis population.   Other prey items included hyperiid amphipods, pteropods, and smaller copepods.  Other food resources for gelatinous zooplankton include detritus and algal cells and in some cases small or juvenile fish (e.g. larval capelin and herring; Raskoff et al. 2010b).

 

2.2.3.5 Fish

For the purposes of this review, fish in the Arctic can be divided into four general groups, the pelagic shallow and midwater species, anadromous species, and demersal fishes.  The bottom fish include nearshore, shelf, and deep water species.  Pelagic fish are the least diverse group representing 13% of the 242 recorded Arctic species (Bluhm et al. 2011; Mecklenberg et al. 2011).   There are 31 species considered to be anadromous.  The remaining species are marine demersal fish.

2.2.3.5.1 Pelagic Fish

Throughout the Arctic, gadoids (cod) represent a critical link between the zooplankton community and higher trophic levels (e.g. seals, White whales).  Arctic cod (Boreogadus saida) and Polar cod (Arctogadus glacialis) are truly pan-arctic species occurring in all marine waters of the Arctic.  Cod are widely distributed throughout the Arctic, occupying nearshore, pelagic, and sea-ice habitats, residing both at depth and near the surface waters, depending upon age and season.  Both B. saida and A. glacialis can be found in small numbers or in large, densely packed schools.  In the scientific literature there is confusion on the common names “Arctic cod” and “Polar cod” at times referring to either B. saida or A. glacialis.  While these species are similar in appearance and life history, gene sequencing has shown that they are genetically distinct (Breines et al. 2008; Madsen et al. 2009). 

Both B. saida and A. glacialis have a diet dominated by pelagic or sympagic components (Sufke et al. 1998; Lonne and Gulleksen 1989; Bradstreet and Cross 1982).  Stomach contents of B. saida cod sampled in the pelagic ranges of the Beaufort Sea, northern Canadian waters, and the Barents Sea were dominated by calanoid copepods and gammarid amphipods.  Other prey items included hyperiid amphipods, mysids, and shrimps (Frost and Lowry 1984; Lonne and Gulliksen 1989).  A similar pelagic diet was observed for A. glacialis in the Northeast Water Polynya near Greenland (Sufke et al. 1998).  In the nearshore zone, the diet is dominated by copepods, gammarid amphipods, and young-of-the-year Arctic cod. 

Both B. saida and A. glacialis spawn under the ice in the winter months; however recent observations indicate that there are regional differences in the hatching season of B. saida (Figure 2-7).  Bouchard and Fortier (2011) noted that cod hatching started as early as January and extended to July in areas with significant freshwater input (Laptev/East Siberian Seas, Hudson Bay, and Beaufort Sea).  In contrast, hatching was restricted to April-July in regions with little freshwater input (Canadian Archipelago, North Baffin Bay, and Northeast Water).  The authors found that the different hatching periods resulted with different length and weight classes co-occurring throughout the Arctic.  Cod larvae generally occupy a depth of 10 – 30 m, settling to the bottom in September (Craig 1984; Graham and Hop 1995). 

As noted above, the distribution of B. saida and A. glacialis includes nearly all marine waters of the Arctic depending upon the age class and the season.  Arctic cod have been found to be a dominant component of both the pelagic and demersal communities at all depths; cod have been collected from the mixed water mass (<200 m), at the sharp halocline (200-300 m), and in the deeper Atlantic water mass (>300 to 1000 m; Majewski and Riest 2013; Norcross et al. 2012). One and two year old fish also occupy fissures and gaps in the ice pack, feeding on the sympagic fauna.  In the summer months, adult cod are dispersed in habitats ranging from coastal brackish waters to the demersal and pelagic zones of the shallow shelves, including the ice-water interface (Craig et al. 1982; Lonne and Gulliksen 1989; Gradinger and Bluhm 2004).  In autumn, as nearshore salinities increase, large shoals of cod are observed in shallow (<10 m) waters (Hop et al. 1997; Welch et al. 1993), presumably following shoreward fronts of plankton.

Acoustic surveys conducted as part of the BREA program as well as others conducted in the US Beaufort have found large shoals of Arctic cod in the water column and near the bottom along the entire shelf of the Beaufort Sea (Geoffrey et al. 2013; Parker-Stettner 2011).  There was a clear segregation between the young-of-the-year (YOY) and age 1+ fish in the summer months.  Age 1+ fish were found to aggregate in the deeper waters of the shelf  and slope at depths ranging from 200 to >1000 m while large shoals of YOY Arctic cod were found nearer to the surface (20 to 100 m) in nearshore waters extending into waters over the outer shelf and slope (Figure 2-8). The near bottom aggregations of Arctic cod at depths of 200 to 400 m appear to span the Canadian Beaufort shelf into the fall and winter months (Geoffrey et al. 2013; Benoit et al. 2008). Acoustic surveys during the winter months have found massive aggregations of adult cod at depths of 140 to 230 m under the pack ice in the Canadian Beaufort waters (Benoit et al. 2008) and at depths of 300 to 1,300 m near the North Pole (Geoffrey 2013). In winter months, adult cod appear to remain in schools at the deep inverse thermocline (160-230 m, -1 to 0°C) throughout the Polar night to avoid seal predation; whereas smaller cod (<25 g) periodically migrate into the isothermal cold intermediate layer (90-150 m) to feed on Calanoid copepods and then return to the deeper layer (Benoit et al. 2010).

An exception to the Arctic cod dominated pelagic food web is in the Barents Sea and White Sea.  The Barents Sea is located between the Arctic and boreal oceanic systems and is influenced by the variations in the Atlantic current and the Polar front.  In the southern portions of the Barents Sea, where the Atlantic water is more predominant, capelin (Mallotus villosus) is the primary link between pelagic crustaceans (e.g. copepods and amphipods) and higher trophic levels (Blanchard et al. 2002; Hamre 1994; Mehl and Yaragina 1992; Titov et al. 2006).  The importance of capelin to the Barents Sea ecosystem was demonstrated when capelin stocks decreased markedly in the 1980s, resulting in decreases in stocks of the commercially important Atlantic cod (Gadus morhua) and ringed seals.  Subsequent studies have demonstrated linkages between the location of the polar front, the population of capelin and subsequent changes in the population of Atlantic cod (Titov et al. 2006).  Capelin are also an important forage fish in the northern boreal waters of Greenland, the Sea of Okhotsk and the Bering Strait.  Unlike Arctic cod, Atlantic cod are more demersal in nature, generally feeding near the bottom.  The diet of Atlantic cod is remarkably varied, with Mehl and Yaragina (1992) reporting with over 180 prey species.  While capable of feeding on a variety of invertebrate and vertebrate prey, capelins appear to be the primary energy source.

Figure 2-8. Arctic cod (Boreogadus saida) vertical distribution during the ice-free season along the Mackenzie Slope 2012 (from Geoffrey et al. 2013).
Figure 2-8. Arctic cod (Boreogadus saida) vertical distribution during the ice-free season along the Mackenzie Slope 2012 (from Geoffrey et al. 2013).

Pacific herring (Clupea harengus pallasi) are another nearshore forage fish that represents an important link between zooplankton and higher level consumers, particularly anadromous fishes of the Bering-Chukchi-Laptev and Lofoten-Barents Sea systems (Mehl and Yaragina 1992; Hamre 1994).  Herring feed primarily in the water column on copepods, euphausiids, and mysids.  Herring spawn en masse and their eggs and larvae are a vital food source for nearshore migrating anadromous fish in the Arctic such as salmonids, cisco, and char. In the Arctic they are generally found in the nearshore waters and avoid the colder, Arctic water (Craig 1984).  While numerically less important than the Arctic cod, capelin and herring are important forage species for upper trophic level consumers.

2.2.3.5.2 Anadromous Fish

Anadromous fish include those species with a life history that includes both freshwater and marine habitats.  In the Arctic, utilization of marine waters is typically limited to the brackish waters found in nearshore corridor immediately following breakup.  Anadromous fishes in the Arctic include Arctic char (Salvelinus alpinus), least and Arctic cisco (Coregonus sardinella and C. autumnalis), broad and humpback whitefish (C. pidschian and C. nasus), Inconnu (Stenodus leucichthys), and several species of salmonids (Craig 1984; Mecklenburg et al. 2011).  These species spawn in fresh water and typically do not enter coastal waters until months, or often years, after hatch.  Thus, the most sensitive life stages for anadromous fish are spent in the Arctic rivers and lakes.  Use of marine waters is limited to feeding and migration.  Coregonids are the most common anadromous fish in the Laptev-Kara-East Siberian waters, with Arctic cisco more commonly found in the marine waters than Least cisco (Sherman and Hempel 2008; Craig 1984).  Arctic char range widely from their stream of origin and might be found in more open water during high flow years (Jarvela and Thorsteinson 1999; Johnson 1980).  When occupying the nearshore brackish water, anadromous fish feed nearly exclusively on epibenthic fauna (e.g. polychaetes, mysids, and amphipods; Dunton et al. 2012).  In turn, anadromous fish become an important food source for seals as well as subsistence fishers.

A number of salmonid species are found in river systems throughout the Arctic; however, their use of marine waters is limited.  Atlantic salmon are the most common form found in the Lapotov-Barents Sea and Hudson Bay river systems (Sherman and Hempel 2008).  Pink and chum salmon (O. gorbuscha) are the most abundant species of Pacific salmon documented in the Arctic, with populations in Russian (Yana and Lena Rivers), Canadian (Mackenzie River), Alaskan, and Norwegian waters (Sherman and Hempel 2008; Mecklenburg et al. 2002).  With the exception of some juvenile use of the brackish nearshore waters, adults are the most common life stage found in marine waters.

2.2.3.5.3 Demersal Fish

Demersal fishes are those that are found in close association with the bottom.  Sculpins (Cottidae) and eelpouts (Zoarcidae) are the most speciose fish taxa in the Arctic, comprising over 50% of the species in polar waters (Mecklenberg et al. 2011).  Both taxa are well adapted to living in the dominant substrates found in the Arctic (sand, silt, and mud), as well as rocky bottoms, with most species spending the majority of their life cycle in close association with the bottoms.  Adults often deposit eggs directly on benthic substrates or on bottom oriented vegetation.  Larval and early juvenile forms may remain on the bottom near the adults or move into the water column or into vegetation before descending to the bottom as young fish.  Common and pan-Arctic sculpin species include the Arctic Staghorn sculpin (Gymnocanthus tricuspis), the large Short-horned sculpin (Myoxocephalus scorpius: 90 cm length), and Spatulate sculpin (Icelus spatula; Mecklenberg et al. 2011).  Two genera account for most Arctic eelpouts, Gymnelus and Lycodes.  Sculpin and eelpout diets vary, but many sculpin and eelpouts feed on benthic infauna and epifauna, including polychaetes, benthic amphipods, small mollusks, and epibenthic crustaceans, with larger species feeding on fish, including cods, flounders, and smelts (Dunton et al. 2012).  Anti-freeze proteins have been found in several species of Myoxocephalus sculpins, showing their adaptation to Arctic waters (Fletcher et al. 1982).

Flatfishes or flounders live on the bottom, usually in shallow marine waters, and burrow into the surface sediment to rest and wait for prey. They eat worms, mollusks, echinoderms, crustaceans, other benthic invertebrates, and fishes. Arctic flounder (Pleuronectes glacialis) are a pan-Arctic species that prefer coastal and nearshore waters (Mecklenberg et al. 2011).

2.2.3.5.4 Deep-Sea Fish

The deep-sea fishes are perhaps the poorest known group in the Arctic.  Recently there have been several targeted efforts to sample the deep basins in the Arctic (Reist and Majewski 2013; Dolgov et al. 2009; Jorgensen et al. 2005). In the deep Canadian Beaufort, Arctic cod were the dominant species in the water column and associated with the demersal community (Reist and Majewski 2013). Other species found in midwater trawls included species found in other deep ocean basins, including Myctophids (lanternfish) and Gonostomids (bristlemouths).  The myctophids, Benthosoma glaciale and Protomyctophium arcticum, spawn above the Polar front (Dolgov et al. 2009) and were found in deep water trawls (Riest and Majewski 2013).  Other species have been recently collected from the Kara Sea, includingMyctophum punctulum (Dolgov et al. 2009; Weinrroither et al. 2010).  The wide-spread gonostomid, Cyclothone microdon, has been observed in trawls in Baffin Bay (Riest and Majewski 2013; Mecklenburg et al. 2011).  Based on observations of myctophids and bristlemouths throughout the world, their diet is dominated by pelagic crustacean, migrating to the surface from depth to feed.  It is not known if the diel vertical migrations occur in the Arctic.

Deepwater demersal fish taxa are generally similar to those found in other oceanic basins and include Zoarcids (eelpouts) and Macrourids (grenadiers).   The globally distributed macrourids, Coryphaenoides rupestris and Macrourus berglax have been found in the Baffin Bay (Jorgensen et al. 2005).  The Glacial eelpout (Lycodes glacialis) is one of the most dominant demersal fishes in the Arctic basins.  This large eelpout (~70 cm) moves along the bottom stirring up the bottom sediments to feed on small benthic crustaceans and mollusks; L. glacialis also feeds on other fishes and cephalopods (Mecklenberg and Mecklenburg 2011).  This is a truly deep water species, spending its entire life cycle at depths >1,000m.  The Greenland halibut (Reinhardtius hippoglossoides) is a right-eyed flounder typically associated with deep waters of the Arctic (200 – 1600 m), as well as deep waters of the Atlantic and Pacific Oceans.  The Greenland halibut is epibenthic, and feeds on epibenthic crustaceans, demersal fish, and other invertebrates.  Deepwater rays found in the southern Arctic waters in Alaska and Baffin Bay include Rajella spp. and Amblyraja radiata (Dolgov et al. 2009; Mecklenburg et al. 2002).

2.2.3.6 Marine Mammals

Marine mammals are both permanent and seasonal members of the Arctic and include baleen and toothed whales, seals, walrus, and Polar bear.  The following section focuses on those mammals closely linked to the marine food webs.

2.2.3.6.1 Bowhead Whale (Balaena mysticetus)

Bowhead whales are circumpolar, residing in the high latitudes from late April to October.  In the spring months, bowheads migrate northward as the sea ice breaks up.  Five stocks have been identified, including the Spitsbergen, Baffin Bay-Davis Strait, Hudson Bay-Fox Basin, Sea of Okhotsk, and Bering-Chukchi-Beaufort stocks (Rugh et al. 2003).  The latter stock overwinters in the Bering Sea.  In the spring, the Bowheads move north and east, past Point Barrow into the western Beaufort Sea (Lowry et al. 2004).  The majority of whales move into the Canadian Beaufort Sea for the summer months; however, some Bowhead whales will either remain in the eastern Alaskan Beaufort or move into the Arctic Ocean.  While Bowhead whales appear to favor continental slope waters in the spring and summer, they appear to favor the inner shelf waters (<200 m depth) of the western Beaufort in September and October (Moore et al. 2000).  As the ice cover increases in late fall, the whales migrate into the Bering Sea for the winter months. 

Bowhead whales are considered to be second order consumers (Hoekstra et al. 2002), with a diet dominated by euphausiids and calanoid copepods (Bluhm and Gradinger 2008).  Stomach content analyses indicate that, while benthic crustaceans and fish occur, their consumption is either occasional or incidental (Frost and Lowry 1984; Lowry et al. 2004).  In the Bering-Chukchi-Beaufort system, there appears to be some seasonality in the diet which is dominated by euphausiids in spring and by copepods in the summer and autumn months.  This distribution of diet is likely to be a result of opportunity, rather than selection.  Dominant species in stomach contents included the euphausiids Thysanoessa raschii and the copepods C. hyperboreus and C. glacialis (Frost and Lowry 1984; Lowry et al. 2004).

2.2.3.6.2 White Whale (Delphinapterus Leucas)

The White whale, or Beluga, is a circumpolar species inhabiting cold waters of the Arctic and subarctic waters (Rice 1998; NAMMCO 2005a).  During the winter months, White whales retreat to subarctic regions with loose pack ice and winter polynyas (Barber et al. 2001).  During the summer months, White whales live in coastal waters, estuaries, shelf breaks and deep basins.  In the Alaskan Arctic, White whales move into the Beaufort Sea in May, moving east into the Canadian Beaufort or north to the Arctic Ocean for much of the summer (Loseto et al. 2006).  In the fall as the Arctic cod congregate in the nearshore waters, White whales cross along the Alaskan mid-shelf in late August and September (Suydam and Moore 2004).  The  Eastern Chukchi White whale typically remains in the more open waters (>200 m depth) throughout the year, whereas the Eastern Beaufort Sea White whales move closer to shore when in the eastern Beaufort Sea (Suydam et al. 2001).

White whales are considered to be third order consumers (Hoekstra et al. 2002), with a diet dominated by Arctic cod (B. saida and A. glacialis), and to a lesser extent whitefish (Coregonidae) in the Russian and Greenland Arctic.  A variety of other prey items have been observed in stomach contents, including capelin, herring, smelt, sculpins, cephalopods and benthic invertebrates (Bluhm and Gradinger 2008).  Nitrogen and carbon isotope signatures indicate that they also feed on copepods and euphausiids, particularly in the spring and fall (Hoekstra et al. 2002; Frost and Lowry 1984).  Known predators of White whales include Orca whales, Polar bears, and humans (NAMMCO 2005a).

2.2.3.6.3 Narwhal (Monodon monoceros)

Narwhal occur in the deep and offshore waters of the Canadian High Arctic, the Barents and Kara Seas, eastern Laptev Sea and the waters surrounding Greenland (Sherman and Hempel 2008).  Narwhal appear to have high site fidelity, remaining in close association with the ice in winter.  Winter feeding grounds appear to be more important than summer feeding areas, with remarkable aggregations of narwhals found in polynyas.  In winter, a large number of whales can share the limited open water areas; near Greenland, between 17,000 and 19,000 narwhals were found to occupy 2% of the surveyed area (approximately 73 whales per km2 of open water (Laidre, personal communication).  Narwhals feed mostly in deep water and possibly at or near the bottom. Dives of up to nearly 1,500 m and 25 minutes are documented (Laidre et al. 2003), and there are some seasonal differences in the depth and intensity of diving (Laidre et al. 2002, Laidre et al. 2003).  Arctic cod (B. saida and A. glacialis) and the squidGonatus fabricii dominate the narwhal diet, with lesser amounts of Greenland halibut and other deep-sea fish (Laidre and Heide-Jørgensen 2005a; Bluhm and Gradinger 2008). Predators include Orca, Polar bears and humans (Hay and Mansfield 1989). 

2.2.3.6.4 Ice Seals

Ice seals of the Arctic include Ringed seals, Ribbon Seals, Spotted Seal, and Bearded Arctic seals.   Ringed seals (Phoca hispida) are the most common and widely distributed ice seal in the Arctic (Reeves 1998).  Ringed seals are relatively small seals (1.5 m) that are generally found on permanent ice or large floes, maintaining breathing holes allowing it to use ice habitats when other seals cannot (NAMMCO 2005b).  Major foods eaten are Arctic cod, nektonic crustaceans (hyperiid amphipods and euphausiids), capelin, sculpin, and sea-ice and benthic crustaceans (Bluhm and Gradinger 2008).  The balance of the diet varies seasonally.  Ringed seals are a primary prey item for Polar bear, Arctic fox and Glaucous gulls (NAMMCO 2005b).

The Bearded seal (Erignathus barbatus) is a solitary seal with a circumpolar distribution.  It is most abundant where it can reach the sea bottom to feed.  The bearded seal is generally found in the pack ice where openings are common since it cannot maintain a breathing hole.  In the Beaufort and Chukchi system, E. barbatus consumed crab, shrimp, and clams (Lowry et al. 1980).  In the Kara and Barents Seas, and Sea of Okhotsk bearded seals fed primarily on Arctic cod (B. saida) as well as shrimp (Sclerocrangon boreas) and mollusks (Finley and Evans 1983).  In the areas of NW Greenland and the Canadian High Arctic, bearded seals had a varied diet including fish, crustaceans, gastropods, cephalopods and polychaetes (Finley and Evans 1983).

Other seals that are found in the Arctic include Ribbon seals (Phoca fasciata), Harp seals (P. groenlandica), and Spotted seals (P. largha).  Both the Ribbon and Spotted seals feed primarily on Arctic cod, capelin, as well as demersal fish and large benthic crustaceans.  Harp seals feed primarily on Arctic cod and shoaling fishes, such as herring and capelin (Bluhm and Gradinger 2008).

2.2.3.6.5 Walrus (Odobenus rosmarus)

Three subspecies of walrus occur in the Arctic: the Pacific walrus (O. rosmarus divergins) in the Bering-Chukchi), the Laptev walrus (O. rosmarus laptevi) in the Laptev Sea, and the Atlantic walrus (O. rosmarus rosmarus) in the Barents-Kara, Greenland, and High Canadian Arctic waters (NAMMCO 2005c).  Walrus are extremely gregarious, often hauled out on land or ice floes, with several thousand individuals in a herd. 

Walrus feed in shallow waters (10-50 m), foraging through bottom sediments with their stiff beard bristles. Their primary diet appears to be dominated by bivalve clams, however, stomach contents analysis also indicates that other benthos are also important, including snails, echinoderms, and crabs (Outridge et al. 2003; Bluhm and Gradinger 2008).  Dominant clam species found in stomachs included Mya truncata, Serripes groenlandica, and Macoma sp.  Due to their size, tusks, and their gregarious behavior walrus predators are limited to Polar bear, Orca, and humans (NAMMCO 2005c).

2.2.3.6.6 Orca Whales (Orcinus orca)

Orcas occur in the Arctic during open water periods.  Orcas move northward from the Bering Sea or the North Atlantic.  Distribution likely varies, with movements tracking those of favored prey species or pulses in prey abundance of availability (such as seal pups or fish runs).  In the Arctic, Orcas rarely move close along or into the pack ice (Reeves et al. 2002).  The frequency and abundance of Orcas in the Arctic appears to increase during years with decreased ice coverage.  Orcas are a top predator and feed on a variety of large vertebrate prey including anadromous and pelagic fish, ringed seals, and whales.  Orcas are considered a significant threat to Bowhead whales (COSEWIC 2009).

2.2.3.6.7 Polar Bear (Ursus maritimus)

Polar bears are linked to the marine habitat through diet and daily or seasonal migration.  Although they reside on the ice during the winter, polar bears are accomplished swimmers and inhabit the open waters along the ice edge; they migrate toward land once the ice melts.  Polar bears are linked to the marine pelagic food web, feeding primarily on ringed seals, although they will also feed opportunistically on other marine mammals (Thiemann et al. 2007). 

2.2.3.7 Birds

Arctic seabirds are dependent on marine resources from the Arctic for all or most of their energy requirements while they are in the region.  Most seabirds are migratory arriving as spring blooms and breakup begins. Arctic birds that forage in the open pelagic are mostly alcids, gulls, skuas, and terns (Huettmann et al. 2011).  Other taxa tied to marine food webs are sea ducks, most notably eider ducks.

2.2.3.7.1 Black-legged kittiwakes (Rissa tridactyla)

Black-legged kittiwakes are one of the most numerous seabirds with a circumpolar distribution.  Arriving in southern portions of the Arctic in February and moving northerly through April.  Kittiwakes feed in ice floes as well as in open water, skimming the water surface or feeding from the surface.  Based on stomach contents analysis from kittiwakes in Lancaster Sound, the summer diet is dominated by Arctic cod (93% to 98%; Bradstreet 1976).  Dahl et al. (2003) indicate that kittiwakes in the vicinity of Svalbard feed primarily on capelin, Arctic cod, and hyperiid amphipods.  A similar diet was observed in Barents Sea.  Isotope analysis in the High Canadian Arctic indicated that amphipods may play an important role to fish over the course of the year (Hobson 1993).

2.2.3.7.2 Black Guillemots (Cepphus grille)

Black guillemots are a common bird in open water and amongst the ice floes.  Black guillemots are generalists.  Stomach contents analysis in Lancaster Sound showed that amphipods, copepods, and Arctic cod were all important components in guillemot diets.  Amphipods and copepods appeared to be a more dominant component of the diet when the guillemots fed along the ice edge in later spring to early summer (Bradstreet 1980), with fish being an important component of the diet throughout the year (Hobson 1993).

2.2.3.7.3 Thick billed Murres (Uria lomvia)

Thick-billed murres or Brunnichs guillemots are members of the auk family (Alcids).  Thick-billed murres over-winter in boreal regions where there are open waters.  In summer, U. lomvia congregate and breed in the Chukchi Sea, the Siberian coast, eastern Canadian Arctic, Greenland and northern Norway.  Murres are agile diving birds that consume both fish and crustaceans, with Arctic cod comprising the majority of the diet in both coastal and offshore ice edges (Bradstreet 1980; Hobson 1993).  Summer diet was more variable feeding on pelagic amphipods when cod are unavailable.  Murre chicks’ diet is dominated by Arctic cod and sculpin.

2.2.3.7.4 Northern Fulmar (Fulmarus glacialis)

Northern fulmars are long-lived (32 years) migratory birds, moving northwards to the Arctic between May and July.  Fulmars are pelagic birds, preferring shelf habitats, particularly shelf breaks or over the continental slope, though they are seldom further than 100 km from shore (Dewey 2009).  As with many other sea birds, the fulmar diet is dominated by Arctic cod, copepods and pelagic amphipods.  Fulmars also prey on the squid, Gonatus fabricii.  Seasonal analysis showed that amphipods and copepods were dominant in the diet of adult and nesting fulmars.  Arctic cod are the primary diet once chicks have hatched and during rearing.  After that time, amphipods (Hyperia sp., Gammarus sp, Themsto sp.) and copepods once again dominated the diet.

2.2.3.7.5 Common Eider (Somateria mollissima)

The Common eider is a large diving duck common in the Arctic, particularly in the High Canadian and Atlantic sectors.  Common eiders feed primarily on benthic prey including mollusks (Buccinum glacialisHiatella arctica), barnacles (Balanus balanus), decapods (e.gHyas araneus), and amphipods (Gamarellus homari; Dahl et al. 2003).  Eiders are an upper level consumer in kelp forest (Fredriksen 2003) and estuarine lagoons (Dunton et al. 2012).  Isotope analysis confirms that eiders feed at the lower trophic levels (Hobson 1993).  Unlike other Arctic species, lipids analysis and stomach contents analysis show that copepods are not an important prey item for eiders.

2.2.3.7.6 Little Auk/Dovekie (Alle alle)

Dovekies are a small marine diving bird that is circumpolar in distribution (Day et al. 1988).  This small auk overwinters in boreal waters, such as the North Sea and Norwegian Sea.  In early spring they migrate northwards to feed on the sympagic copepods and amphipods (Dahl et al. 2003).  The Little auk feeds on the herbivorous sea-ice amphipod, Apherusa glacialis (Kramer 2010).  Dovekies also rely heavily on Calanus copepods, relying on the lipid rich C. glacialis and C. hyperborealis to successfully raise their chicks (Falk-Petersen 2007).   Breeding colonies can be quite large with 30 million birds observed in northwestern Greenland. 

2.2.3.7.7 Glaucous gull (Larus glaucescens)

Glaucous gulls are pan-Arctic and are a primary avian predator in the Arctic that feed on a wide variety of fish, mollusks, crustaceans, eggs, chicks, and adult seabirds, as well as carrion.  They will often prey on young and adult birds, as well as the catch from other birds.  Arctic fox are important predators of gulls and skua, as well as other nesting birds, preying on eggs and chicks.

2.2.3.7.8 Arctic jaeger (Stercorarius parasiticus)

Arctic jaegers are a top avian predator during the summer months, migrating annually to overwinter in the Antarctic.  Jaeger primarily on fish, though they will also feed on insects and berries.  While they can catch their own food, they will often steal fish from other birds.  They will also prey on the nests of waterfowl, eating the eggs and young, as well as small mammals (e.g. lemmings).

2.2.4 Benthic Realm

Benthic communities are strongly influenced by the substrate-type and sediment grain size.  Benthic substrates in the Arctic are dominated by fine-grained sands, silts, and clays on the shelves and fine clays and silts in the oceanic basins (Bluhm et al. 2010).  Sediment in the expansive shelves of the Barents, Kara, and Laptev Seas are typically dominated by fine grained sediments (Stein et al. 2004, Semiletov et al. 2011; Cochrane et al. 2009), with areas of sandy substrate occurring in the nearshore areas.  Similarly the narrowing shelves of the Bering and Chukchi seas are fine in nature, with some sandy areas occurring in the Bering Strait.  In the central reaches of Russian, Greenlandic, and Baffin Island fjords, sediments are dominated by very fine, unconsolidated clays and silts (Aitken and Fournier 1993, Stein et al. 2004).  Sands and gravelly substrate is more common in the nearshore zone and in erosional areas.  Hard substrate is localized and includes subtidal boulder fields as well as nearshore rocky outcrops.  Notable examples are the boulder fields in the Beaufort Sea, portions of the Barents Sea, and the rocky shoreline in northern Greenland and Svalbard.

2.2.4.1 Intertidal Communities

Intertidal benthic communities in the Arctic are generally thought to be less diverse than those found in lower latitudes due to ice scour and UV exposure.  The disturbance from ice comes from direct contact of foot or anchor ice and scouring during breakup.  Such disturbance results in a continually changing benthic community dominated by fast-settling, opportunistic meiofauna (Barnes 1999).  Estimates suggest that Arctic intertidal macrofaunal communities typically have no more than 100 species, with some areas nearly devoid of intertidal macrofauna (animals larger than 0.5 mm).  In contrast intertidal boreal communities in Alaska and Norway may have over 200 to over 300 species (Weslawski et al. 2011).  Most intertidal species are circumpolar; however, their regional distribution can be highly variable.  For example, areas of the Beaufort and Siberian coasts are devoid of intertidal macro-organisms as beaches are predominately gravelly with little stability and fast ice in the winter months (Weslawski et al. 2011).  Conversely, estuarine lagoons along the Beaufort are relative benthic hotspots, with well-developed benthic infaunal communities that are protected from fast ice and coastal erosion (Dunton et al. 2012).

Rocky intertidal and shallow subtidal communities are an important component in certain portions of the Arctic, particularly the Atlantic sector (Figure 2-9).  Benthic communities in these areas are more akin to those of north Atlantic rocky intertidal and shallow subtidal habitats.  For example, in the steep rocky substrate in Svalbard (Kuklinski and Barnes 2008) the most common species include the macroalgaeFucus spp., sessile (i.e., non-mobile) barnacles (Balanus balanoides), and motile gastropods (Littorina saxatilis) and amphipods (Gammarus setosus and G. oceanicus).  Associated subtidal communities are dominated by the barnacle Balanus balanus, brittle stars (Ophiopholis sp.), motile amphipods (Calliopidae sp.), isopods (Munna sp. and Janira maculosa), sipunculid polychaete worms, and snails (Alvaniasp.).

Figure 2-9. Locations with Shoreline Dominated by Hard Substrate
Figure 2-9. Locations with Shoreline Dominated by Hard Substrate (Red outline, from Lantuit et al. 2012)

In the western Arctic, hard-substrate communities are uncommon in intertidal and subtidal waters.  However, there are isolated communities associated with patches of boulders and cobble.  Dunton et al. (1982) characterized an Arctic kelp community associated with subtidal (<10m) boulder patches in the Alaskan Beaufort Sea.  Exposed boulders and cobble provided an attachment point for Laminarian kelp (Saccharina latissima, L. solidungula, Alaria esculenta).  The community associated with the kelp beds were characterized as being similar to those found in northern Atlantic waters.  Large sponges and cnidarians were common, as were the chitons and mussels; barnacles (B. balanus), snails, and sipunculid polychaete worms were also common.  The crab Hyas coarctatus was the dominant crustacean, along with mysids, amphipods, and hermit crabs (Pagurus trigonocheirus).

The benthic communities of the coastal nearshore zone (<50 m) can be variable depending upon the salinity regime and substrate.  Infaunal biomass is variable across the Arctic, ranging from 41 gC/m2 in the Beaufort Sea to over 250 gC/m2 in the Baffin Island, Lancaster Sound area (Thompson 1982).  The nearshore zone in the Beaufort Sea has relatively low biomass and highly dynamic communities due to the gravelly nature of the shallow subtidal substrate.  In gravelly or cobble substrate meiofauna and barnacles dominate.  In sand and sandy silts, the sediment fauna is dominated by small polychaetes (Scoloplos armiger, Spio filicornis, and Chaetozone setosa) and oligochaetes.  Feder and Schamel (1976) found that species diversity, abundance and biomass increased with distance from shore, which was attributed to ice and wave action affecting stability.  Nearshore coastal areas of Lancaster Sound, the high Canadian Arctic, and Greenland, infaunal communities have higher biomass, with bivalve dominated communities including Astarte borealis, A. motagui, Serripes groenlandicus, Mya truncata, Cistenides granulata, and Macoma calcarea (Thompson et al. 1986).  Similar species complexes that also included the bivalvesPortlandia arctica, and Nuculana sp. are found along the Russian Arctic (Gebruk 2004).

Estuarine flats and lagoons can have abundant and well developed benthic communities.  Lagoon areas are often protected from fast ice and wave action, are enriched by terrestrial sources of organic matter, and have warmer water temperatures.  However they may also have highly variable salinities that can limit species diversity.  In lagoons along the Beaufort Sea and in the Kara Sea, infaunal diversity was noticeably lower in estuarine lagoons than the neighboring marine waters.  However, abundance and biomass were equivalent or greater.  As in temperate waters, the lagoons act as a rich habitat for phytoplankton, harpacticoid copepods, calanoid copepods, polychaetes (Nephtys sp., Prionospio sp., Spio sp., Terebellides sp., and Travisia sp.), Pandalus shrimp, mysids, clams (Astarte sp., Yoldia sp.,Macoma sp., Portlandia sp.), a variety of amphipods (Anonyx sp., Gammarus spp.), and isopods, anadromous fishes (Arctic and Least cisco, Arctic char, salmonids) and birds (Dunton et al. 2012; Gebruk 2004).  Many of the species found in the lagoon communities are similar to those found in boreal and temperate waters.

2.2.4.2 Shelf and Deepwater Communities

As indicated above, the majority of the benthic habitat in the Arctic basins is fine grained sand, silt, and clay.  In general, soft bottom infaunal and epifaunal communities in the Arctic are similar to those of other oceanic basin, being dominated by polychaetes, amphipods, mollusks, and echinoderms.  Abundance, biomass, and species diversity in Arctic shelf communities is considered to be similar to the lower range for boreal and temperate basins.  Community abundance, biomass, and diversity are variable throughout the Arctic, and are controlled by factors such as substrate type and availability of organic carbon appear to be the primary drivers for abundance and biomass (Thompson 1982; Grebmeier and Cooper 2012; Piepenburg et al. 2011).   In the Barents Sea, Cochrane et al. (2009) found that ice cover was inversely proportional to organic carbon and abundance.  A number of studies have noted the importance of pelagic-benthic coupling in determining the benthic assemblage (Grebmeier and Cooper 2012; Carmack et al. 2006; Carmack and Wassmann 2006).  Pelagic sources of organic carbon in the Arctic includes both water-column and sea-ice production.  In the high Arctic, the ice algae are the primary source of carbon for the benthic communities.  However, productivity in these regions is limited due to the reduced daylight and ice cover and this appears to limit abundance and biomass in the associated benthic communities.  Species assemblages have also been shown to vary across the shelf from east to west, with the southern Greenland, Norwegian, and Barents Seas affected by the species in the North Atlantic.  Similarly, the Siberian-Chukchi-Beaufort system is affected by Pacific species associated with the Bering Sea and Strait (Grebmeier 2012).

Beyond the shelf, species diversity, abundance and biomass decreases markedly with depth.  Unlike other oceanic basins, there was no increase in diversity and biomass at the mid-depths, rather all three benthic measures decrease steadily with depth (Table 2-2; Bluhm et al. 2011).  Many of the species that comprise the deep water communities are eurybenthic and are found on the outer continental shelves.  Similarly, many of the dominant deep water species are similar to those of the boreal and temperate deep water communities.  When considering deep water benthos, it is important to bear in mind that data at depth in the Arctic is limited.  Sampling methods commonly deployed in other ocean basins cannot be deployed in the Arctic.

Table 2-1.  Abundance and Biomass for Arctic Deep-Sea Benthos (Bluhm et al. 2011).

Depth Range (m)Abundance (ind. per m2)Biomass (mg C per m2)

500-1,000

2,295

436

1000-2,000

840

157

2,000 – 3,000

791

116

3,000 – 4,000

271

19

>4,000

104

10

The general taxa groups that dominate the benthic macro- and megafauna in the Arctic are similar to those of other oceanic basins, namely polychaetes, amphipods, isopods, mollusks, and echinoderms.  Other epibenthic crustaceans such as crab and shrimp are also found in the Arctic but are not as common throughout the region.  While polychaetes are the most abundant of the major taxa throughout the shelf, Arctic clams are perhaps a more important component in the Arctic food webs, relative to other regions, with a number of larger and important predators (e.g. walrus, bearded seal).  The following section discusses each of these taxa groups.

2.2.4.3 Mollusca

Bivalve mollusks are an important component of Arctic shelf communities, serving as a primary prey item for walrus and bearded seals, among other higher vertebrates.  Benthic communities of the inner shelf are often defined by the dominant bivalve species.  Bivalves (clams and mussels) can comprise up to 15% of the infauna in slope and plain sediments, but often dominate the biomass.  In substantial portions of the East Siberian Sea-Chukchi-Beaufort and Barents-Kara-Laptev shelves, bivalves were the dominant taxa (Grebmeier and Cooper 2012; Denisenko 2007; Filatova and Zenevich 1957).

The Greenland cockle, Serripes groenlandicus, is a common circumpolar bivalve found up to 100 m depth on a variety of substrates.  A fairly large (up to 112 m) and long-lived clam (up to 39 years) it is a main component of the diet for walrus and Bearded seals.  Macoma calcarea are also a common component in the shallower portions of the Siberian-Chukchi- Beaufort and Barents-Kara-Laptev seas (<50 m depth), as well as the fjords of northern Canada and Baffin Island.  Areas with organic enrichment had higher Macoma abundance (Grebmeier and Cooper 2012; Filatova 1957).  Other dominant clams in the shelf include Astarte sp., Portlandia, Mya truncata, Telina sp. and Yoldialla solidula. 

Deep-sea species are generally smaller and less abundant.  Common deep sea species include the taxa Axinopsidae, Nucula, and Nuculanacia while other species are unique to the deep sea (e.g. Bathymodiolus spp. and Abra spp.; Gage and Tyler 1991). Clams are generally suspension or deposit feeders that are well adapted to processing particulate organic matter (POM) in the deep sea.  The digestive tract in many deep-sea species is elongated with intra and extracellular digestive enzymes that allows for efficient digestion of recalcitrant forms of carbon.  In addition, deep-sea clams have modified palps that sort particles before sending them to the mouth (Allen 1979).  Some deep-sea species are carnivorous, with modified siphons that allow for predation on copepods and ostracods. 

2.2.4.4 Polychaetes

Polychaetes dominate the infaunal community on the slope, rise and abyssal plain.  MacDonald et al. (2010) found dominant families in the Arctic including Cirratulidae and Paraonidae (Aricidea sp).  Other species considered to have pan-oceanic distributions based are Capitella capitataLumbrineris minuta, Maldanid, Oweniid, and Chaetozone complexes.  There are a number of different feeding strategies used by polychaetes, including filter feeders, surface deposit feeders, subsurface-burrowing feeders, and carnivores.  On the shelf and in fjords, the species assemblages change with feeding strategies.  Renaud et al. (2007) found that the inner fjords and nearshore areas were typically dominated by few species and species that were highly motile or surface deposit feeders (e.gLumbrineris sp., Chaetozonesp.). In the mid- to outer fjords, the polychaete community was more diverse with an increase in subsurface deposit feeders.  There is a strong indication for pelagic-benthic coupling in the Arctic with the distribution of benthic fauna strongly associated with patterns in the associated water column, particularly near polynyas where regionally high production in the water column is reflected in the benthic community (Piepenburg et al. 2005).

2.2.4.5 Amphipods

Amphipods are common members of the benthic community, including both infaunal and epibenthic species.  Amphipods are broadly distributed throughout the world’s shelf and deep sea basins.  The most widely distributed species across the Arctic are Ampelisca eschrichti, Anonyx nugax, Arrhis phyllonyx, Pontoporeia femorata, Gammarus setosus, and Byblis gaimardi (Piepenburg et al. 2011; Dunton et al. 2012).  Petrova and Dzhurinskiy (2012) found that Ampeliscidae were more common in the inner shelf areas, while Pontoporeiidae were more common in Bering Strait.  In the Gulf of Finland, a similar trend was observed with Pontoporeiidae more abundant at the mouth of the Gulf and Ampelisca more common near the head of the Gulf.  Amphipods are primarily deposit feeders or active scavengers, either free-burrowing or living in tubes.  Infaunal amphipods are commonly used in toxicological evaluations of whole sediment, with established methods for testing whole sediments or spiked water samples.

Lysianassid amphipods are a species rich group of epibenthic omnivorous amphipods that are key scavengers in the Arctic and deep sea waters.  Many are especially adapted to scavenging with specialized mouthparts, extendable guts for food storage, and a highly motile foraging behavior.  Premke (2003) has reported on food-fall scavenging activities of the Lysianassid amphipod Eurythenes gryllus  in deep Arctic waters of the Norwegian Deep.  However, in shallower waters, Lysianassid amphipods may have a more diverse diet.  Lysianassid amphipods  dominate invertebrate macrofauna in certain environments.  On tidal flats, Onismus litoralis constitutes up to 95% of the macrofaunal density (Weslawski et al. 2000). 

Many species of Lysianassid amphipods are barotolerant. E. gryllus may be easily retrieved and decompressed from deep water with no apparent deleterious effects, as long as temperature is kept below 4 °C (Sainte-Marie 1992).  Their geographic distribution and depth range is also considered to be extensive, being found throughout the world to depths up to 2,500 m above the bottom (Krapp et al. 2008, Sainte-Marie 1992, Premke 2003).  Anonyx nugax is also a common benthic amphipod in the Arctic found associated with the sea ice and with abyssal communties as deep as 1700 m.

2.2.4.6 Decapod Crustaceans

In general, decapod crustaceans are less common in the Arctic.  There are however, several notable exceptions, particularly in the boreal-Arctic waters of the Bering Sea-Chukchi and the Barents-Greenland Seas.  Crab and shrimp are the primary decapods in the Arctic. These include the shrimp, Pandalus borealis and the deep water prawns Pandalopsis dispar.  In the eastern Arctic, the densest concentrations are found in the central region of the Barents Sea, Hopen Deep, Thor Iversen Bank and near the western Murman coast at depths from 200 to 350 meters.  The caridean shrimp, Nematocarcinus ensifer is found from the North Sea into the Arctic.  Epibenthic shrimp are primarily detritivores; however the diet is augmented with slow-moving prey such as gastropods, ostracods, and hydrozoan polyps (Cartes 1993).  In general, larval and juvenile demersal shrimp appear to occur deeper than adults, migrating to the shallower end of their distribution as they grow older.  Demersal shrimp are an important food source for demersal fish, in particular Alepocepalus and Macrouridae.  There is a substantial fishery for P. borealis in the Russian Arctic.

The crab, Chionoecetes spp., is native to waters in Alaska, the east coast of Canada and west of Greenland, and is an invasive species in the Barents Sea.  Main items in the Chionoecetes spp. diet in the southeastern Barents Sea are polychaetes, mollusks, crustaceans and echinoderms.  Chionoecetes spp.in the Barents Sea were recorded in waters from 39 to 387 m depth, predominantly on muddy or sandy and muddy grounds, at temperatures from –1.6° C to 5.9° C and salinity from 34.5 to 35.1 ‰ in the near-bottom layer.  In the Bering Strait and Chukchi Sea, Chionoecetes spp. has shown increased abundance (Bluhm et al. 2009; Iken et al. 2010). 

The Red king crab (Paralithodes camtschaticus) occurs in portions of the Barents Sea and was deliberately introduced to the Barents Sea at several locations during the 1960s and 1970s from the northern part of the Pacific (Orlov and Ivanovo 1978). It has continuously spread to new areas and is now distributed from the Kluge Island to east, the Goose Bank to north, and west to Lofoten and Kvænangen to west along the Norwegian coast. The expansion of the area inhabited by red-king crabs occurred during years when water temperature in Atlantic currents was higher than normal (Pinchukov and Karsakov, in press). Several studies have revealed that the crab besides being an important fishing resource, also significantly impact the bottom ecosystem in areas of high densities of crabs (Sundet and Berenboim 2008).

In the Russian waters of the Barents Sea, red-king crabs occur in areas from shallow waters to the depths below 335 m. In spring, April-May, they form spawning aggregations of individuals of both sexes, whereas in August-September, red-king crabs form separate aggregations where males aggregate in concentrations within the temperature range 4-6° С and females within 5-7° С.  Red-king crabs are benthic predators (Gerasimova and Kachanov 1997; Manushin 2008), but in areas with intensive fishing, they predominantly feed on fish offal (Pinchukov and Pavlov 2002; Anisimova and Manushin 2003). The main red-king crab predators in the Barents Sea are cod and skates (Matyshkin 2003).

Galatheid crabs (Munida spp) are widely distributed on bathyal bottoms in most deep ocean regions (Cartes et al. 2004).  Munida tenuimana is common in the north Atlantic along the middle and lower slope from depths of 300 to 1900 m (Cartes 1993).  The diet of the galatheid crabs includes polychaetes, crustaceans, and fish remains.  Galatheid crabs are also found in the hydrothermal vent communities, living within the large tube worms (Martin and Haney 2005).

2.2.4.7 Echinoderms

As in other oceanic basins, Ophiuroid brittle stars are among the most common megafauna occurring in Arctic shelf habitats.  Dominant species include Ophiocten sericeum, Ophiura sarsi, Ophiura robusta,Ophiopleura borealis, and Ophiacantha bidentata (Piepenburg et al. 1997; Gebruk 2004; Thomson 1982 Anisimova 1989).  In the Laptev Sea, populations of O. sericeum and O. sarsi were abundant (as high as 36 ind/m2) but highly variable with nearby areas devoid of brittle stars (Piepenburg et al. 1997; Sirenko et al. 2010).  Similar trends were observed in shelf habitats in Greenland and Barents Sea and the outer shelf in the Siberian Sea (Bluhm et al.2009). 

Photo 2-4. Dense aggregations of Brittle Stars
Photo 2-4. Dense aggregations of Brittle Stars

Most ophiuroids are motile epifaunal grazers using their flexible arms to feed on detritus, suspended organic material, small epifauna, and infaunal organisms.  Ophiuroids have the ability to aggregate in areas of organic carbon deposition and have been associated in the Arctic with polynyas and marginal ice zones, where ice algae and associated detritus are deposited at higher rates (Photo 2-4).  Predators of ophiuroids include shrimp, crabs, and epibenthic feeding fish such as Zoarcidae and small Macrouridae.

Other important echinoderms found on the shelf are the sea urchin, Strongylocentrotus droebachiensis, and holothoroids (sea cucumbers).  Urchins can move over the seabed rapidly to form dense aggregations when responding to food, such as patches of “phytodetritus”.  Urchins are a prey source for crab and larger predatory fish species.

At deeper sites, holothoroids become a dominant component of the demersal invertebrate community.  Holothoroids, or sea cucumbers, are ubiquitous on the abyssal plain and relatively abundant.  Sea cucumbers are detritivores, ingesting sediment and digesting the incorporated organic material.

2.2.5 Sea-Ice Realm

The sea-ice realm is defined by a complex of permanent or multi-year ice that can reach thicknesses in excess of 5 m and seasonal, first-year ice that is generally <1 to 2 m in thickness (Melnikov 1997).  The structure of sea-ice varies with the depth and age of the pack ice.  At the ice-water interface, the sea-ice surface is uneven and porous, with brine channels that extend upwards nearly 1 m into the ice pack.  Brine channels are created as the water freezes and salts are excluded from the structure of the ice.  In first-year ice, brine channel density can be 50 to 300 channels per m2 and average 0.4 cm diameter (Arndt et al. 2009).  The ice temperature and interstitial water salinity increases with increasing distance from the ice-water interface.  Temperatures in the bottom meter of ice are relatively stable, remaining between 0 and -2°C and salinities ranging from brackish to marine (4 to 40‰; Petrich and Eiken 2010).  In the middle depths of the ice pack, ice density increases, brine channels become smaller, and interstitial water salinity increases (>100 ppt).  Temperatures in the mid to upper ice pack are more variable and can range from -35°C to >5°C.  In older multi-year ice, the upper layers of ice are comprised of fresh water ice, with more fully developed structure and no brine channels.   With the exception of some microbial and bacterial species, the majority of the flora and fauna associated with the sea ice are found on or in the bottom ~20 cm of ice.

Sympagic organisms are those species that live in and on sea ice and include autochthonous species (those that spend their entire life history in the ice) and allochthonous species (those that migrate to the sea ice from the benthic or pelagic realm to spend a portion of their life cycle associated with the sea ice.  The sympagic community includes ice algae, ciliates, nematodes, rotatorians, acoel turbellarians, cyclopoid and harpactacoid copepods, amphipods, and polychaetes (Melnikov 1997).  These species in turn support larger pelagic and avian predators that are closely associated with the ice (e.g. Arctic and polar cod).  The species composition and distribution of sympagic fauna can exhibit large spatial and inter annual variations due to the origin and history of the ice, the water depth, the physical and biological characteristics of the underlying water.  Despite variation, the dominant components of the sea-ice community are similar throughout the Arctic, and include the ice algae, cyclopoid and harpacticoid copepods, sea-ice amphipods, and polar or Arctic cod.  The following section will focus on these species, with reference to the associated upper trophic levels. 

2.2.5.1 Ice Algae

ce algae form the base of the sea-ice food web.  Although primary production by sea ice algae is generally low compared to phytoplankton, they comprise the primary source of fixed carbon to higher trophic levels in ice covered waters (Arndt and Swadling 2006).  Gradinger (2010) noted that in portions of the Arctic Ocean, sea ice primary production accounts for 50% of the total annual production.  In addition, the sea-ice blooms in the spring coincide with the ice melt, representing an important early source of nutrition for zooplankton (Bluhm et al. 2011).  Ice algae primary production is controlled by available light and nutrients.  During the Polar night, production is limited by light availability.  However, increases in light during April and May initiates algal blooms.  As with phytoplankton production, nutrient limitation becomes a controlling force during these blooms.

The sea algal ice communities are diverse and variable throughout the Arctic, with hundreds of reported sympagic species (Sakshaug et al. 2009).  While centric diatoms are a dominant form in the phytoplankton community, ice algae are dominated by pennate diatoms (Melnikov 1997; Sakshaug et al. 2009).  Of 21 species observed in sea ice near Barrow, Alaska, only one centric diatom (Thalassiosirasp) was found in the sea ice.  Common diatoms found in sea ice include Nitzchia spp, Navicula spp, Pinnularia, Pleurosigma, Gomphonema, and Surirella.  Sea ice algae often exists as single cells within brine channels, with pennate diatoms being well suited to living in the limited space of brine channels.  Along the bottom of the ice, sea ice algae may also form dense mats or long strand communities (Melosira arctica).  Melnikov (1997) found long strands of the diatom Melosira arctica under multiyear ice that supports faunal communities.

Ice algae also form communities in melt-ponds of the pack-ice surface.  The melt-pond communities occur in the summer and autumn in brackish to marine waters that are created by melting snow and ice and marine water that penetrate upwards through channels in the ice.  Melt pond communities are dominated by unicellular green algae and flagellates that are commonly found in Arctic freshwater basins at altitudes from sea level to 3000 m (Melnikov 1997).  The communities shift to diatom-dominated communities as salinity in the melt ponds increases (von Quillfelt et al. 2009).

2.2.5.2 Sympagic Copepods

The sympagic mesofaunal community is dominated by harpacticoid and cyclopoid copepods (Kramer 2010).  These smaller copepods are generally considered to be epibenthic in nature, with feeding habits and morphology that makes them well suited to living in and under the sea-ice.  Harpacticoid and cyclopoid copepods can occur in populations as high as tens of thousands per m2 in pack ice and are several orders of magnitude greater in abundance than the surrounding water column (Kramer 2010).  The highest population densities generally coincide with the highest algal densities and the more moderate temperatures and salinities, within the bottom 20 cm of the ice pack (Gradinger et al. 1999; Bluhm et al. 2010).  As with ice algae, sympagic copepods biomass and abundance is highly variable depending upon the ice age and location.  The copepod genera Harpacticus, Halectinosoma, Tisebe and Cyclopina appear to have a circumpolar distribution (Arndt and Swadling 2006).  However, the relative proportion of the dominant taxa varies with the type of ice, the region, and with season.  For example, the copepod H. superfluxus and other Harpacticus species appear the be nearly absent from the interstices of perennial ice in the Arctic ocean and the northern Barents and Greenland Seas (Melnikov 1997) and are scarce underneath old ice.  In contrast, in the seasonal fast ice of Frobisher Bay Canada, H. superfluxuspopulations can reach >380 individuals per sq. m.  Another harpacticoid species, Halectinosoma sp. is typically found in multi-year ice in particular near Svalbard and northern Greenland.  Both species can be found in areas where young ice and multi-year ice mix (Kern and Carey 1983). 

In Frobisher Bay, Grainger (1991) found two dominant copepods, Tisbe furcata and Cyclospina schneideri move to the ice in the winter months at a time when the algal production in the ice exceeds that of the sea bottom or water column.  The timing of the ice-ward migration for the two species differed, with Tisbe migrating gradually from February to April at which time, a new generation was produced.  The migration of Cyclopina consisted of only young copepods moving to the ice in early winter and remaining there until April, emerging as mature adults. Despite phytoplankton blooms in the water column in early June, the two ice copepods descend to the bottom to take advantage of the organic material dropping from the melting ice and the water column (Grainger 1991).

The feeding habits of harpacticoid copepods is considered to be herbivorous with little selective feeding (Grainger and Hsiao 1990; Kramer 2011), though many species will supplement their diatom-based diet with detritus during periods of low productivity (Arndt and Swadling 2006).  Tisbe spp. has also been found with fish larvae and copepod eggs and copepodites (Grainger et al. 1985).  Cyclopoids appear to be more omnivorous, with gut contents and lipid biomarkers showing a diet of diatoms, copepod eggs, and detritus.

Cyclopoid copepods are known to use sea ice for reproduction and development.  A large portion of the Cyclopina population is comprised of ovigerous females and up to three cohorts (Arndt and Swadling, Kern and Carey 1983) with a generation cycle of 31 days.  While the full life history of harpacticoid copepods is less well known, Arndt and Swadling (2006) infer that the ice is used for reproduction and growth, given that the eggs, nauplii and copepodite stages, as well as gravid females are found in the ice.  Furthermore, harpacticoid copepods can have >10 broods per year and a generation cycle of 20 days (Tisbe furcata).

2.2.5.3 Ice Amphipods

The sympagic macroinvertebrate community is dominated by ice amphipods, in particular the pan-Arctic species Gammarus wilkitzkii, Apherusa glacialis, Onismus nanseni and O. glacialis (Hop et al. 2000; Melnikov 1997; Arndt and Swadling 2006).  These autochthonous amphipods reside primarily in sea ice, occupying the three dimensional structure under the ice, as well as somewhat limited use of the brine channels and structure created by the algal mats and strands.  Amphipod abundance is highly variable across the Arctic.  While all five of the dominant species are found in areas impacted with the perennial Arctic ice, Onismus spp. is more commonly associated with young, seasonal ice (Arndt and Swadling 2006).  The amphipod G. wilkitzkii is more abundant in multiyear ice, but will move from drifting multi-year ice to first year ice.

Allochthonous amphipods that are pelagic or benthic in origin are also found under the pack ice, generally taking advantage of the high spring production associated with the sympagic communities (Melnikov 1997).  Planktonic amphipods may include Parathemisto spp. which can be common at the ice-water interface and may occur in swarms of several hundred individuals per m2 (Dalpadao et al. 2001).  However, most allochthonous species are benthic in origin, occurring in the land fast ice, or seasonal ice forming over shallow coastal areas.  Species may include Anonyx nugax, Anonyx sarsi, and G. setosuswhich may occur in abundance of tens per m2 (Arndt and Swadling 2006)

Apherusa glacialis is an herbivorous amphipod and an important grazer of ice algae (Arndt et al. 2005).  This species concentrates along ice edges and beneath more translucent new ice, where the onset of primary production takes place (Hop et al. 2000).  The diet may shift towards detritus when algae become less abundant in winter.  In areas with high amphipod abundance, the ice algae biomass may decrease at a rate of 30% to 60% of the standing stock per day (Arndt and Swadling 2006).  Phytodetritus is a major food item for Onismus spp.; however, O nanseni is a species repeatedly collected by baited traps, scavenging on carcasses and live prey including sympagic harpacticoid and cyclopoid copepods.  Gammarus wilkitzkii is primarily a predator, catching copepods, chaetognaths and other live prey, as well as amorphous organic debris, diatoms and microflagellates.  Other Lysianassid amphipods that move from pelagic or benthic communities to the pack ice are typically generalists feeding on ice algae, detritus, copepods, and fish eggs. (Arndt and Swadling 2006)  These species can occur in large swarms and in turn become an important food source for Arctic and Polar cod. 

Sympagic amphipods have perennial life cycles and are thought to spawn once a year. Apherusa glacialis can reach 2 years in age but has a high fecundity (>500 eggs develop for 6 months in a female’s brood pouch).  Offspring are released between March and August, presumably to take advantage of the spring growth of ice algae (Arndt and Swadling 2006).  The life span for Onisimus spp. is 2.5 to 3.5 years (Hop et al. 2000).  The reproductive cycles for the two species are offset, with O. glacialis spawning a few months earlier than O. nanseni.  Both species have one brood per year and produce approximately 100 eggs over the female’s life span. G. wilkitzkii is the longest lived ice amphipod, having a 5 to 6 year life span.  This species matures after two years and has one brood per year, releasing 90 to 250 eggs per year.   Eggs are deposited into the females marsupium during winter and release them in April and May when primary production peaks in ice-filled waters.  However, they can be flexible as the amphipods (such as G. wilkitzkii) are generalists and can release young April to September.

2.2.5.4 Pelagic Copepods

The pelagic calanoid copepods common in the Arctic are commonly found under the pack ice, particularly during the spring bloom (Melnikov 1997; Arndt and Swadling 2006).  While they do not appear to colonize the ice, the eggs and nauplii may be advected to the ice sheet.  Calanus glacialisC. hyperboreusPseudocalanus acuspesMetridia longa, and Oithona similis perform diel vertical migrations to the ice surface at dusk (Fortier et al. 2001) to feed at the ice-water boundary.  Biomass estimates for pelagic copepods are highly variable and are typically highest during the spring bloom, particularly during the ice melt (Arndt and Swadling 2006).  There are some indications that during this time of both ice algae and phytoplankton blooms, calanoid copepod build up a significant portion of their lipid reserves.  The ice-water interface can be an important nursery ground for Calanoid eggs and nauplii, providing food and shelter.  Other omnivorous copepods, such as M. longa come to the sea-ice to feed on Calanus eggs and the small ice crustacean, as well as sympagic diatoms.

2.2.5.5 Sympagic Fish

Both B. saida and A. glacialis spend a portion of their life history in close association with the sea-ice, utilizing cavities and ledges on the underside of the pack ice, as well as the edges of melting ice floes (Lonne and Guilliksen 1983; Gradinger and Bluhm 2005).  The larval and juvenile life stages are commonly found individually or in small groups, with adults are seldom found in close association with the ice.  In multiyear and first year ice, Lonne and Guilliksen (1983) found both one and two-year old fish living in and under the pack ice.  No fish older than 2-year-olds were observed. 

The diet of Arctic cod in multiyear ice is comprised of a mixture of sympagic amphipods as well as pelagic copepods (Lonne and Gulliksen 1989; Melnikov 1997).  In first-year ice, Lonne and Gulliksen (1989) found the Arctic cod diet dominated by calanoid copepods, however, this was likely due to the distribution and abundance of food items.  Arctic cod have been observed feeding on sympagic amphipods in first year ice in other regions of the Arctic.  Younger cod tended to favor harpacticoid and cyclopoid copepods, shifting to the larger prey with age (Bradstreet 1979).  The amphipod, G. wilkitzkii, was seldom found in Arctic cod stomach contents.  This is likely due to the large size and spiny morphology of the G. wilkitzkii.

2.2.5.6 Mammals

Ice amphipods and the more energy rich polar cod are subject to strong predation by top carnivores.  The sympagic macrofauna is a major link in the transfer of energy from sympagic primary producers to the ice-associated sea birds and marine mammals. 

Ice seals are perennial predators under the ice.  The seal diet is variable and is based on food availability.  Based on stomach contents and fecal analysis from seals in the high Canadian Arctic, Arctic cod comprised the majority of the adult ringed seal (P. hispeda) diet, with small proportions (approximately 8%) of amphipods and larger pelagic copepods (Bradstreet and Cross 1982).  Immature seals had a diet that was numerically dominated by ice amphipods, with approximately 3% cod.  However, the cod comprised 62% of the biomass in the stomach contents.  Both A. glacialis and B. saida were present in the stomach contents of young seals.  When Arctic cod are less common, a variety of crustacean species dominate the stomach contents, including ice-amphipods, mysids and other under-ice fauna (Melnikov 1997).

Other seals living amongst the ice include the leporine or bearded seal (E. barbatus), the harp seal (P. groenlandica), and hooded seal (C. cristata).  Erignatus barbatus is primarily a benthic feeder primarily on mollusks, crustaceans, and demersal fish including B. saida and A. glacialis (Melnikov 1997; Finley and Evans 1983).  Four taxa dominate the hooded and harp seal diets in Greenland: Arctic cod, capelin, the squid Gonatus fabricii and the pelagic amphipod, Parathemisto sp. (Haug et al. 2004).  Based on the stomach contents and satellite tracking data, the feeding habits of these two species is more pelagic in nature and not necessarily associated with the communities under the ice.

2.2.5.7 Birds

Sea birds are a primary consumer of sympagic fauna while their access is limited to areas with open water along ice edges.  Important avian predators include northern fulmars, black legged kittiwakes (Rissa tridactyla), guillemots (Uria lombia and Ceppus grille), murres, thick-billed murres, Little auks (Alle alle), and gulls.

Kittiwakes, Black guillemots (C. grille) and Brunnichs guillemots (U. lombia) and murres feed mainly on polar cod.  Other guillemots feed on fammasur in the Canadian Arctic and Themisto spp when the water is ice free.  In the Barents Sea, Little auks forage in multiyear ice mainly on A. glacialis (makes up 80% of their diet).  In the Canadian and Norwegian Arctic, the diet of Little auks is primarily C. glacialis (Falk-Petersen et al. 2007). 

2.2.6 VECs of Arctic Marine Environments

The key VECS for four Arctic realms are presented in Table 2-1.  In many cases, the species overlap between different realms.  For example Calanoid copepods, Arctic cod, and amphipods play a key role throughout Arctic food webs.  Additionally, interface habitats that contribute to the functioning and diversity of the Arctic should be further studied. These habitats include: surface microlayer, ice edges and ice margins, under-ice flora and fauna, water mass convergence zones, demersal communities and shorelines.

2.2.6.1 Seasonal Distribution Patterns of Arctic Marine Populations

Determining areas of seasonal population aggregations of VECs is important to inform NEBA decision making. However, these efforts require information on life-history and presence/absence data for each VEC.  Such analyses have been done for certain portions of the Arctic including the US and Canadian Beaufort.    An example of such an effort is the Beaufort Regional Environmental Assessment (BREA) program in the eastern Beaufort.  Seasonal movements of VEC species have been overlain with traditional hunting grounds and other data to create VEC vulnerability profiles that indicate specific locations and time periods where the population may be sensitive to certain OSRs (Trudel 2013).  As an example, data has been summarized for White whales (D. leucas) as they enter the eastern Beaufort in early June, with adults and young congregating at the mouth of the Mackenzie River delta (Figure 2-10).  During July, population densities are highest close to the mouth of the river, in areas used by indigenous fisheries.  In September and October, all whales leave the area.  Based on the vulnerability analysis, the White whales would be most vulnerable close to the mouth of the river during July. This approach has been used in different portions of the Arctic; a pan-Arctic compilation of such data would be useful to OSR strategizing.

 

Table 2-2.  Valuable Ecosystem Components of Arctic Communities

Valuable Ecosystem ComponentsAssociated Communities
PelagicBenthicSea-IceDeepwater

Phytoplankton

 

 

Sympagic copepods

Gammarus wilkitzkii

Apherusa glacialis

Onismus spp.

   




 

Calanoid copepods

Calanus hyperboreus

Calanus glacialis

Calanus finmarchicus




 






Euphausiids

Thysanoessa spp


 


 

Hyperiid amphipods

Themisto libellula


 


 

Cephalopods

Gonatus fabricii



 


Pelagic Fish

Arctic cod - Boreogadus saida

Polar cod - Arctogadus glacialis

Capelin - Mallotus villosus

Myctophids

Gonostomids
















Clams

Serripes groenlandica

Macoma sp.




   

Benthic and Epibenthic Amphipods

Ampelisca sp.

Anonyx nugax

Eurythenes gryllus

 




 




Decapod crustaceans

Shrimp -  Pandalus borealis

Crab -  Chionoecetes spp.




   

Echinoderm

Urchin - Strongylocentrotus droebachiensis

 


   

Epibenthic Fish

Sculpin – Myoxocephalus spp.

Eelpout – Lycodes spp.

Greenland halibut - Reinhardtius sp.







 




Mammals

Ringed seal – Phoca hispeda

Walrus - Odobenus rosmarus

Narwhal – Monodon monoceros

White whale – Delphinapterus leucas

Bowhead whale – Balaena mysticetus

Polar bear – Ursus maritimus
















 

Seabirds

 

 

● Integral component of the community

℗ Prey item or predator, but not an integral component of the community

Figure 2-10. Distribution of adult and young White whales (D. leucas)
in the eastern Beaufort Sea. [Shaded areas represent whale densities of 1 ind (blue), 5 (green), and 50 (red); Source: Trudel 2013]
Figure 2-10. Distribution of adult and young White whales (D. leucas) in the eastern Beaufort Sea. [Shaded areas represent whale densities of 1 ind (blue), 5 (green), and 50 (red); Source: Trudel 2013]

2.3 Future Research Considerations

There has been a dramatic increase in data available on various aspects of marine Arctic communities over the last decade, particularly on the pan-Arctic distribution of species and trophic interactions.  New studies include investigations of the trophic links within different systems (e.g. the central role of Arctic and Polar cod and Calanoid copepods in the pelagic food webs).  The review of Arctic ecosystems and VECs described by the authors in this section led to suggestions of further research which can reduce remaining uncertainties.  The more generic recommendations for further research compiled from this review are summarized below while recommendations that are important for improving Arctic NEBA are listed separately.

  1. Continue studies on Arctic faunal groups.  Some Arctic populations are now well understood in terms of natural history and toxicological profiles; however, groups of species require further examination regarding trophic roles, distributions, abundances, and ecotoxicological profiles based on annual and Interannual patterns.
    1. Knowledge of the interrelationships of Arctic species in areas of high productivity could benefit from further attention, especially with respect to migration and emigration through river systems, lagoons, and polynyas.
    2. The distributions and abundances of benthic and bathypelagic communities of the Arctic are not well known.  Boulder patch and other isolated areas of hard substrate as well as lagoon systems have proven to be important areas of increased production in the Arctic, but have received little attention.  While there have been some studies evaluating the deep sea communities particularly in the Norwegian Deep, eastern Beaufort, and parts of Baffin Bay, there are substantial portions of the Arctic deep water that have not been assessed.  This is a difficult area to characterize; however, recent studies in the High Canadian Arctic (Geoffroy et al. 2011, 2013; Reist and Majewski 2013) indicate that during certain portions of the year, Arctic cod can be found in large numbers in deeper waters of the Arctic.  There are indications that VECs known to occupy the deep sea habitats in other oceanic basins are present in the Arctic (e.g. myctophids, deep-sea corals), however, these communities are not likely to be disrupted by near-term oil and gas activities.  Deep water assessments would become important if there were a deep water release from a drilling platform.
  2. Continue pan-Arctic data collation.  Data from holistic efforts, such as BREA and RUSALCA could be collated and put into a GIS platform.  VEC species or regions for which there is not sufficient data may require additional data collection efforts.  These types of efforts generally occur as new areas are open for exploration or development.
  3. Evaluate ecosystem services.  Ecosystem services are the conditions and processes through which natural ecosystems and the species that comprise them sustain and fulfill human needs (Daily 1997).   Marine ecosystem services include functions that support human life, such as the production of ecosystem goods (e.g. seafood) and cleansing and sequestering wastes (e.g. uptake of excess nutrients by phytoplankton).  The marine ecosystem confers intangible aesthetic and cultural benefits (Kaufman and Dayton 1997; Peterson and Lubchenco 1997) to residents of the Arctic.

2.3.1 Priority Recommendations to Enhance NEBA Applications in the Arctic

There are only a handful of studies useful to understanding the trophic interactions of emerging habitats of concern (e.g. interface habitats and deep-sea basins of the Arctic).  The recommendations presented below indicate where increased knowledge of Arctic ECs and VECs would result in reducing existing uncertainties in NEBA assessments. No prioritization has been made to the list; for some of the recommendations, surrogate data may be already available.

  1. Expand knowledge of Arctic ECs.  Assessment of Arctic ECs should be expanded to include the communities populating interface habitats. These habitats include: sea surface microlayer, ice edges and margins, under-ice flora and fauna, water mass convergence zones, demersal communities, and shorelines.  These specialized habitats and resident or transient species may contribute significantly to the overall functioning, diversity, and resilience of the Arctic.  While the effects caused by individual OSR actions to key VECs living in the open water pelagic environment has been examined, repercussions of oil exposure to aggregating communities within convergent interface habitats is less well understood. 
    1. The surface microlayer (SML).  The surface microlayer refers to the uppermost surface layer(s) of the ocean.  The depth of the layer(s) is defined differently by physical oceanographers, chemists and biologists based on their conceptual model developed to address their different fields of interest.  Physical oceanographers and atmospheric scientists view the layer as the interface between the air and water while chemical oceanographers describe the layer based on the behavior of hydrophilic and hydrophobic moieties of chemical compounds.  Biological oceanographers define these layers based on where organisms and life stages reside or interact with the sea surface.  Certain communities of plants, invertebrates and vertebrates spend all of their life history at the sea surface and these are typically referred to as neuston.  The SML also acts as a nursery ground for many larval fish and invertebrate species, including larval species settling onto intertidal surfaces.  This group of surface oriented species represents a community of organisms that is most closely associated with surfaced oil as the oil sheen spreads over the water’s surface.  Whether the oil sheen is only a few microns or centimeters thick the organisms that contact this layer are exposed to the highest oil concentrations with the potential of activating multiple modes of toxic action.  In some cases larger marine organisms can skim feed on the concentrated masses of food and contaminants (certain fish, birds, and mammals) while others swim through the layer(s) to breath.  An understanding of the role of the neuston in the pelagic and intertidal food webs is needed to better characterize the impacts of surfaced, untreated oil and potential recovery rate of this vital micro-compartment.  Exposure to oil at the upper sea surface layer may result in additional toxic stress via different modes of toxic action, including fouling and respiratory stress from evaporating volatile compounds.
    2. Ice edges and margins, polynyas and other interface communities.  Polynyas have been identified as areas of enriched abundance and production during the Arctic winter.  Pelagic-benthic coupling is also showing that the increased pelagic activity is mirrored in the benthic community.  These different communities are typically an aggregation of species already known to be important in other portions of the Arctic.  Identifying and further studying the importance of these areas is of importance for the selection of OSR alternatives.
  2. Emphasize functional role of faunal groups.  The list of VEC species to be included in NEBA analyses is not static for all areas of the Arctic.  Emphasis will be placed on functional roles while addressing regional differences. New information regarding trophic food webs, population abundances and distribution patterns as well as toxicological profiles of VECs should be continuously expanded and updated (e.g. for ophiuroids, hard corals, jellyfish and neuston).  Population size estimates of VECs that occupy interface habitats compared to bulk pelagic waters is needed to determine the relative impact of the various OSR options.
  3. Increase understanding of resiliency and potential for recovery of Arctic species and populations.  An evaluation of the resiliency of potentially impacted populations of VEC species within Arctic ECs is critical in determining the ultimate biological consequences of each oil spill response considered during emergency oil spill response planning. Generic metrics for resilience should be developed and scored for keystone VECs. Refer to Sections 7, 8 and 9 for further concept development (Population Effects Modeling, Ecosystem Recovery, and NEBA for Oil Spill Response Options in the Arctic, respectively).

2.4 Further Information

Authors William Gardiner and Dr. Jack Q Word (ENVIRON), Ida Breathe Øverjordet (SINTEF), Dr. Oleg Titov and Andrei Zhilin (PINRO), Dr. Thierry Baussant (IRIS) 

References

Aitkenv AE, Fournier J

1993

Macrobenthos communities of Cambridge, McBeth and Itirbilung Fiords, Baffin Island, Northwest Territories, Canada

Arctic 46(1):60-71

Allen JA

1979

The adaptations and radiation of deep-sea bivalves

Sarsia. 64:19-27

Anisimova NA

1989

Distributional patterns of echinoderms in the Eurasian sector of the Arctic Ocean

In: The Arctic Seas. Climatology, Oceanography, Geology, and Biology (ed. By Yvonne Herman). N-Y:Van Nostrand Reinhold Compa

Anisimova NA , IE Manushin

2003

Feeding of the red king crab in the Barents Sea.

In: The Red King Crab in the Barents Sea, 2nd edn, pp. 170–189. Ed. by BI Berenboim, VA Borovkov, SS Drobysheva, GI Nesvetova,

Arndt CE, B Gulliksen, OJ Lønne, J Berge

2009

Sea ice fauna

In: E Sakshaug, G Johnsen, K Kovacs (eds) Ecosystem Barents Sea, Chap 13. Academic Press, Tapir.

Arndt CE, KM Swadling

2006

Crustacea in Arctic and Antarctic sea ice: Distribution, diet and life history strategies.

Adv. Mar. Ecol. 51:144-320

Arndt CE, O Pavlova

2005

Origin and fate of ice fauna in the Fram Strait and Svalbard area

Mar Ecol Prog Ser 301:55-66

Ashjian C, S Smith, F Bignami, T Hopkins, P Lane

1997

Distribution of zooplankton in the Northeast Water Polynya during summer 1992

J Mar. Syst. 10: 279-298.

Auel H, M Harjes, R daRocha, D Stubing, W Hagen

2002

Lipid biomarkers indicate different ecological niches and trophic relationships of the Arctic hyperiid amphipods Themisto abyssorum and T. libellula

Polar Biol. 25(5): 374-383

Barber DG, Saczuk E, Richard PR

2001

Examination of Beluga-Habitat Relationships through the Use of Telemetry and a Geographic Information System

Arctic 54: 305–316

Barnes DKA

1999

The influence of ice on polar benthos

J Mar Biol Assoc UK 79:401–407

Benoit D, Y Simard, J Gagne, M Geoffrey, L Fortier L

2010

From polar night to midnight sun: photoperiod, seal predation, and the diel vertical migrations of polar cod (Boreogadus saida) under landfast ice in the Arctic Ocean

Polar Biol 33:1505–1520

Benoit-Bird KJ, Au WWL

2006

Extreme diel horizontal migrations by a tropical nearshore resident micronekton community

Marine Ecology Progress Series 319:1-14

Blanchard, JL, JK Pinnegar and S Mackinson

2002

Exploring marine mammal-fishery interactions using “Ecopath with Ecosim”: Modelling the Barents Sea ecosystem

Sci Ser Tech Rep, CEFAS Lowestoft 117: 52 pp.

Bluhm B, D Watts, F Huettmann

2010

Free database availability, metadata and the internet: an example of two high latitude components of the census of marine life.

Chapter 13. In: Spatial complexity, informatics and wildlife conservation. Eds. S Cushman and F Huettmann. Springer, Japan, pp 2

Bluhm BA, Gebruk AV, Gradinger R, Hopcroft RR, Huettmann F, Kosobokova KN, Sirenko BI, Weslawski JM

2011

Arctic marine biodiversity: An update of species richness and examples of biodiversity change

Oceanography 24(3):232-248

Bluhm BA, Gradinger R

2008

Regional variability in food availability for Arctic marine mammals

Ecological Applications 18(2) Supplement:S77-S96

Bluhm BA, K Iken, SM Hardy, BI Sirenko, BA Holladay

2009

Community structure of epibenthic megafauna in the Chukchi Sea

Aquatic Biology 7: 269-293

Bodil BA, Ambrose Jr WG, Bergmann M, Clough LM, Gebruk AV, Hademann C, Iken K, Klages M, MacDonald IR, Renaud

2011

Diversity of the arctic deep-sea benthos

Marine Biodiversity 41:87-107

Bouchard C, L Fortier

2011

Circum-arctic comparison of the hatching season of polar cod Boreogadus saida: a test of the freshwater winter refuge hypothesis.

Prog Oceanogr 90:105–116

Bradstreet, MSW

1976

Summer feeding ecology of seabirds in eastern Lancaster Sound

Report by LGL Ltd, Toronto Canada for Norlands Petroleum Ltd. 187 p.

Bradstreet, MSW

1979

Feeding ecology of seabirds in northwest Baffin Bay, 1978

Report by LGL Ltd, Toronto Canada for Petro-Canada Corporation. 65 p.

Breines R, A Usvik, M Nymark, SD Johansen, DH Coucheron

2008

Complete mitochondrial genome sequences of the Arctic Ocean codfishes Arctogadus glacialis and Boreogadus saida reveal oriL and tRNA gene duplications.

Polar Biol. 31:1245-1252

Brinton E

1962

The distribution of Pacific euphausiids.

Bulletin of the Scripps Institution of Oceanography. 8(2):51-269.

Carmack E, Barber D, Christensen J, MacDonald R, Rudels B, Sakshaug E

2006

Climate variability and physical forcing of the food webs and the carbon budget on panarctic shelves

Prog Oceanogr 71(2-4):145-181

Cartes JE

1993

Diets of two deep-sea decapods: Nematocarcinus exilis (caridea: nematocarcinidae) and Munida tenuimana (anomura: galatheidae) on the western Mediterranean slope.

Ophelia. 37(3):213-229

Cartes JE, Maynou F, Sarda F, Company JB, Lloris D, and S Tudela

2004

The Mediterranean deep-sea ecosystems: an overview of their diversity, structure, functioning and anthropogenic impacts.

In: The Mediterranean deep-sea ecosystems: an overview of their diversity, structure, functioning and anthropogenic impacts, wit

Cochrane, SKJ, SF Denisenko, PE Renaud, CS Emblow, WG Ambrose Jr, IH Ellingsen, and J Skardhamar

2009

Benthic marcofauna and productivity regines in the Barents Sea – Ecological implications in a changing Arctic.

J Sea Res 61:222-233.

COSEWIC

2009

COSEWIC assessment and update status report on the Bowhead Whale Balaena mysticetus, Bering-Chukchi-Beaufort population and Eastern Canada-West Greenland population

Committee on the Status of Endangered Wildlife in Canada. Ottawa. vii + 49 pp.

Coupel P, HY Jin, M Joo, R Horner, HA Bouvet, M-A Sicre, J-C Gascard1, JF Chen, V Garcon, D Ruiz-Pino

2012

Phytoplankton distribution in unusually low sea ice cover over the Pacific Arctic

Biogeosciences 9:4835–4850

Craig PC

1984

Fish use of coastal waters of the Alaskan Beaufort Sea: A Review

Trans Amer Fish Soc 113(3): 265-282.

Dahl TM, S Falk-Petersen, GW Gabrielsen, JR Sargent, H Hop, RM Millar

2003

Lipids and stable isotopes in common eider, black legged kittwake and north fulmar: a trophic study from an Arctic fjord

Mar Ecol Prog Ser 256:257-269

Daily GC

1997

Nature’s Services: Societal Dependence on Natural Ecosystems.

Island Press

Dalpadado P, Yamaguchi A, Ellertsen B, Johannessen S

2008

Trophic interactions of macro-zooplankton (krill and amphipods) in the Marginal Ice Zone of the Barents Sea.

Deep-Sea Res II 55:2266–2274

Darnis G, Robert D, Pomerleau C, Link H, Archambault P, Nelson RJ, Geoffroy M, Tremblay JE, Lovejoy C, Ferguson SH,

2012

Current state and trends in Canadian Arctic marine ecosystems: II. Heterotrophic food web, pelagic-benthic coupling, and biodiversity

Climate Change 115:179-205

Day RH, AR DeGrange, GV Divoky, DM Troy

1988

Distribution and subspecies of the Dovekie in Alaska

Condor 90:712–714

Dehn LA, Follmann EH, Thomas DL, Sheffield GG, Rosa C, Duffy LK, OHara TM

2006

Trophic relationships in an Arctic food web and implications for trace metal transfer

Science of the Total Environment 362:103-123

Denisenko SG

2007

Zoobenthos of the Barents Sea in a changing climate and human impact// Dynamics of marine ecosystems and modern problems of conservation of biological resources of the Russian Seas

Vladivostok, Dalnauka. P. 418 – 511. (In Russian.)

Dewey T

2009

Fulmarus glacialis (On-line)

Animal Diversity Web. Accessed February 12, 2013 at http://animaldiversity.ummz.umich.edu/accounts/Fulmarus_glacialis/

Dolgov AV, OV Smirnov, KV Drevetnyak, OY Chetyrkina

2009

New data on composition and structure of the Kara Sea ichthyofauna.

ICES CM 2009/E:32

Dunton KH, Reimnitz E, Schonberg SV

1982

An Arctic kelp community in the Alaskan Beaufort Sea

Arctic 35(4):465-484

Dunton KH, SV Schonberg, LW Cooper

2012

Food web structure of the Alaskan nearshore shelf and estuarine lagoons of the Beaufort Sea.

Estuaries and Coasts 35:416-435

Engelhardt FR

1982

Hydrocarbon metabolism and cortisol balance in oil-exposed ringed seas, Phoca hispida

Aquat Toxicol 4(3):199-217

Ershova E, Hopcroft R, Kosobokova K

2012

A multi-year census of the zooplankton communities in the Pacific Arctic

Falk-Petersen S, V Pavlov, S Timofeev, JR Sargent

2005

Climate variability and possible effects on arctic food chains: The role of Calanus.

In: Arctic alpine ecosystems and people in a changing environment. (eds JB Orbaek, R Kallenborn, I Tombre, EN Hegseth, S Falk-Pet

Feder HM, D Schamel

1976

Shallow water benthic fauna of Prudhoe Bay

Assessment of the Arctic Marine Environment: Selected Topics, Chapter 22. Institute of Marine Science, University of Alaska, Fairba

Filatova ZA

1957

Overview of bivalve fauna of the northern seas of the USSR

Processing of Institution Of Oceanology of AS USSR, 20: 3-59. (In Russian.)

Filatova ZA, Zenkevich LA

1957

The quantitative distribution of the benthic fauna of the Kara Sea

Processing of Hydrobiology Society of USSR, 8: 3-67. (In Russian.)

Finley KJ, Evans CR

1983

Summer ‘Diet of the Bearded Seal (Erignuthus burbutus) in the Canadian High Arctic

Arctic 36(1):82-89

Fletcher GL, RF Addison, D Slaughter, CL Hew

1982

Antifreeze proteins in the Arctic Shorthorn Sculpin (Myoxocephalus scorpius)

Arctic 35(2): 302-306

Fortier M, Fortier L, Michel C, Legendre L

2002

Climatic and biological forcing of the vertical flux of biogenic particles under seasonal Arctic sea ice

Mar Ecol Prog Ser 225:1-16.

Fredriksen S

2003

Food web studies in a Norwegian kelp forest based on stable isotope (δ13C and δ15N) analysis

Marine Ecology Progress Series 260:71-81

Frost KJ, LF Lowry

1984

Trophic Relationships of Vertebrate Consumers in the Alaskan Beaufort Sea.

The Alaskan Beaufort Sea, Ecosystems and Environments. PW Barnes et al. ( eds). Academic Press, pp 381-401.

Fujiwara, A, T Hirawake, K Suzuki, I Imai, and SI Saitoh

2014

Timing of sea-ice retreat can alter phytoplankton community structure in the western Arctic Ocean

Biogeosciences. 11(7):1705-1716

Gage JD, PA Tyler

1991

Deep-sea biology: a natural history of organisms at the deep-sea floor.

Cambridge University Press, Cambridge

Gardiner K, Dick TA

2010

Arctic cephalopod distributions and their associated predators

Polar Research 29:209-227

Gebruk A

2004

Arctic marine fauna: Data accumulated in Russia

Census of Marine Life (COML)

Geoffrey MA, S Majewski, W Rousseau, J Waslkusz, J Reist, L Fortier

2013

Active acoustics mapping of fish: an answer to the mystery of the missing Arctic cod?

Presentation at the BREA 2013 Results Forum. Beaufort Regional Environmental Assessment, Gatineau, Quebec City, Canada.

Geoffroy M, A Majewski, S Gauthier, M LeBlanc, W Walkusz, JD Reist, and L Fortier

2014

Target Strength, vertical distribution and migrations of pelagic fish in the Canadian Beaufort Sea from spring to fall,

Polar Biol, submitted

Geoffroy M, A Majewski, S Rousseau, W Walkusz, J Reist, L Fortier

2013

Active Acoustic Mapping of Fish.

Presentation at BREA Results Forum February 19-21, 2013.

Geoffroy M, D Robert, G Darnis, L Fortier

2011

The aggregation of polar cod (Boreogadus saida) in the deep Atlantic layer of ice-covered Amundsen Gulf (Beaufort Sea) in winter

Polar Biol 34(12):1959-1971

Geoffroy M, Majewski A, Rousseau W, Walkusz W, Reist J, Fortier L

2013

An answer to the mystery of the missing polar cod (Boreogadus saida)?

2013 Arctic Frontiers Annual Conference

Gerasimova O, M Kochanov

1997

Trophic relations of the red king crab (Paralithodes camtschaticus) in the Barents Sea.

Researches on invertebrates in the Barents Sea. Selected papers. PINRO. Murmansk. PINRO Press: 35-58

Gradinger R

2008

Sea ice algae: Major contributors to primary production and algal biomass in the Chukchi and Beaufort sea during May/June 2002

Deap-Sea Research II: Topical Studies in Oceanography 56(17):1201-1212

Gradinger R, BA Bluhm

2005

Susceptibility of sea ice biota to disturbances in the shallow Beaufort Sea: Phase 1: Biological coupling of sea ice with the pelagic and benthic realms

Final report, OCS Study MMS 2005-062.

Gradinger R, Bluhm BA, Hopcroft RR, Gebruk AV, Kosobokova K, Sirenko B

2010

Marine life in the arctic

Life in the world's oceans: Diversity, distribution, and abundance. Wiley-Blackwell, 354pp.

Gradinger R, C Friedrich, M Spindler

1999

Abundance, biomass and composition of the sea ice biota of the Greenland Sea pack ice

Deep-Sea Res 46:1457-1472.

Gradinger RR, BA Bluhm

2004

In-situ observations on the distribution and behavior of amphipods and Arctic cod (Boreogadus saida) under the sea ice of the High Arctic Canada Basin

Polar Biol 27:595-603.

Gradinger RR, K Meiners, G Plumley, Q Zhang, BA Bluhm

2005

Abundance and composition of the sea-ice meiofauna in off-shore pack ice of the Beaufort Gyre in summer 2002 and 2003

Polar Biol., 28, 171-181

Graham M and H Hop

1995

Aspects of Reproduction and Larval Biology of Arctic Cod (Boreogadus saida)

Arctic 48(2):130-135

Grainger EH

1991

Exploitation of arctic sea ice by epibenthic copepods

Mar Ecol Prog Ser 77:119-124

Grainger EH, AA Mohammed, JE Lovrity

1985

The sea ice fauna of Frobisher Bay, Arctic Canada

Arctic 38:23-30

Grainger EH, SIC Hsiao

1990

Trophic relationships of the sea ice meiofauna in Frobisher Bay, Arctic Canada

Polar Biol 10:283-292

Grebmeier JM

2012

Shifting patterns of life in the Pacific Arctic and Sub-Arctic Seas

Annual Reviews of Marine Science 4:63-78

Grebmeier JM, Cooper L

2012

Benthic population dynamics within the RUSALCA program

Russian-American Long-Term Census of the Arctic (RUSALCA) 2012 PI Meeting

Griffiths W, DH Thomson

2002

Species Composition, Biomass, and Local Distribution of Zooplankton Relative to Water Masses in the Eastern Alaskan Beaufort Sea.

In: Bowhead Whale Feeding in the Eastern Alaskan Beaufort Sea. Richardson WJ, and Thomson, LGL Report TA2196-7; OCS study

Hamre J

1994

Biodiversity and exploitation of the main fish stocks in the Norwegian-Barents Sea ecosystem

Biodiversity and Conservation 3:473-492

Hardy JT

1982

The sea surface microlayer: Biology, chemistry, and anthropogenic enrichment

Proc. Oceanogr 11: 307-328.

Haug T, Nilssen KT, Lindblom L

2004

Feeding habits of harp and hooded seals in drift ice waters along the east coast of Greenland in summer and winter

Polar Research 23(1):35-42

Hay KA, AW Mansfield

1989

Narwhal Monodon monoceros Linneaus, 1758

In: S H Ridgway and R Harrison (eds), Handbook of marine mammals, pp. 145-176. Academic Press, London, UK

Hjermann DO, A Mersom, GE Dingsor, JM Durant, AM Eikeset, LP Reed, G Ottersen, G Storvik, NC Stenseth

2002

Fish and oil in the Lofoten–Barents Sea system: synoptic review of the effect of oil spills on fish populations

Mar Ecol Prog Ser 339:283-299

Hobson KA

1993

Trophic relationships among high Arctic seabirds: Insights from tissue-dependent stable-isotope models

Marine Ecology Progress Series 95:7-18

Hoekstra PF, LA Dehn JC George, KR Solomon, DCG Muir, TM Ohara

2002

Trophic Ecology of Bowhead Whales (Balaena mysticetus) compared with that of other Arctic marine biota as interprested from Carbon, Nitrogen, and Sulfur-Isotope signatures.

Can J Zool 80(2):223-231

Hop H, HW Welch HE, RE Crawford

1997

Population structure and feeding ecology of Arctic cod schools in the Canadian high Arctic.

In: Reynolds JB (ed.) Fish ecology in arctic North America. American Fisheries Society Symp 19, Bethesda, Maryland, pp 68-80.

Hop H, M Poltermann, OJ Lonne, S Falk-Petersen, R Korsnes, WP Budgell

2000

Ice amphipod distrubtion relative to ice density and under-ice topography in the northern Barents Sea.

Polar Biol 23:357-367.

Hopcroft R, Bluhm B, Gradinger R

2008

Arctic Ocean synthesis: Analysis of climate change impacts in the Chukchi and Beaufort Seas with strategies for future research

Arctic Ocean synthesis: Analysis of climate change impacts in the Chukchi and Beaufort Seas with strategies for future research

Hopcroft RR, Clarke C, Nelson RJ, Raskoff KA

2005

Zooplankton communities of the Arctic's Canada Basin: The contribution by smaller taxa

Polar Biology 28:198-206

Horner R

1984

Phytoplankton Abundance, Chlorophyll a, and Primary Productivity in the Western Beaufort Sea.

In: The Alaskan Beaufort Sea Ecosystems and Environments. Barnes PW (ed). Academic Press.

Horner R, Schrader GC

1982

Relative contributions of ice algae, phytoplankton, and benthic microalgae to primary production in nearshore regions of the Beaufort Sea

Arctic 35(4):485-503

Huettmann F, Y Arukhin, O Gilg and G Humphries

2011

Predictions of 27 Arctic pelagic sea bird distributions using public environmental variables, assessed with ecology data: a first digital IPY and GBOF open access synthesis platform.

Mar Biodiv 41:141-179.

Iken K, BA Bluhm, K Dunton

2010

Benthic food-web structure under differing water mass properties in the southern Chukchi Sea

Deep-Sea Res. II. 57:71-85

Iken K, Bluhm BA, Sirenko BI, Gagev S, Hardy SM, Holladay BA, Dunton K

2012

Temporal trends in epibenthic megafauna and food web structure in the Chukchi Sea

Russian-American Long-Term Census of the Arctic (RUSALCA) 2012 PI Meeting

IOC et al.

Arctic ocean bethymetry map

http://geology.com/world/arctic-ocean-bathymetry-map.shtml

Jarvela LE, Thorsteinson

1999

The epipelagic fish community of Beaufort Sea coastal waters, Alaska.

Arctic 52:80-94

Joensen J.

2008

Comparative feeding ecology of the sympatric cod fishes Arctogadus glacialis and Boreogadus saida in North-East Greenland evaluated from diet and stable isotope analyses

University of Tromso Masters Dissertation

Johnson L

1980

The Arctic char, Salvelinus alpines

In Chars: Salmonid fishes of the genus Salvelinus. Edited by EK Balon. Dr. W. Junk Publishers, The Hague, The Netherlands, p 15-9

Jørgensen OA, C Hvingel, PR Møller, MA Treble

2011

Identification and mapping of bottom fish assemblages in Davis Strait and Southern Baffin Bay

Can J Fish Aq Sci. 62(8):1833-1852

Jørgensen OA, Hvingel C, Møller PR, Treble MA

2005

Identification and mapping of bottom fish assemblages in Davis Strait and southern Baffin Bay.

Can J Fish Aquat Sci 62:1833–1852

Kaufman LS, PJ Dayton

1997

Impacts of marine resource extraction on ecosystem services and sustainability.

In: Daily, G. (Ed.), Nature's Services. Island Press. pp. 275-293

Kern J, Carey Jr AG

1983

The faunal assemblage inhabiting seasonal sea ice in the nearshore Arctic Ocean with emphasis on copepods

Marine Ecology Progress Series 10:159-167

Koski WR, GW Miller

2002

Habitat use by different size classes of bowhead whales in the central Beaufort Sea during late summer and autumn

In: Richardson, W.J and Thomson, D.H. (eds.) Bowhead whale feeding in the eastern Alaskan Beaufort Sea:

Kraft A, E Bauerfeind, EM Nothig, M Klages, A Beszczynska-Moller, and UV Bathmann

2012

Amphipods in sediment traps of the eastern Fram Strait with focus on the life-history of the lysianassoid Cyclocaris guilelmi.

Deep-sea Res. Part I: Oceanogr Res Papers:73:62-72

Kramer M

2011

The role of sympagic meiofauna in Arctic and Antarctic sea-ice food webs

Christian-Albrechts-Universität zu Kiel Doctoral Dissertation

Kramer, M, K Swadling, KM Meiners, R Kiko, A Scheltz, M Nicolaus, I Werner

2010

Antarctic sympagic meiofauna in winter: comparing diversity, abundance and biomass between perennially and seasonally icecovered regions.

Deep Sea Res. II: Topical Studies in Oceanography, 58(9-10):1062-1074

Krapp RH, J Berge, H Flores, B Gulliksen, I Werner

2008

Sympagic occurrence of Eusirid and Lysianassid amphipods under Antarctic pack ice.

Deep Sea Res. II. 55(2008):1015-1023

Kuklinski P, DKA Barnes

2008

Structure of intertidal and subtidal assemblages in Arctic vs temperate boulder shores

Pol Polar Res. 29(3):203-218

Laidre KL, MP Heide-Jorgensen

2005a

Winter feeding intensity of narwals (Monodon Monoceros).

Mar Mammal Sci 21(1):45-57

Laidre KL, MP Heide-Jorgensen, R Dietz

2002

Diving behaviour of narwhals (Monodon monoceros) at two coastal localities in the Canadian Arctic.

Can J Zool 80:624–635

Laidre KL, MP Heide-Jorgensen, R Dietz, RC Hobbs, OA Jorgensen

2003

Deep-diving by narwals Monodon monoceros: differences in foraging behavior between wintering areas?

Can J Zool 80, 624–635

Lantuit H, Overfuin PP, Couture N, Wetterich S, Are F, Atkinson D, Brown J, Cherkashov G, D Drozdov, Forbes DL, raves-

2012

The Arctic Coastal Dynamics Database: A New Classification Scheme and Statistics on Arctic Permafrost Coastlines

Estuaries and Coasts. 35(2): 383-400.

Letessier TB, MJ Cox, AS Brierly

2009

Drivers of euphaussid species abundance and numerical abundance in the Atlantic Ocean.

Mar. Biol. 156:2539-2553

Li WKW, McLaughlin FA, Lovejoy C, Carmack E

2009

Smallest algae thrive as the Arctic Ocean freshens

Science 326(5952):539-539

Lin L, Lian Y, Zhang J, Zheng S, Xiang P, Yu X, Wu R, Shao K

2012

Composition and distribution of fish species collected during the fourth Chinese National Arctic Research Expedition in 2010

Advances in Polar Science 23(2):116-127

Loseto LL, P Richard, GA Stern, J Orr, SH Ferguson

2006

Segregation of Beaufort Sea beluga whales during the open-water season

Can Jnl Zool 84 (12):1743–1751

Lowry LF, G Sheffield, JC George

2004

Bowhead Whale Feeding in the Alaska Beaufort Sea Based on Stomach Contents Analyses.

Cetacean Resources & Management 6(3): 215-224

Lowry LF, KJ Frost, JJ Burns

1980

Feeding of bearded seals in the Bering and Chukchi seas and trophic interaction with Pacific walruses.

Arctic 33:330-342

MacDonald IR, BA Bluhm, K Iken, S Gagaev, S Strong

2010

Benthic macrofauna and megafauna assemblages in the Arctic deep-sea Canada Basin.

Deep-Sea Research II. 57:136-152

Mackinson S, Pinnegar JK, Blanchard JL

2002

Exploring marine mammal-fishery interactions using the 'Ecopath with Ecosim': Modelling the Barents Sea ecosystem

CEFAS Loestoft 117:52pp.

Madsen ML, SE Fevolden, JS Christiansen

2009

A simple molecular model for Arctogadus and Boreogadus saida.

Polar Biol 32(6):937-939

Majewski A, J Reist

2013

Offshore fish populations, habitat and ecosystems. Presentation at the BREA 2013 Results Forum

Presentation at the BREA 2013 Results Forum. Beaufort Regional Environmental Assessment, Gatineau, Quebec City

Manushin I

2008

Rhythm of Red King Crab feeding activity

In: The Red King Crab in the Barents Sea, 2nd edn, pp. 20-28. Ed. by BI Berenboim, VA Borovkov, SS Drobysheva., GI Nesvetova, MS

Martin JW, TA Haney

2005

Decapod crustaceans from hydrothermal vents and cold seeps: a review through 2005.

Zool J Linnean Soc. 145: 445-522

Matyushkin VB

2003

Juveniles of the red king crab in the West Murman areas.

In: The Red King Crab in the Barents Sea, 2nd edn, pp. 140–152. Ed. by BI Berenboim, VA Borovkov, SS Drobysheva., GI Nesvetova,

McGuire A D, Anderson L G, Christensen T R, Dallimore S, Guo L, Hayes D J, Heimann M, Lorenson TD, Macdonald RW,

2009

Sensitivity of the carbon cycle in the Arctic to climate change

Ecological Monographs 79(4):523-555

Mecklenburg CW, Chernova NV

2012

Focal areas: Fish diversity and otter trawl

Russian-American Long-Term Census of the Arctic (RUSALCA) 2012 PI Meeting

Mecklenburg CW, TA Mecklenburg, LK Thorsteinson

2002

Fishes of Alaska

American Fisheries Society, Bethesda MD

Mecklenburg, CW and TA Mecklenburg

2011

Glacial eelpout: Lycodes frigidus

Arctic Ocean Diversity webpage: http://www.arcodiv.org/Fish/Lycodes_frigidus.html

Mehl S

1991

The Northeast Arctic cod stock's place in the Barents Sea ecosystem in the 1980s: An overview

Proceedings of the Pro Mare Symposium on Polar Marine Ecology, Trondheim, 12-16 May 1990

Mehl S, NA Yaragina

1992

Methods and results in the joint PINRO-IMR stomach sampling program.

In: B Bogstad and S Tjelmeland (eds.): Interrelations between fish populations in the Barents Sea. Proceedings of the fifth PINRO-I

Melnikov A

1997

Arctic Sea Ice Ecosystem

CRC Press

Moore SE

2000

Variability in cetacean distribution and habitat section in the Alaskan Arctic, autumn 1982-91.

Arctic 53(4): 448-460

NAMMCO (North Atlantic Marine Mammal Commission)

2005b

Status of Marine Mammals in the North Atlantic – The Atlantic Walrus..

NAMMCO, Tromsø, pp. 1-8.

NAMMCO (North Atlantic Marine Mammal Commission)

2005c

Status of Marine Mammals in the North Atlantic – The Atlantic Walrus ©

NAMMCO, Tromsø, pp. 1-7.

NAMMCO (North Atlantic Marine Mammal Commission)

2005a

Status of Marine Mammals in the North Atlantic – The Atlantic Walrus

NAMMCO, Tromsø, pp. 1-12.

Navarro J, M Coll, C Somes, RJ Olsen

2013

Trophic niche of squids: Insights from isotopic data in marine in systems worldwide.

Deep Sea Res II: Topical Studies in Oceanogr. In press

Niehoff B

2007

Life history strategies in zooplankton communities: The significance of female gonad morphology and maturation types for the reproductive biology of marine calanoid copepods.

Prog Oceanogr 74:1-47

Norcross BL, Holladay BA, Gleason C

2012

Fish ecology - Baseline for assessing effects of climate change on fishes

Russian-American Long-Term Census of the Arctic (RUSALCA) 2012 PI Meeting

Orlov YI, BG Ivanov

1978

On the introduction of the Kamchatka King Crab Paralithodes camschatica (Decapoda: Anomura: Lithodidae) into the Barents Sea

Mar Biol 48:373-375

Outridge PM, WJ Davis, REA Stewart, EW Born

2003

Investigation of the stock structure of Atlantic walrus (Odobenus rosmarus rosmarus) in Canada and Greenland using dental Pb isotopes derived from local geochemical environments

Arctic 56:82-90

Parker-Stetter SL, JK Horne, TJ Weingartner

2011

Distribution of polar cod and age-0 fish in the US Beaufort Sea

Polar Biol. 34:1543-1557

Percy JA, FJ Fife

1985

Summertime abundance and depth distribution of macrozooplankton in Fobisher Bay NWT from 1978 to 1983

Canadian Data Report of Fisheries and Aquatic Sciences No 503 63 p.

Peters J, Tuschling K, Brandt A

2004

Zooplankton in the arctic Laptev Sea—feeding ecology as indicated by fatty acid composition

Journal of Plankton Research 26(2):227-234

Peterson CH, J Lubchenco

1997

Marine ecosystem services

In G. Daily, editor. Natures’ services: societal dependence on natural ecosystems. Island Press, Washington D.C., USA. pp. 177–194

Petrova D, Dzhurinskiy

2012

Amphipods (Amphipoda: Gammaridae) of the Chukchi Sea

Russian-American Long-Term Census of the Arctic (RUSALCA) 2012 PI Meeting

Piepenburg D

2005

Recent research on Arctic benthos: common notions need to be revised.

Polar Biol 28:733–755

Piepenburg D, Archambault P, Ambrose W G, Blanchard A L, Bluhm B A, Carroll M L, Conlan KE et al.

2011

Towards a pan-Arctic inventory of the species diversity of the macro- and megabenthic fauna of the Arctic shelf seas

Marine Biodiversity 41:51-70

Piepenburg D, Schmid MK

1997

A photographic survey of the epibenthic megafauna of the Arctic Laptev Sea shelf: Distribution, abundance, and estimates of biomass and organic carbon demand

Marine Ecology Progress Series 147:63-75

Piepenburg D, WG Ambrose, A Brandt, PE Renaud, MJ Ahrens, P Jensen

1997

Benthic community patterns reflect water column processes in the Northeast Water polynya (Greenland).

J Mar Syst 10:467-482.

Pinchukov M, A Karsakov

2009

Red king crab setting and condition of habitation in Russian waters of the Barents Sea.

http://barentsportal.com/barentsportal_v2.5/index.php/en/barents-sea-status-report/introduction/report-references

Pinchukov MA, Pavlov VA

2002

Food composition and its spatial variability in red-king crab (Paralithodes camtschaticus Tilesius) in the Barents Sea.

In Abstracts from reports for the 4th Russian national conference on commercial invertebrates, pp. 59-61. PINRO Press. Moscow.

Poltermann M

2000

Growth, production and productivity of the Arctic sympagic amphipod Gammarus wilkitzkii

Mar Ecol Prog Ser 193:109-116

Poltermann M

2001

Arctic sea ice as feeding ground for ampipods - food sources and strategies

Polar Biol 24:89-96

Premke K

2003

Aggregations of Arctic deep-sea scavenging amphipod at large food falls.

Dissertation. University of Bremen. 134 pp.

Purcell JE, RR Hopcroft, KN Kosobokova, TE Whitledge

2009

Distribution, abundance, and predation effects of epipelagic ctenophores and jellyfish in the western Arctic Ocean.

Deep Sea Res. Topical Studies in Oceanogr. 57(1-2):127-135.

Raskoff KA, RR Hopcroft, KN Kosobokova, JE Purcell, M Youngbluth

2010b

Jellies under ice: ROV observations from the Arctic 2005 hidden ocean expedition

Deep-Sea Res II 57:111-126

Raskoff, K.A.

2010

Bathykorus bouilloni: A new genus and species of deep-sea jellyfish from the Arctic Ocean (Hydrozoa, Narcomedusae, Aeginidae

Zootaxa 2361:57-67

Reeves RR

1998

Distribution, abundance and biology of ringed seals (Phoca hispida): an overview

NAMMCO Sci. Publ. 1:9-45

Reeves RR, BS Stewart, PJ Clapham, PA Powell

2002

Marine Mammals of the World.

Knopf: New York, NY.

Reist J, A Majewski

2013

Offshore fish populations, habitat, and ecosystems.

Presentation at BREA Results Forum February 19-21, 2013.

Renaud PE, Włodarska-Kowalczuk M, Trannum H, Holte B, Weslawski JM, Cochrane S, Dahle S, Gulliksen B

2007

Multidecadal stability of benthic community structure in a high-Arctic glacial fjord (van Mijenfjord, Spitsbergen)

Polar Biology 30:295-305

Rice DW

1998

Marine mammals of the world: systematics and distribution

Wartzok D (ed); Lawrence, KS. USA: Society for Marine Mammalogy, Special Journal. p. 231

Roper CFE and RE Young

1975

Vertical distribution of pelagic cephalopods

Smithson Contrib Zool 209:1-51

Rugh D, D DeMaster, A Rooney, J Breiwick, K Shelden, S Moore

2003

A review of bowhead whale (Balaena mysticetus) stock identity.

J. Cet. Res. Manage. 5:267–279.

Sainte Marie B

1992

Foraging of scavenging deep-sea lysianassoid amphipods.

In: Deep-Sea Food Chains and the Global Carbon Cycle. NATO Science Series. Rowe GT and V Pariente (eds.). Springer Netherlands.

Sakshaug E

2004

Primary and secondary production in the Arctic Seas.

In: The organic carbon cycle in the Arctic Ocean (eds R Stein and R MacDonald). pp. 57-82. Berlin. Springer.

Sakshaug E, G Johnsen, S Kristianse, C von Quillfeldt, F Rey, D Slagstad, F Thingstad.

2009

Phytoplankton and primary production.

In: Ecosystem Barents Sea. Tapir Academic Publishers, Trondheim. pp. 167-208

Sakshaug E, G Johnsen, S Kristiansen, C vonQuillfeldt, F Rey, D Slagstad, F Thingstad

2009

Phytoplankton and primary production

IN: Shakshaug E, G Johnsen, K Kovacs (eds); Ecosystem Barent Sea. Tapir Academic Press, Trondheim pp. 167-208

Semiletov I, II Pipko, ME Shakhova, OV Dudarev

2011

On the biogeochemical signature of the Lena River from its Headwaters to the Arctic Ocean.

Biogeosci. Disc. 8:2093-2143

Sherman K and G Hempel

2008

The UNEP Large Marine Ecosystem Report: A perspective on changing conditions in LMEs of the worlds regional seas

UNEP Regional Seas Report and Studies No. 182. United Nations Environment Programme. Nairobi, Kenya

Sieburth J

1971

Distribution and activity of oceanic bacteria.

Deep Sea Resources 18: 1111-1121

Sirenko B, Piepenburg D, Spindler M

1994

Current knowledge on biodiversity and benthic zonation patterns of Eurasian Arctic Shelf seas with special reference to the Laptev Sea

Proceedings of the Russian-German Cooperation in and around the Laptev Sea, St. Petersburg, Russia, May 1993

Sirenko BI, C Clarke, RR Hopcroft, F Huettmann, BA Bluhm, R Gradinger (eds).

2010

The Arctic Register of Marine Species (ARMS) compiled by the Arctic Ocean Diversity (ArcOD) project.

Accessed at http://www.marinespecies.org/arms on 2013-01-24

Southward AJ, Sims DW

2006

Advances in Marine Biology

Elsevier, 326pp

Stein R, K Dittmers, K Fahl, M Kraus, J Matthiessen, F Niessen, M Pirrung, Ye Polyahova, F Schoster, T Steinke, DK Futter

2004

Arctic (palaeo) river discharge and environmental change: evidence from the Holocene Kara Sea sedimentary record.

Quaternary Science Reviews. 23(11-13):1485-1511.

Sufke L, Piepenburg D, von Dorrien CF

1998

Body size, sex ratio and diet composition of Arctogadus glacialis (Peters, 1874) (Pisces: Gadidae) in the Northeast Water Polynya (Greenland)

Polar Biology 20:357-363

Sundet JH, B Berenboim

2008

Research on Red King Crab (Paralithodes camtschaticus) from the Barents Sea in 2005-2007.

IMR-PINRO Joint Report Series 3/2008. 71 pp.

Suydam R, S Moore

2004

Changes in Arctic Sea Ice: What Are the Consequences for Whales.

North Slope Borough; Department of Wildlife Management Barrow AK. NOAA/Alaska Fisheries Science Center, NMML Seattle WA.

Suydam RS, Lowry LF, Frost KJ, O' Corry-Crowe GM, Pikok D JR

2001

Satellite Tracking of Eastern Chukchi Sea Beluga Whales into the Arctic Ocean

Arctic 54 (3): 237–243

Thiemann GW, Iverson SJ, Stirling I

2008

Polar Bear diets and Arctic marine food webs: Insights from fatty acid analysis

Ecological Monographs 78(4):591-613

Thompson DH

1982

Marine benthos in the Eastern Canadian High Arctic: Multivariate analyses of standing crop and community structure.

Arctic 35(1):61-74.

Thompson DH, CM Martin, WE Cross

1986

Identification and characterization of Arctic nearshore benthic habitats.

Can Tech Rep Fish Aqua Sci 1434.

Titov OV, Ozhigin VK, Gusev EV, Ivshin VA

2006

Theory of functioning of the Barents Sea ecosystem: Fisheries and oceanographic aspects

International Council for the Exploration of the Sea (ICES) 2006 Conference

Tremblay JE, D Robert, DI Varela, C Lovejoy, G Darnis, RJ Nelson, AR Sastri

2012

Current state and trends in Canadian Arctic marine ecosystems: I. Primary production.

Climatic Change 115(1):161-178

Trudel K

2013

Assemble biological data and traditional knowledge needed for net environmental benefit analysis for Beaufort Sea oil spill response planning.

BREA Forum. Inuvik Northwest Territories, February 21, 2012.

von Quillfeldt, C, EN Hegseth, G Johnsen, E Sakshaug, EE Syvertsen

2009

Ice algae.

In: Ecosystem Barents Sea. Tapir Academic Publishers, Trondheim. Pp 285-302

Welch HE, RE Crawford, H Hop

1993

Occurrence of Arctic cod (Boreogadus saida) schools and their vulnerability to predation in the Canadian high Arctic,

Arctic 46(4): 331–339

Werner I

1997

Grazing of Arctic under=ice ampipods on sea-ice algae

Mar Ecol Prog Ser

Weslawski JM

2011

Biodiversity of the Arctic Ocean in the face of climate change

Papers on Global Change 18:89-91

Weslawski JM, Kendall MA, Włodarska-Kowalczuk M, Iken K, Kedra M, Legezynska J, Sejr MK

2011

Climate change effects on Arctic fjord and coastal macrobenthic diversity - observations and predictions

Marine Biodiversity 41:71-85

Weslawski JM, KW Opalinski, J Legezynska

2000

Life cycle and production of Onisimus litoralis Crustacea, Amphipoda): the key species in the Arctic soft sediment littoral.

Pol Arch Hydrobiol 47:585–596.

Wienerroither RM, K Nedreaas, F Uiblein, JS Christiansen, I Byrkjedal, O Karamushko

2010

The marine fishes of Jan Mayen Island, Northeast Atlantic—past and present.

Mar Biodivers.

Word JQ, LS Word, JN McElroy, RM Thom

1986

The Surface Microlayer: Review of Literature and Evaluation of Potential Effects of Dredge Activities in Puget Sound.

PSDDA Report, Region 10.

Zaitsev YP

1971

Marine Neustonology.

Translated from Russian by Israel Program for Scientific Translation, Jerusalem.

Zeller D, S Booth, E Pakhomob, WE Swarz and D Pauly

2011

Arctic fisheries in Russia, USA, and Canada: baselines for neglected ecosystems

Arctic fisheries in Russia, USA, and Canada: baselines for neglected ecosystems

Petrich C, H Eicken

2010

Growth, structure, and properties of sea ice

In: Sea Ice, 2nd ed., Thomas, D.N. and G.S. Dieckmann (eds), Wiley-Blackwell.

Zinjarde SS, Pant AA

2002

Hydrocarbon degraders from tropical marine environments

Mar Pollut Bull 44: 118-121

Engelhardt FR

1983

Petroleum effects on marine mammals

Aquat Toxicol 4(3):199-217

Hsiao SI

1978

Effects of crude oils on the growth of arctic marine phytoplankton2

Environ Pollut 17(2): 93-107

Bradstreet MSW

1980

Thickbilled murres and black guiellemots in the Barrow Strait area, NWT during spring: diets and food availab ility along ice edges

Can J Zool 58:2120-2140

Bradstreet MSW and Cross WE

1982

Thickbilled murres and black guiellemots in the Barrow Strait area, NWT during spring: diets and food availab ility along ice edges

Can J Zool 58:2120-2140

Bradstreet MSW and Cross WE

1982

Trophic relationships of high-arctic ice edges

Arctic 35:1-12

Fortier M, Fortier L, Hattori H, Saito H, Legendre L.

2001

Visual predators and the diel vertical migration of copepods under Arctic sea ice during the midnight sun

Jnl Plankton Res 23(11):1263-1278

Lonne OJ, Gulliksen B

1989

Size, age and diet of Polar Cod, Boreogadus saida (Lepechin 1773), in ice covered waters

Polar Biology 9:187-191

Mecklenburg CW; Moller PR

2011

Biodiversity of arctic marine fishes: Taxonomy and zoogeography

Marine Biodiversity 41:109-140

3.0 THE TRANSPORT AND FATE OF OIL IN THE ARCTIC

Executive Summary

Photo 3-1. Marginal Ice (NOAA)
Photo 3-1. Marginal Ice (NOAA)

Oil is composed of many compounds resulting in mixtures containing a wide range of volatile, semi-volatile, soluble and more recalcitrant compounds. During a spill some of these oil components separate from the oil mixture by evaporation into the air or by solubilizing into the water leaving the more recalcitrant compounds in the weathered oil product where wind and wave action adds water to the mixture (emulsification). Spilled oil in the Artic undergoes the same processes occurring in environments throughout the world. However, the colder temperatures increase the viscosity of oil resulting in slower spreading rates and increased thickness of oil on the sea surface. The slower spreading rates also reduce the area covered by surface oil which may be further reduced by the presence of ice and snow. The thicker layers of oil on the sea surface decrease the surface area of the oil that is available for losses due to evaporation or solubilization and decrease the potential for emulsification especially in heavy ice with dampened wind and wave action. Furthermore, oil that is encapsulated in ice during the fall and winter provides an extended period for oil spill response actions (OSR) to proceed prior to and during spring melt. The most efficient biodegradation of oil occurs within the water column and, to a somewhat lesser degree, at the sea surface. Maximizing the surface area of the oil will increase the efficiency of microbes to reduce oil concentrations. Biodegradation of oil present in or on shorelines or encapsulated in ice is less efficient. The confinement of surface oil along the edges of ice can result in both the encapsulation of oil in freezing ice or the ‘booming’ of oil along ice edges increasing the opportunities for response actions.

Evaluation of the transport and fate of oil residues that remain after application of an OSR is a key component to assessment of long-term ecosystem consequences. In order to formulate a proper net environmental benefit analysis (NEBA) decision framework for Arctic oil spill response it is important to understand the differences in transport and fate of the remaining oil components after OSR actions to and within Arctic environmental compartments (ECs). Important processes include physical, chemical, and biological complexities associated with the transport, fate, exposure, and effects of oil and OSR residuals in the Arctic environment. It is crucial that the underlying processes are understood in a quantitative as well as a qualitative way. Key recommendations therefore include the development of a better understanding of environmental compartment attributes that affect the dissolution, mineralization and biodegradation of oil especially at ice/water interfaces and that decrease or increase resiliency.

3.1 Introduction

The type of oil designates the relative spectrum of chemical components present in the oil, e.g. fuel oil versus heavy crude oil.  Table 3-1 summarizes the various chemical components that may be present in oil.  The properties of the oil itself will define the physico-chemical interactions that will ensue following release to the marine environment, and will influence the important processes represented in Figure 3-1 (e.g. spreading, dissolution, evaporation).

When oil is spilled at sea, a number of natural processes change the physical and chemical properties of the oil and influence interactions with seawater and ice. These natural processes are spreading, drifting, evaporation, dissolution, photolysis, biodegradation and formation of oil-in-water (dispersion) and water-in-oil (w/o) emulsions. The relative contribution of these processes varies during a spill under cold conditions.  Figure 3-1 illustrates how the different weathering processes vary with time. Evaporation is, for example, most important the first days after the spill, while biodegradation is a significant process one to two weeks after the spill (see Section 5). The behavior of spilled crude oils and petroleum products depends on the oils physical and chemical properties, the release conditions, and the environmental conditions (e.g. temperature, waves, wind, currents, presence of ice).  Factors that increase the spread of an oil slick are gravity and net interfacial tension whereas inertia and viscosity retard its motion (Hollebone 1997).

Table3-1.  Oil constituents based on carbon composition

ClassMass
(LMW to HMW)
FractionRepresentative CompoundsSolubility mg/LVapor  Pressure atmHenry’s  Law Constant cm3/cm3Log Koc1

Aromatics

Total Petroleum Hydrocarbons (TPH)

Total    Polynuclear Aromatic Hydrocarbons (TPAH)

Low    to High

C5-C7

Benzene

220

0.11

1.5

3

C>7-C8

Toluene

130

0.035

0.86

3.1

C>8-C10

Isopropyl    Benzene, Naphthalene

65

0.0063

0.39

3.2

C>10-C12

25

0.00063

0.13

3.4

C>12-C16

5.8

0.000048

0.028

3.7

C>16-C21

Fluorene, Fluoranthene, Benzo(a)pyrene

0.65

0.0000011

0.0025

4.2

C>21-C35

0.0066

0.00000000044

0.000017

5.1

Aliphatics

Total    Saturated Hydrocarbons (TSH)

Low    to High

C5-C7

n-Hexane

36

0.35

47

2.9

C>7-C8

5.4

0.063

50

3.6

C>8-C10

n-Alkanes

Kerosene,    Dearomatized petroleum

0.43

0.0063

55

4.5

C>10-C12

0.034

0.00063

60

5.4

C>12-C16

0.00076

0.000076

69

6.7

C>16-C35

Mineral    Oils

0.0000025

0.0000011

85

8.8

1Organic carbon/water  partition coefficient

Weathering processes relative importance with timeFigure 3-1. Weathering processes relative importance with time (Adapted from Mackay et al. 1983)

A key determinant of oil weathering is based on the viscosity of the oil mixture.  This property largely affects the tendency of the specific oil to spread out on surfaces under cold conditions.  Oils have a range of viscosities; less viscous oils cover larger surface areas than the more viscous oil types which also tend to spread more slowly.  As the oil spreads and the thickness of the oil slick becomes thinner, the release of soluble and volatile compounds is increased.  The cold temperatures in the Arctic increase the viscosity of the oil and thus reduce the rate of spreading on surfaces.  The reduced spreading and increased thickness of the oil decreases the loss of volatile and soluble components to the air or water, respectively.  The cold temperature also slows the rate of volatility and solubility of the chemical compounds. 

These changes in viscosity and the impact that has on volatile and soluble compound releases is one factor that extends the time scales portrayed in Figure 3-1, allowing an extended time period for OSR actions to be implemented.  Since the Arctic is significantly influenced by seasonally prevailing climate conditions, oil behavior changes according to summer, winter, and transitional seasonal conditions.  During the summer periods surface oil may encounter shorelines, lagoons, estuaries, and convergence zones in open waters.  Although the fate of oils during this season is generally similar to non-Arctic environments, factors such as lower temperatures and decreased viscosity of the oil reduce the speed of volatilization and the occurrence of 24-h light periods increasing the potential for UV activation on constituents such as PAHs that can impact organisms, especially those living near surface waters.  The influence of these factors on biodegradation and exposure to organisms is similar to other high latitude non-Arctic environments. The influence of cold water on shoreline stranding to various substrates during ice-free periods is an area of ongoing study. Oil may become entrained within or under the ice during winter conditions as seawater becomes covered in ice.  In ice-covered waters, several studies have indicated that the time-dependent weathering processes can be slower as a result of less energy input, increased oil film thickness, and lower temperatures.  Pack ice consists of a variety of ice types depending on the season.  In the fall, pack ice may contain a mix of older ice from the previous winter, vast floes of thin new ice known as “nilas”, and patches of newly forming ice.  In the spring, pack ice consists of deteriorating first year floes.  Frazil or slush ices are typically associated with freeze-up whereas brash ice is typical of spring melt (Buist et al. 2003). 

The bioavailability of oil components in the water column is governed by processes both generating and depleting bioavailable oil fractions.  While the process of generating oil droplets and dissolution of water soluble components are reasonably well understood, the potential importance and kinetics of the depletion processes determining the fate of oil in ice and oil droplets in the water column are not well understood.  Chemical dispersion, which increases the fraction of dispersed oil, is now more frequently included as one of the response actions after an oil spill, but not in all countries.

3.2 Knowledge Status

This section focuses on the physical fate of oil, relevant for Arctic conditions, with emphasis on identification of literature and studies of operative importance. An extensive body of research performed both in the field and in the laboratory over the last 30 years has aimed at understanding the fate, behavior and weathering processes for oil spilled under Arctic conditions.  Most of this research has been performed in USA, Canada, and Norway.   Numerous field investigations including spills of opportunity, laboratory and tank tests, and theoretical modeling of the fate and behavior of oil under Arctic conditions have been summarized by Dickins and Fleet (1992), Hollebone (1997), Fingas and Hollebone (2003), Brandvik (2007), Sørstrøm et al. (2010), SL Ross et al. (2010), and Dickins (2011).  These compilations of salient technical elements from this broad range of historical studies contributes to the current knowledge base of information on the physico-chemical interactions of oil and sea ice. 

Experimental oil releases under solid ice (up to 6 tons per spill) were performed in the 1970s and early 1980s (reviewed by Hume et al. 1983).  The first experimental oil spill in broken ice was carried out in 1986 on the Canadian East Coast (SL Ross and Dickins 1987).  This project was followed by the first large-scale experimental oil spill (26 m3 oil) in April 1993 in the Barents Sea marginal ice zone (Vefsnmo and Johannesen 1994; Singsaas et al. 1994). These field experiments proved invaluable in understanding the weathering processes of oil in a variety of spill scenarios and environmental conditions (including wind, waves, ice conditions, drift and spreading in the marginal ice zone).  The studies clearly showed that marine oil spills have different weathering properties under Arctic conditions compared to spills under more moderate temperatures (Brandvik et al. 2004).

The different aspects of oil weathering in Arctic environments have been summarized by Payne et al. (1991a) and field observations regarding weathering at low temperatures and in broken ice were also studied in small- and meso-scale lab facilities (Singsaas et al. 1994).  The findings from an experimental oil spill of 3400 liters of Statfjord crude under first-year sea ice on Svalbard in March 2006 are described in Dickins et al. (2007).  The experiment demonstrated the ability of ground penetration radar to detect and map oil under natural sea ice from the surface.  It also documented oil weathering beneath relatively warm ice sheets under spring conditions.

Several reports provide good overviews of the state of knowledge, e.g. the final report from the Norwegian program “Oil spill contingency in cold and Arctic areas” – ONA I and ONA II (Løset et al. 1994) and “Oil spill response in ice infested waters” (Vefsnmo et al. 1996). The ONA research program included information on physical environment, behavior and properties of oil and oil spill response (biological, burning, dispersing, emulsion breaking, and mechanical oil recovery).  Later work evaluating the status and research needs for future oil in ice research was presented in Dickins (2004), Owens (2004), and Brandvik et al. (2005 and 2006).

Photo 3-2. Oil-in-Ice JIP field investigation (SINTEF, Liv-Guri Faksness)
Photo 3-2. Oil-in-Ice JIP field investigation (SINTEF, Liv-Guri Faksness)

Fingas (1992) summarized accidental spills and the experiments conducted from the 1970s up to 1990 to gain knowledge and understanding of the interactions occurring when oil is discharged in water where ice is present.  A more recent review (Fingas and Hollebone 2003) summarized the studies of oil behavior in ice infested environments related to the specific ice situation or behavioral mechanism.  The authors concluded that the knowledge basis of the complex behavior and fate processes associated with oil spills in ice infested water should be improved in order to better predict oil behavior and fate in an ice infested environment.

The latest large-scale field experiment took place in the marginal ice zone in the Barents Sea in May 2009 as a part of the Joint Industry Program (JIP) to develop and advance the knowledge, methods and technology for oil spill response in Arctic and ice-covered waters (Oil-in-Ice JIP; Brandvik et al. 2010).  Fresh oil (7000 liters) was released uncontained between the ice floes to study oil weathering and spreading in ice and surface water throughout the six-day experiment. In addition, meteorological and oceanographic data that were monitored included wind speed and direction, air temperature, currents and ice floe movements.  The monitoring showed low concentrations of dissolved hydrocarbons; the predicted acute toxicity associated with that level of exposure indicated that the potential for acute toxicity was low. The ice field drifted nearly 80 km during the experimental period, and the oil drifted with the ice and remained contained between the ice floes (Faksness et al. 2011a).  The JIP summary report gives an overview of the total program, main findings, and the technical reports (Sørstrøm et al. 2010). Findings from the program show that the presence of cold water and ice can enhance response effectiveness by limiting oil spreading and slowing down the weathering processes, resulting in an increased window of opportunity for in-situ burning or use of dispersants compared to ice-free conditions.

3.2.1 Weathering of Oil Spilled in Ice

After release to the Arctic marine environment, oil will gradually weather due to natural physical and chemical processes including spreading, evaporation, dispersion, emulsification, dissolution, and chemical modification by oxidative processes (refer to Figure 3-1). From earlier studies it is known that this weathering is influenced by the oil composition, release conditions, temperature, water quality, and solar radiation (Duursma and Dawson 1981, Sydnes et al. 1985, and Sydnes 1991).

Environmental conditions will influence an oil spill in the Arctic differently compared to temperate regions, particularly under conditions of lower temperatures, the presence of ice, and different light conditions. Low temperatures and lack of waves in ice act to reduce oil spreading, evaporation, emulsification and dispersion (e.g. Brandvik et al. 2010).  Wax and asphaltene content of oils that act in combination with reduced spreading due to temperature-dependent oil viscosity differences combined with reduced weathering that occurs within the more stable ice fields are important aspects of the fate of oil in the Arctic.  These factors extend the period of time when various response actions can be deployed.   

Different locations of spilled oil associated with ice are: oil between broken ice, oil under the ice, submerged oil, and oil on the surface of the ice including oil in melt ponds as shown in Figure 3-2. . The oil can be entrapped in ice and become difficult to track particularly during the winter darkness. The sequestering of oil also results in a secondary discharge situation during the spring season, where the oil released from the ice appears relatively un-weathered (Payne et al. 1991a and 1991b).  In the spring and summer season chemical photo-oxidation of oil may become an important hydrocarbon degradation process.  The water-soluble components released from the encapsulated oil may be transported through the brine channels, thereby contacting sea-ice microbes in the brine and the underlying water.  Ice fauna do not avoid the exposure by moving to non-oiled ice locations and may be exposed to toxic water-soluble components for a prolonged period of time while the oil is present within the brine channels, crack and crevices of the ice.

In the summer, local melting enhanced by changes in albedo due to the presence of oil causes the ice to break up weeks earlier than normal seasonal break-up.  Once the oil reaches the surface of the ice or within leads or polynyas, evaporation and other weathering processes commence.  However, low temperatures and increasing film thickness due to confinement in ice reduce the rate and degree of evaporation (e.g. Brandvik and Faksness 2009, and Brandvik et al. 2010).  As the ice cover dwindles, the oil slick begins to act similarly to open water conditions.  Since biological activity is high in early spring, and the amount of open water available for birds and mammals is small, the release of oil into leads and polynyas at this time would be expected to be damaging to resident and migratory species (AMAP 1998).

The presence of ice reduces the wave activity (breaking waves), and emulsification usually decreases with increasing ice coverage.  The water uptake rate decreases with increasing ice coverage due to the wave damping effects, and will be slower in dense sea ice.  The viscosity increases with increasing water uptake and evaporation as in open sea, but the increase will be slower due to reduced evaporation and water uptake. The rate of natural dispersion decreases with increasing ice coverage and could be very low due to reduced energy conditions in the ice (Singsaas et al. 1994, and Brandvik et al. 2010).  Spreading of oil in ice is dependent on ice type and ice coverage with oil film thickness increasing with increasing ice coverage (Vefsnmo and Johannesen 1994, Fingas and Hollebone 2003, and Brandvik and Faksness 2009).

Figure 3-2. Weathering of oil spilled in ice infested environment (AMAP 1998)
Figure 3-2. Weathering of oil spilled in ice infested environment (AMAP 1998).

3.2.2 Oil in Ice Interactions

Spill response strategies in open waters versus the ice seasons (spring breakup, fall freeze, and winter) are expected to be fundamentally the same.  Table 3-2 presents general characteristics of oil movement with associated ice conditions.  Oil drifting in ice-covered waters is very dependent on the amount of surface area covered by ice.  The knowledge about oil drift in ice and oil-ice interactions is limited, but the current assumption is that at ice coverage less than 30%, the drifting of oil will be independent of the ice.  At ice coverage greater than 60-70%, the oil will mainly drift with the ice (Vefsnmo and Johannesen 1994).  Under pack ice conditions, oil will move with the ice.  The rate of drift of pack ice influences film thickness and areal distribution of spilled oil.  Ultimately, the rate of pack ice drift determines the magnitude of offshore logistics required to recover released oil since the ice can travel hundreds of km in a few months.  Open drift ice presents a challenge for spill containment and recovery.  The ice itself can contain oil as a natural boom and encapsulate oil as the ice begins to freeze resulting in transporting the oil away from the active response areas.

Photo 3-3. Field experiment during Oil-in-Ice JIP program (SINTEF)
Photo 3-3. Field experiment during Oil-in-Ice JIP program (SINTEF)

During a six-day experimental field trial in the Barents Sea in May 2009 the ice concentration in the area varied from 70 to 90% (Faksness et al. 2011a).  During this experiment a storm occurred which resulted in transport of the oil and ice over a long distance. However, the key observation is that even under storm conditions the oil slick was still contained within the ice field, thus the monitoring during the experiment verified that in high ice concentrations oil drifts with the ice.

Prevalence of ice is indicative of environmental conditions that do not facilitate weathering of oil:  low air and water temperatures as well as a relative lack of waves significantly reduce the rate of evaporation, natural dispersion and emulsification.  Additionally, oil spilled in the Arctic marine environment can be rapidly frozen into the ice as the ice grows downwards and encapsulates oil beneath or within the ice during freezing.  Once the oil becomes fixed within ice, it moves only as the ice moves. The oil will in this way be preserved, in the sense that evaporation, dissolution, and degradation are expected to be reduced (Hollebone 1997).

Table 3-2.  Sea ice and behavior of oil

Type of Ice FieldDescriptionConc1Characteristics

Fast Ice

Ice attached to or contiguous with the shore

 
  • Does not move with currents and wind
  • Most stable (extensive in areas with broad shallow shelf extending offshore)
  • Extends to 10 - 12m (water depth); potentially to 30m (but rougher terrain)

Open Drift Ice

Floats freely

1-5/10

  • Oil spreads similarly to open water conditions
  • Would not provide for stable operations; also little natural ‘containment’ properties
  • Relatively short term occurrences (e.g. 25 d/year)

Pack Ice

6/10

  • Unpredictable movement
  • Tends to corral oil

Close Pack Ice

7-8/10

Very Close Pack Ice

9-9+/10

1Ratio of pack ice to visible open water; adapted from Dickins 2011

This implies that the oil will retain much of its potential toxicity upon release from the ice, either via transport in brine channels, and/or eventual breakup and melting of the ice.  Work to date has led to a good general understanding of the key processes controlling the behavior of fresh and emulsified crude oil in a variety of ice conditions, including landfast and broken pack ice.    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 and ablation of the ice surface down to the oil lens in the ice (Fingas and Hollebone 2003).  Oil, initially trapped under or in the ice, may appear on the upper ice surface due to density-mediated migration up through the open brine channel pathways.  Migration rates of oil in brine channels during the spring thaw may vary as a function of oil chemistry and viscosity, as controlled by water contents due to previous emulsification and temperature gradients within the ice, as well as depth in the ice canopy (Martin 1979, and Payne et al. 1991a).

When oil is released or drifts into an area with ice, spill responders will face a complex interaction between oil, water and ice.  A review by Fingas and Hollebone (2003) summarized the studies of oil behavior in ice environments under the topics of specific ice situations or behavioral mechanisms.  The oil will be absorbed by snow on the ice edges, it may be trapped in the ice in brine channels (Faksness and Brandvik, 2008b) or it may be moved underneath the ice.  The ice field will also be under a constant transformation driven by wind, currents and temperature.  The result over time is that the individual ice floes may change their relative position and melt or freeze.  Some ice floes may be transported relatively far from both their original position and their original neighbors, which may be a strong driving force for the drift and spread of oil after oil has been released in an ice field.

Until recently, most of the research conducted on migration of oil through ice has focused on bulk oil (e.g. NORCOR 1975, Martin 1979, Payne et al. 1991a Reed et al. 1999, Fingas and Hollebone 2003, and Brandvik et al. 2004).  Existing conceptual models for oil-in-ice mainly discuss upward migration of oil through brine channels during melting, due to density differences, solar radiation and heat capacity of the oil.  Very few studies have actually attempted to determine the transport and fate of individual compounds, such as PAHs or changes in oil compositions in ice scenarios.  The field and tank studies of Payne et al. (1991a and 1991b) show that dissolved aromatic hydrocarbons in brine waters are effectively transported downward with the dense brine water.  The knowledge of the migration process of the water soluble components from oil encapsulation in ice has been very limited. However, recent work from three field seasons with oil encapsulated in first-year Arctic sea ice on Svalbard has shown that the more soluble oil components are transported downwards through brine channels (Faksness and Brandvik, 2005, 2008a, 2008b, and Faksness et al. 2011b).  Furthermore, the results from this research suggests that the presence and dynamics of brine channels transports the bioavailable oil components downward from an “encapsulated slick” through diffusion and advection. These observations are in accordance with the findings in laboratory experiments with sea ice columns, which have shown that there is a downwards migration of water soluble oil components from oil encapsulated in the ice, and that the migration starts after spring thawing has increased the porosity of the ice (Faksness et al. 2011b).  However, Faksness and Brandvik (2008b) observed an upward migration of oil during their field experiments on Svalbard, but no upward transport was observed during the melting process in these laboratory experiments, indicating that the tide movement, sunlight, and albedo effect are important factors that should be taken into account.  The differences in the oil’s migration rates as a function of oil properties and chemical composition are illustrated in Photos 3-4 through 3-7.  As can be seen in the upper photos in this figure, the major part of the encapsulated oils of Heidrun Åre (naphthenic oil) and Goliath had migrated through the brine channels to the surface.  For Kyrtael (bottom left photo), which has high wax contents and a pour point of 8 °C, there had been some vertical rise through the ice, but mainly only melting of the ice over the oil.  The major part of the oil was still encapsulated in the ice. The heavy fuel oil, IFO 180 (bottom right photo), was still encapsulated in the ice, and no upward migration had occurred.  IFO 180 was the oil with the highest viscosity in these experiments. The transport and fate of oil components within different oil types need to be performed in order to determine the controlling aspects of what portions of the oil move within an ice column.

Photos 3-4—3-7. Details from sampling of ice cores with encapsulated oil in June, 2004. Melt ponds with Heidrun Åre and Goliat in the upper photos, and Kyrtael and IFO 180 in the bottom photos (From Faksness 2008c)
Photos 3-4—3-7. Details from sampling of ice cores with encapsulated oil in June, 2004. Melt ponds with Heidrun Åre and Goliat in the upper photos, and Kyrtael and IFO 180 in the bottom photos (From Faksness 2008c).

3.2.3 Oil on Arctic Shorelines

The short and long term fate of stranded oil in the Arctic has not been extensively studied.  What is known from other environments is that the slope and sediment characteristics of the intertidal environment has a strong influence on the retention of stranded oil.  Additionally, those studies in non-Arctic environments also indicate that oil stranded in deep sediments, especially within cobble intertidal environments can retain oil that continues to impact the environment for decades (Etkin et al. 2007; Peterson et al. 2003).  In contrast, stranded oil on sandy intertidal environments with ice is subject to large physical disturbances when the ice begins to break up and is remobilized during that period (Potter et al. 2012; see also Section 4).  Therefore, the preferred OSR option should protect cobble beach environments from oiling because of the extended residence time of oil that does not undergoing significant weathering.  In most instances, the presence of ice-covered water or ice in the shore zone prevents surface oil from making contact with the shoreline substrate.  Owens (2004) summarized the following relevant scenarios for oil on icy shorelines:

  • Where oil comes in contact with exposed ice surfaces, the oil is unlikely to adhere except in cold temperatures when the air, water, and oil surface temperatures are below 0 °C.
  • During freeze-up, oil present on the shore or stranded on the shore-zone ice during a period of freezing temperatures can become covered and encapsulated within the ice.
  • During a thaw cycle, or if the surface of the ice is melting and wet, oil is unlikely to adhere to the ice surface and would remain on the water surface or in shore leads.
  • In broken ice, without a landfast ice cover, oil may reach the shore and be stranded on the substrate in between the ice floes.
  • If continuous shore-fast ice (an ice foot) is present, the ice may protect the shore zone, but in the few instances when near shore ice is present adjacent to the shore zone as a solid floating ice layer, oil can migrate through ice cracks and accumulate under the ice.
  • Ice in beach sediments (frozen groundwater) can prevent the penetration of stranded oil.

3.2.4 Oil-Sediment Interactions

The fate of oil that becomes associated with intertidal and subtidal sediment has been extensively studied in many areas throughout the world and there are good predictive models that describe the uptake of toxic components from sediment based on the concentrations of organic carbon and the octanol-water partition coefficient (Kow) of the compounds.   There are also excellent studies on the reduction in bioavailability of the toxic compounds based on the decreased desorption that occurs with time.  These studies are not a part of the current review but demonstrate that oil associated with sediment becomes less bioavailable to processes of toxicity and presumably biodegradation.

The first association of oil with sediment occurs within the water column where it comes in contact with suspended sediment grains or solids.  If the specific gravity of the suspended sediment is sufficient to counteract the buoyancy of oil droplets then the combined materials will settle to the substrate at rate that is based on Stoke’s Law.  The current literature review identified a number of studies on oil-mineral-aggregation (OMA) with focus on use of fine, clay-sized particles.  Some studies have included systematic variation of oil type, salinity, presence of dispersants and mineral type, thereby increasing our understanding of the importance of these variables in OMA.  Many of the studies have focused on the potential use of fine mineral aggregation as a combat technique for free floating and beached oil spills, but there is a lack of systematic studies on the adsorption and trapping of oil in bottom sediments under turbulent conditions in shallow waters. The size and density of these mineral particles and their association with oil droplets counteract part of the buoyancy of the oil droplets, suspending the mixtures in the water but the buoyancy is not sufficiently impacted to have them settle rapidly.   

Etkin et al. (2007) has prepared a literature review on state-of-the-art on modelling interactions between spilled oil and shorelines.  They state, in general, that OMA does not play a significant role in the fate of oil in the early stages after oil deposition on the shoreline, and, as such, is of relatively minimal importance in shoreline-oil interaction modelling efforts; however, OMA may play a role in longer-term shoreline processes.  Oil/sediment interactions may be very important in areas where there are higher concentrations of fines or in areas with coarse gravel and cobble with sub-surface open spaces that can trap oil.  A number of studies focused on OMA as a pathway for transport of oil from shorelines, sea-surface and the water columns to the seafloor.  These studies emphasize the modification to the settling rates that occur with well mixed oil and OMA and tend to over predict the amount of oil that would settle to the bottom.  Lee (2002) has also given an overview of OMA formation, including the formation of OMA as a method to mobilize and remove stranded oil from low-energy, intertidal environments.  Increasing knowledge of this process has fostered the development and evaluation of oil spill countermeasure strategies based on the promotion of oil-particle interactions.  Muschenheim and Lee (2002) provide a comprehensive literature review of the role of oil-particle interactions in removal of petroleum hydrocarbons from the sea surface and provide estimates of the small degree to which presence of particulate matter augments the deposition of oil. They discuss the interaction between oil weathering, placement methods of the OMA on the oil, sinking, adsorption, microbial processes, flocculation and ingestion by zooplankton, which all contribute to packaging oil and suspended particulate matter (SPM) into settling aggregates.  Findings from the literature covering many of these processes are described in the following.  The majority of published literature on characterization of adhesion of oil to particulate matter has focused on OMA, where much of the work has been directed towards investigation of OMA as a response action.  Delvigne (2002) investigated the physical appearance of oil in three types of oil-contaminated sediment.  Microscopic observations showed possible presence of three oil phases in the sediment as distinct oil droplets, oil coating on sediment particles and as 'oil patches'.  It was concluded that the division of oil in the different phases affects the oil-sediment interaction processes.  A number of studies have focused on adhesion properties in general, uncoupled to a mechanistic understanding of the process (Sterling et al. 2004, Ajijolaiya et al. 2006, Ma et al. 2008, Devadoss et al. 2009, Lee et al. 2008, Lee et al. 1996, and Weise et al. 1999).  These studies show that OMA can be an important process for determining the fate of oil in scenarios where the properties of oil are favorable, and mineral fines are present in necessary quantities.

Studies on the effect of viscosity of oil on OMA (Kepkay et al. 2002, Danchuk and Wilson 2011, Omotoso et al. 2002, Payne et al. 2003, and Stoffyn-Egli and Lee 2002) have shown that oils with low density have a higher propensity to form OMA than oils with higher density.  Other studies have focused on the effect of salinity on OMA (Danchuk and Wilson 2011, Guyomarch et al. 2002, Le Floch et al. 2002, Khelifa et al. 2005, Lebedeva and Fogden 2010, and Sterling et al. 2004), where the major findings are that the effect of salinity on droplet size distribution is strongly influenced by clay type, and that there is a minimum salinity threshold for obtaining significant OMA.

The effect of mineral type and surface properties has been investigated in a few studies (Danchuk and Wilson 2011, Omotoso et al. 2002, Stoffyn-Egli and Lee 2002, Zhang et al. 2010, and Wang et al. 2011). These studies show an effect of mineral type and mineral grain size on the adhesion strength, where better adhesion properties are found for the smallest particles.  The surface energy seems important for understanding the adhesion strength, and hydrophobic minerals will give rise to better adhesion, compared to hydrophilic minerals.  The strength of adhesion is also a function of ion strength (salinity), oil type (viscosity and content of natural surface active components), pH, and presence of chemical dispersants.

The effect of mixing energy has been investigated and discussed in some studies (Devadoss et al. 2009, Sterling et al. 2004, Cloutier and Doyon 2008, and Stoffyn-Egli and Lee 2002), where they find a strong positive correlation between mixing energy (i.e. breaking waves, strong flood currents) and OMA formation.

Effort has been made to develop models that include the process of OMA, and use these for oil mass balance calculations, and as a decision support tool for oil spills (Khelifa et al. 2002, Hill et al. 2002, Bandara et al. 2011, Lee et al. 2002, Sterling et al. 2005, and Niu et al. 2010 and 2011).  A number of papers describe field observations of OMA formation, both after accidental oil spills and controlled field experiments (Owens et al. 1994, Owens 1999, Lee and Lunel 1997, Lee et al. 2002, Lee et al. 2003, Owens and Lee 2003, and Wood et al. 1997). These field studies all show the potential for OMA formation to occur, and to be an important factor for fate of spilled oil in the Arctic.

The use of OMA as a OSR technique has been investigated and discussed in several studies (Ajijolaiya et al. 2007, Lee and Lunel 1997, Owens 1999, Owens and Lee 2003, and Sun and Zheng 2009), and these studies clearly conclude that enhancing OMA formation in certain cases might serve as a sound response option for oiled shorelines, with references to field studies at two spill events the Tampa Bay response in Florida and the Sea Empress operation in Wales and at a controlled oil spill experiment in the field [the 1997 In Situ Treatment of Sediment Shorelines (ITOSS) Program (Guenette et al. 2003)].

Investigations of oil-particle interaction after the Deepwater Horizon accident in the Gulf of Mexico are described by Passow et al. (2012). They proposed a mechanism for understanding the formation of marine snow on the basis of the complex processes involved in bacterial decomposition of oil and subsequent interaction with organic matter and phytoplankton.  The findings related to oil-particle interactions from the Deepwater Horizon spill demonstrate that oil-particle interactions are important also for deep water and water column processes affecting dispersed oil, and that knowledge gaps related to these processes remain.

3.3 Future Research Considerations

The review of the transport and fate of oil in the Arctic described by the authors in this section led to suggestions of further research which can reduce remaining uncertainties.  The more generic suggestions compiled from this review are summarized below while recommendations that are important for improving Arctic NEBA are listed separately.

Behavior of Oil in the Arctic. Develop a better understanding of the key environmental variables that alter the dissolution, mineralization and biodegradation of oil.  As an example, it has been well demonstrated that small droplets of oil are able to more rapidly undergo degradation than larger oil droplets and that this is based on the surface area exposed to microbes.  Using this basic understanding it is reasonable to conclude that concentrated oil masses (at the air/water or ice/water interfaces, convergence zones, shorelines, and within shoreline beaches) will be less bioavailable and decrease the ability of microbes to effectively degrade the oil.  Habitats that accumulate oil are more likely to experience longer term impacts on resident biota at the concentration zone or in adjacent areas as a result of oil remobilization.   The ability for these environments to recover is therefore based not only on the resilience of the organisms at these locations but on the duration of contact with concentrated oil.

  1. Impact of seasons:  Develop a better understanding of the seasonal physical and chemical weathering processes on oil that occur in brine channels under ice and in polynyas to provide better exposure estimates and increase the accuracy of impact assessments of these specialized ECs. 
  2. Shoreline stranding:  The seasonal difference in stranding processes for oil on shorelines is beginning to be well understood.  Ice can act as a protective barrier, minimizing contact and stranding of oil in the presence of ice.  During open water periods the stranding process is different and the effects on shorelines need to be characterized to provide an assessment of the potential for stranded oil to reside for extended periods of time with little to no additional weathering.
    1. As part of an input to tactical OSR plans mapping of shorelines with a propensity to recover rapidly and those that are projected to retain unweathered oil should take place preferably using GIS.  GIS is recommended because the seasonal use of the various types of environments by valuable ecosystem component (VEC) species could be added as an information layer.  One example of a shoreline is a sandy beach that may be able to recover rapidly from stranded oil but it may also be located in areas of marine mammal (e.g. walrus) haul-outs.  Knowing the location of this type of EC combined with seasonal patterns of haul out use would provide needed information to protect VECs from this type of exposure.  Similarly, a cobble shoreline may have an extended time for recovery with a release of oil on a regular basis that could influence the spawning habitat for VEC species that use the area.  The regional identification of shoreline ECs and the spatial and temporal use of local intertidal environments will provide a specialized component for regionally adapted NEBA assessments. 
    2. Experimental studies examining OMA suggest that OMA may play a useful role in longer-term shoreline processes. The use of OMA on stranded oil to remobilize the oil is an intriguing response option that needs to be characterized more completely at mesocosm and field scales. 
  3. Oil-particle interactions:  Interactions between dispersed oil and other types of particles, e.g. marine snow, microorganisms, phytoplankton, and zooplankton have not been well characterized. These interactions need to be investigated in order to produce new data to improve fate modeling tools.
    1. Similarly, whether OMA are predominantly consumed as food and/or transferred as contaminants into the Arctic food webs is unknown.  Establishing the rate of biodegradation and fate of oil contained in OMA are fundamental steps for evaluating the potential positive and negative effects of OMA to pelagic, demersal and intertidal communities.

3.3.1 Priority Recommendations for Enhanced NEBA Applications in the Arctic

The recommendations presented below indicate where increased knowledge of oil transport and fate processes would result in reducing existing uncertainties in NEBA assessments.  No prioritization has been made to the list; for some of the recommendations, surrogate data may be already available.

  1. Attributes of ECs that impact fate of oil.  Develop a better understanding of environmental compartment attributes that affect the dissolution, mineralization and biodegradation of oil.  Two primary compartments for study include the sea surface and ice/water interfaces.
  2. Attributes of ECs that affect resiliency and recovery.  Habitats that accumulate oil are more likely to experience longer term impacts on resident biota including additional effects of oil remobilization.   The ability for these environments to recover is therefore based on the resilience of the compartment and removal of oil.  Define environmental compartment attributes the decrease or increase resiliency and ultimately recover.
  3. Remobilization of oil.  Information on the chromatographic separation of oil components as they become sequestered into ice and brine channels and the extent of biodegradation and physical/chemical changes that occur during ice encapsulation would provide information needed to assess remobilization of oil during ice break-up and provide justification for extending response action periods.
  4. Fate of oil associated with ice communities.  The fate and effects of oil under ice where ice algae and species living in that compartment become exposed need further characterization.
  5. Fate of OSR residues.  Each OSR action produces residual materials that behave differently.   Modeled transport processes should include settling of particles derived from OMA, residues from in-situ burning (e.g. increased volatilization and ash formation), natural or enhance dispersion of oil, transport of surface oils, retention of stranded oil on or in shorelines, and convergence of water masses to adequately predict the fate of all oil components.
  6. Role of biological processes on degradation of OSR residues.  The physical and microbial processes that control the fate of OSR residues also need to be incorporated in transport and fate models.

3.4 Further Information

Authors Drs. Alf Melbye and Liv-Guri Faksness (SINTEF), Dr. Janne Fritt-Rasmussen (ARTEK), Dr. Torgeir Bakke (NIVA), Dr. Oleg Titov (PINRO) 

References

Ajijolaiya LO, Hill PS, et al.

2006

Laboratory investigation of the effects of mineral size and concentration on the formation of oil-mineral aggregates

Mar Pollut Bull 52(8):920-927

Ajijolaiya LO,Hill PS, et al.

2007

Qualitative understanding of the mechanism of oil mineral interaction as potential oil spill countermeasure - A review

Energy Sources Part a-Recovery Utilization and Environmental Effects 29(6): 499-509.

ATSDR

1999

Toxicological profile for total petroleum hydrocarbons (TPH).

US Health and Human Services, Agency for Toxic Substances and Disease Registry (ATSDR). Atlanta, Georgia. 315 p.

Bandara UC, Yapa PD, et al.

2011

Fate and transport of oil in sediment laden marine waters

J Hydro-Environ Res 5(3):145-156

Bragg JR, Yang SH

1995

Clay oil flocculation and its role in natural cleansing in Prince William Sound following the Exxon Valdez oil spill

In: Exxon Valdez Oil Spill: Fate and Effects in Alaskan Waters. P. G. Wells, J. N. Butler and J. S. Hughes. W Conshohocken, American

Brandvik PJ

2007

Behaviour and weathering of oil spills under Arctic conditions and implications for response

Proceedings, International Oil & Ice Workshop; MMS, Herndon VA

Brandvik PJ, Singsaas I, Daling PS

2004

Oil Spill R & D in Norwegian Arctic Waters with Special Focus on Large-scale Oil Weathering Experiments

Paper presented at The Inter Spill Conference, Trondheim, Norway, 14-17 June, 2004

Brandvik PJ, Sørheim KR, Singsaas I, Reed M

2006

Short state-of-the-art report on oil spills in ice-infested waters. Oil behaviour and response options

SINTEF report STF80MKA06148, 59 p

Cloutier, D. and B. Doyon

2008

OMA formation in ice-covered brackish waters: Large-scale experiments.

In: Oil Spill Response: A Global Perspective. NATO Science for Peace and Security Series C: Environmental Security. W. F. Davidson,

Danchuk S, Willson CS

2011

Influence of Seasonal Variability of Lower Mississippi River Discharge, Temperature, Suspended Sediments, and Salinity on Oil-Mineral Aggregate Formation

Water Environ Res 83(7):579-587

Delvigne GAL

2002

Physical appearance of oil in oil-contaminated sediment

Spill Sci & Tech Bull 8(1):55-63

Devadoss RS, Sannasiraj SA, et al.

2009

Laboratory Investigation of Clay Mineral Fines as Dispersant for Combating Oil Spill

New York, Amer Soc Mechanical Engineers

Dickins DF Assoc and Fleet Technology Limited

1992

Behaviour of spilled oil at sea (BOSS): Oil-in-ice fate and behavior.

Draft report submitted to Environment Canada, U.S. Minerals Management Service and American Petroleum Institute.

Duursma EK, Dawson R (eds)

1981

Marine Organic Photochemistry: Evolution, Composition, Interactions and Chemistry of Organic Matter in Seawater

Elsevier, Amsterdam

Etkin DS, D French McCay, J Michel

2007

Review of the State of the Art on Modeling interactions between spilled oil and shorelines for the development of algorithms for oil spill risk analysis modeling

OCS Study, MMS 2007-063

Faksness LG, Brandvik P, Reed M

2006

Chemical characterization of water soluble components from !rctic marine oil spills – a combined laboratory and field study

Paper presented at the 9th International Marine Environmental Modelling Seminar (IMEMS), Rio de Janeiro, 9-11 October, 2006

Fingas MF

1992

The behaviour of oil in ice

Proceedings, Seminar Combatting Marine Oil Spills in Ice and Cold Conditions. National Board of Waters and the Environment, Hel

Fingas MF, BP Hollebone

2003

Behaviour of oil in freezing environments -review

Mar Poll Bull 47:333-340

Fingas MF, Hollebone BP

2003

Review of Behaviour of Oil in Freezing Environments

Mar Pollut Bull 47:333-340

Fingas MF, Hollebone BP

2001

The Fate and Behaviour of oil in Freezing Environments

Proceedings, Seminar on Combatting Marine Oil Spills in Ice and Cold/Arctic Conditions. Finnish Environment Institute, Helsinki, 81

Fu Y

2011

The Forming Mechanism of a New Oil-Mineral Aggregate

Mechatronics and Intelligent Materials, Pts 1 and 2. R. Chen. Stafa-Zurich, Trans Tech Publications Ltd. 211-212: 1176-1181

Fu Y, Chung DDL

2011

Coagulation of oil in water using sawdust, bentonite and calcium hydroxide to form floating sheets

Appl Clay Sci 53(4): 634-641

Guenette CC, GA Sergy, EH Owens, RC Prince, and K Lee

2003

Experimental design of the Svalbard shoreline field trials.

Spill Sci Tech Bull 8 (3):245-256.

Hill PS, Khelifa A, et al.

2002

Time scale for oil droplet stabilization by mineral particles in turbulent suspensions.

Spill Sci & Tech Bull 8(1):73-81

Hollebone BP

1997

The fate and behavior of oil in freezing environments.

Available on-line

Hume HR, I Buist, D Betts, R Goodman

1983

Arctic marine oil spill research

Cold Regions Sci & Tech

Kepkay PE, Bugden JBC, et al.

2002

Application of ultraviolet fluorescence spectroscopy to monitor oil-mineral aggregate formation

Spill Sci & Tech Bull 8(1):101-108

Khelifa A, Hill PS, et al.

2005

Chapter 7 Assessment of minimum sediment concentration for OMA formation using a Monte Carlo model

Developments in Earth and Environmental Sciences; M. Al-Azab and W. El-Shorbagy, Elsevier. Volume 3:93-104

Khelifa A, Stoffyn-Egli P, et al.

2005

Effects of salinity and clay type on oil-mineral aggregation

Mar Environ Res 59(3):235-254

Khelifa A., Stoffyn-Egli P, et al.

2002

Characteristics of oil droplets stabilized by mineral particles: Effects of oil type and temperature

Spill Sci & Tech Bull 8(1):19-30

Lebedeva EV, Fogden A

2010

Adhesion of Oil to Kaolinite in Water

Environ Sci Tech 44(24):9470-9475

Lee K

2002

Oil-particle interactions in aquatic environments: influence on the transport, fate, effect, and remediation of oil spills

Spill Science and Technology Bulletin

Lee K, Li Z et al.

2008

Wave tank studies on formation and transport of OMA from the chemically dispersed oil

In: Oil Spill Response: A Global Perspective. W. F. Davidson, K. Lee and A. Cogswell. Dordrecht, Springer: 159-177 of 388

Lee K, Lunel T, et al.

1997

Shoreline cleanup by acceleration of clay-oil flocculation processes

Amer Petroleum Inst; Washington DC

Lee K, Stoffyn-Egli P, et al.

2002

The OSSA II pipeline oil spill: Natural mitigation of a riverine oil spill by oil-mineral aggregate formation

Spill Sci Tech Bull 7(3-4):149-154

Lee K, Stoffyn-Egli P, et al.

2003

Oil-mineral aggregate formation on oiled beaches: Natural attenuation and sediment relocation

Spill Sci Tech Bull 8(3):285-296

Li Z, P Kepkay, K Lee, T King, MC Boufadel and AD Venosa

2007

Effects of Chemical Dispersants and Mineral Fines on Dispersion in a Wave Tank Under Breaking Waves

Marine Pollution Bulletin 54(7):983-993

Løseth S, Singsaas I, Sveum P, Brandvik PJ, Jensen H

1994

Oils spill contingency in cold and !rctic areas (ON!) – Status volume I and II

SINTEF report STF60 F94089 (in Norwegian)

Ma X, Cogswel A, et al.

2008

Particle size analysis of dispersed oil and oil-mineral aggregates with an automated ultraviolet epi-fluorescence microscopy system

Environ Tech 29(7):739-748

Mackay D, W Stiver, PA Tebeau.

1983

Testing of crude oils and petroleum products for environmental purposes.

Proceedings at the 1983 Oil Spill Conference, February 28-March 3, San Antonio, Texas, pp. 331-337.

Martin S

1979

A field study of brine drainage and oil entrainment in first-year sea ice

J Glaciology 22:473-502

Muschenheim DK, Lee K

2002

Removal of oil from the sea surface through particulate interactions: Review and prospectus

Spill Sci Tech Bull 8(1):9-18

Niu HB, Li ZK, et al.

2011

Modelling the Transport of Oil-Mineral-Aggregates (OMAs) in the Marine Environment and Assessment of Their Potential Risks

Environ Modeling & Assmt 16(1):61-75

Niu HB, Li ZK, et al.

2010

A Method for Assessing Environmental Risks of Oil-Mineral-Aggregate to Benthic Organisms

Human Ecoll Risk Assmt 16(4):762-782

Omotoso OE, Munoz VA, et al.

2002

Machanisms of crude oil-mineral interactions

Spill Sci Tech Bull 8(1):45-54

Owens EH

2007

Stranded Oil On Shorelines With Ice or Snow – ! Review

Unpublished report prepared for SINTEF

Owens EH

1999

The interaction of fine particles with stranded oil

Pure Appl Chem 71(1):83-93

Owens EH, Bragg JR, et al.

1994

Clay-Oil Flocculation as a Natural Cleansing Process Following Oil Spills .2. Implications of Study Results in Understanding Past Spills and for Future Response Decisions

Seventeenth Arctic and Marine Oil Spill Program (AMOP) Technical Seminar, Vols 1 and 2: 25-37.

Owens EH, Humphrey B, et al.

1994

Natural Cleaning of Oiled Coarse Sediment Shorelines in Arctic and Atlantic Canada

Spill Sci Tech Bull 1(1):37-52

Owens EH, Lee K

2003

Interaction of oil and mineral fines on shorelines: review and assessment

Mar Pollut Bull 47(9-12):397-405

Owens, CK

2004

Regional considerations influencing oil spill response in Arctic offshore environments

Paper at the 2004 Interspill Conference, Trondheim, June 2004

Passow U, Ziervogel K, et al.

2012

Marine snow formation in the aftermath of the Deepwater Horizon oil spill in the Gulf of Mexico

Environ Res Lett 7(3)

Payne JR, Clayton JR, et al.

2003

Oil/suspended particulate material interactions and sedimentation

Spill Sci Tech Bull 8(2):201-221

Payne JR, GD McNabb Jr., and JR Clayton Jr.

1991

Oil-weathering behavior in Arctic environments

Polar Research 10(2):631-662

Payne JR, Hachmeister LE, McNabb Jr GD, Sharpe HE, Smith GS, Manen CA

1991

Brine-induced advection of dissolved aromatic hydrocarbons to Arctic bottom waters

Environ. Sci. Technol. 25:940-951

Reed M, Johansen Ø, Brandvik PJ, Daling PS, Lewis A, Fiocco R, Mackay D, Prentki R

1999

Oil Spill Modeling towards the Close of the 20th Century: Overview of the State of the Art

Spill Sci Tech Bull 5(1):3-16

SL Ross Environmental Research Ltd. and DF Dickins Associated Ltd

1987

Field research spills to investigate the physical and chemical fate of oil in pack ice

Environmental Studies Revolving Funds Report no. 062, 95 p.

Sterling MC, BonneJS, et al.

2005

Application of fractal flocculation and vertical transport model to aquatic sol-sediment systems

Water Res 39(9):1818-1830

Sterling MC, Bonner JS, et al.

2004

Modeling crude oil droplet-sediment aggregation in nearshore waters

Environ Sci Tech 38(17):4627-4634

Sterling MC, Bonner JS, et al.

2004

Characterizing aquatic sediment-oil aggregates using in situ instruments

Mar Pollut Bull 48(5-6):533-542

Stoffyn-Egli P, Lee K

2002

Formation and characterization of oil-mineral aggregates

Spill Sci Tech Bull 8(1):31-44

Sun J, Khelifa A, et al.

2010

A laboratory study on the kinetics of the formation of oil-suspended particulate matter aggregates using the NIST-1941b sediment

Mar Pollut Bull 60(10):1701-1707

Sun J, Zheng XL

2009

A review of oil-suspended particulate matter aggregation-a natural process of cleansing spilled oil in the aquatic environment

J Environ Monit 11(10):1801-1809

Sydnes LK

1991

Oil, water, ice and light

Polar Res 10:609-618

Sydnes LK, Hemmingsen TH, Skare S, Hansen SH, Falk, Petersen IB, Lønning S, Østgaard K

1985

Seasonal variations in weathering and toxicity of crude oil on seawater under Arctic conditions

Environ. Sci. Technol. 19:1076-1081

Thibodeaux LJ, Valsaraj KT, John VT, Papadopoulos KD, Pratt LR & Pesika NS

2011

Marine Oil Fate: Knowledge Gaps, Basic Research, and Development Needs; A Perspective Based on the Deepwater Horizon Spill.

Environmental Engineering Science 28: 87-93.

Thorpe SA

1995

VERTICAL DISPERSION OF OIL DROPLETS IN STRONG WINDS - THE BRAER OIL-SPILL

Mar Pollut Bull 30(11):756-758

Valette-Silver N, Hameedi MJ, et al.

1999

Status of the Contamination in Sediments and Biota from the Western Beaufort Sea (Alaska)

Mar Pollut Bull 38(8):702-722

Vefsnmo S, Jensen H, Singsaas I, Guenette C

1996

Oil spill response in ice infested waters

SINTEF report STF22 F96202

Vefsnmo S, Johannessen BO

1994

Experimental oil spill in the arents Sea – Drift and spread of oil in broken ice

Proceedings, 17th Arctic Marine Oil Spill Program Technical Seminar, Vancouver, BC, Canada, pp. 1331-1343

Wang WZ, Zheng Y, et al.

2011

PIV investigation of oil-mineral interaction for an oil spill application

Chem Engineer Jnl 170(1):241-249

Weise AM, Nalewajko C, et al.

1999

Oil-mineral fine interactions facilitate oil biodegradation in seawater

Environ Tech 20(8):811-824

Wood P, Lunel T, et al.

1997

Clay-oil flocculation during surf washing at the Sea Empress incident

Ottawa, Environment Canada

Zhang HP, Khatibi M, et al.

2010

Investigation of OMA formation and the effect of minerals

Mar Pollut Bull 60(9):1433-1441

Zhou ZZ, Guo LD

2012

Evolution of the optical properties of seawater influenced by the Deepwater Horizon oil spill in the Gulf of Mexico

Environ Res Lett 7(2)

Brandvik PJ, Faksness LG

2009

Weathering processes in Arctic oil spills. Meso-scale field experiments with different ice conditions

Cold Region Sci Tech 55:160-166

Faksness LG, Brandvik PJ

2005

Dissolution of Water Soluble Components from Oil Spills Encapsulated in Ice

Proceedings, 28th Arctic and Marine Oilspill Program Technical Seminar, Calgary, Canada. Vol 1, 59-73

Faksness LG, Brandvik PJ, Daae RL, Leirvik F, Børseth JF

2011

Large-scale oil-in-ice experiment in the Barents Sea: Monitoring of oil in water and MetOcean interactions

Mar Pollut Bull 62:976-984

Guyomarch J, Le Floch S, Merlin F

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 Sci Tech Bull 8(1):95-100

NORCOR

1975

The interaction of crude oil with Arctic sea ice

Beaufort Sea Project NORCOR Engineering & Research Ltd, technical report no. 27, Canada Department of the Environment, Victor

Singsaas I, Brandvik PJ, Daling PS, Reed M, Lewis A

1994

Fate and behavior of oils spilled in the presence of ice – a comparison of the results from a recent laboratory, meso-scale flume and field tests

Proceedings, 17th Arctic Marine Oil Spill Program Technical Seminar, Vancouver, BC, Canada, pp. 355-370

Faksness LG, Brandvik PJ

2008b

Distribution of water soluble components from oil encapsulated in Arctic sea ice: Summary of three field seasons

Cold Regions Sci.Technol. 54: 114-116

Bragg, JR, Owens EH, et al

1994

Clay-Oil Flocculation as a Natural Cleansing Process Following Oil Spills .1. Studies of Shoreline Sediments and Residues from Past Spills

Seventeenth Arctic and Marine Oil Spill Program (AMOP) Technical Seminar, Vols 1 and 2: 1-23

Brandvik PJ, Faksness L-G, Daling PS, Singsaas I

2005

Fate and behaviour of oil spills under Arctic conditions. Earlier results compared with new field experiments on Svalbard

Paper presented at the AMAP International Symposium on Oil and Gas Activities in the Arctic, St. Petersburg, 13-15 September 200

Faksness L-G

2008

Weathering of oil under Arctic Conditions.

PhD dissertion at the University of Bergen, Norway

Faksness LG, Brandvik PJ

2008a

Distribution of water soluble components from !rctic marine oil spills – ! combined laboratory and field study

Cold Regions Sci.Technol. 54: 97-105

Faksness, L.-G., Brakstad, O.G., Reed, M., Leirvik, F., Brandvik, P.J., Petrich, C

2011

Oil in ice: Transport, Fate and Potential Exposure

Final Report Submitted to The Coastal Response Research Center. SINTEF Report no. A19275, ISBN 978-82-14-05136-0. Unrestricted

Lee K, Weise AM, St-Pierre S

1996

Enhanced oil biodegradation with mineral fine interaction

Spill Sci. Technol. 3: 263-267

Owens EH, Sergy GA, Guenette CC, Prince RC, Lee K

2003

The reduction of stranded oil in in-situ shoreline treatment options

Spill Sci Technol Bull 8:257-272

Sergy GA, Guenette CC, Owens EH, Prince RC, Lee K

2003

In situ treatment of oiled sediment shorelines

Spill Sci Technol Bull 8:237-244

Peterson CH, Rice SD, Short JW, Esler D, Bodkin JL, Ballachey BE, Irons DB.

2003

Long-term ecosystem response to the Exxon Valdez oil spill

Science 302, 2082-2086.

Brandvik PJ, Daling PS, Myrhaug JL

2010

Mapping weathering properties as a function of ice conditions: A combined approach using a flume basin verified by large-scale field experiment

Proceedings, Thirty-third AMOP Seminar on Environmental Contamination and Response, Environment Canada, Ottawa, ON, pp. 7

4.0 OIL SPILL RESPONSE STRATEGIES

Executive Summary

Some description

Some description

Some description

Some description

Photos 4-1 – 4-4. OSR deployment (NOAA)

Four main oil spill response (OSR) strategies exist: natural attenuation, mechanical recovery and containment, in-situ burning, and physical and chemical dispersion of oil. All four are often used in combination and can be used in the Arctic. Selection and application of an oil spill response strategy should be based on both the effective removal of oil for the specific oil and weather conditions and consideration of the information on potential impacts to valuable ecosystem components (VECs) since application of response options will influence the fate of oil in the environment and concomitantly potentially alter the impact to different VECs. The influence of chemically and/or physically dispersed oil on pelagic species is well documented, but biological responses to oil at interfaces (air/water, ice/water, sediment/water and shorelines) has been less documented. Understanding the consequences of OSR actions on the impacts and resilience of VECs within these interface layers needs to be further developed in order to strengthen our ability to select a preferred OSR strategy under each spill scenario using the net environmental benefit analysis (NEBA) process (see Section 9).

Environmental effects related to OSR options have been studied extensively. In order to readily synthesize information that is already available on exposure potential, sensitivity and resilience of VECs, this information should be collated for each OSR technology and corresponding VECs that these technologies potentially impact to improve application of NEBA processes in the Arctic. This compilation of technical data will also facilitate the identification of remaining uncertainties.

4.1 Introduction

This chapter explains the characteristics of the main OSR technologies and summarizes the current knowledge on the potential environmental impacts of the various OSR options to an oil spill in the Arctic region.  The focus is on the impact of an at-sea response (as opposed to shoreline clean-up) to support the NEBA process.  The response techniques discussed are: natural attenuation, mechanical recovery and containment, in-situ burning (ISB, with and without the use of herders), and physical and chemical dispersion.   The equipment and strategies used for these techniques have been reviewed in detail in a recent publication (Potter et al. 2012).  Strategies and techniques for recovering and remediating oil in ice have been extensively studied over the past 45 years (Dickins 2004). Two oil spill research projects conducted in the Canadian Beaufort Sea from 1974 to 1981 contributed to the acceptance of in-situ burning as a primary response strategy to deal with spills in ice (Norcor 1975; Dickins and Buist 1981).  Examples of previous field experimental studies and accomplishments are presented in Table 1-3 (next page).  Of note, is a recent program, oil spill contingency for Arctic and ice-covered waters (JIP on Oil in Ice; Sorstrom et al. 2010) conducted several large scale field experiments assessing the complex interactions of oil, water, and ice with state of the art data collection techniques (Photo 4-5). 

Photo 4-5. Field experiment during Oil-in-Ice JIP program (SINTEF)
Photo 4-5. Field experiment during Oil-in-Ice JIP program (SINTEF)

This review focuses on the effectiveness and associated environmental impacts of these response options when deployed under a variety of Arctic conditions and identifies areas for future research.  Special attention is paid to how responses in the Arctic may differ from non-Arctic areas of the world. Increasing collections of experience and data will provide a broader foundation for spill response decisions and help to reduce the overall environmental consequences of a spill.

The main sources of information were technical documents and incident reports supplied by CEDRE’s library, NewFields, and documents available at the Norwegian Environment Agency within the Norwegian Ministry of Climate and Environment.

These documents were categorized into:

  1. Review or synthesis documents dealing with issues of Oil Spill management in Arctic, or with research needs on the same topic
  2. Experimental studies of OSR actions that targeted the effects of those responses in cold environments
  3. Recorded case stories of incidents relevant to assessing impacts of response technologies or Arctic geographies.

Table 4-1.  Landmark Field Studies

ProjectYearLocationSpill CharacteristicsIce TypeProcess or ResponseAchievementsAuthors

Interaction of crude oil with Arctic sea ice

1975

Canadian Beaufort

340 bbl/ 9 spills

Under fast ice

Burning & Mechanical

Oil injected under ice, Oct – May. Oil spreading and entrainment documented by divers and video. Assessed migration, degree of weathering and dissolution. Min oil thickness was 0.8 cm at ice-water interface. Also assessed currents under solid ice.

Norcor 1975

Oil and Gas Under Ice

1979/80

Canadian Beaufort

116 bbl/ 3 spills

Landfast Ice

Subsea Blow-out

Determined  area impacted by a subsea blowout; assessment of in-situ burning as spill response option

Dickins et al. 1981

Baffin Island Oil Spill Project

1980

Baffin Island

2-8 m3

Seasonal

In-situ burning; natural recovery of shoreline

5 core studies; field based trials w 3 similar bays w untreated oil, treated oil, and control (included biological, chemical, and physical measurements; burning in melt pools; shoreline study w/ oil and emulsion in intertidal zone

Blackall and Sergy 1987

Behavior of crude oil in pack ice

1986

Canadian East Coast

18 bbl/ 3 spills

Pack ice, leads, between floes

Burning

The presence of ice dramatically reduced the spreading of oil compared to open water.

Buist IA, Dickins DF 1987

Emulsions in Ice

1983

 

192 L

   

Emulsions were stable:

did not inhibit ice formation; retarded natural migration through ice sheet

Buist et al. 1983

Marginal Ice Zone

1993

Norway; Barents Sea

164 bbl

Marginal between floes

Oil Tracking in Ice

Applied oil spill trajectory model (OILMAP™) to forecast trajectories of oil in pack ice.  When wind was off-ice wind speed drift (2.5%) with Ekman veering angle of 35°; with on-ice wind drift was 1.5%, veering angle of 60°

SINTEF; Singsaas, Brandvik, Daling

Under-Ice Spill

2001

Sea of Okhotsk

Light oils

Under ice floes

Vertical Migration

Oil fills under-ice cavities;  very small amount migrates to the surface (<1% with 7 -10 cm rise)

Ohtsuka et al.1999; Ohtsuka et al. 2001; Karlsson et al. 2011

Diesel and Fuel Oil

2005

Russian Arctic

Light Oils

Ice floe surface

Evaporation

Complete evaporation during spring and summer; photo-oxidation more significant with 24 h daylight

Serova 1992; Ivanov et al.2005

Encapsulated Oil

2005

Svalbard

6 Crude Oils

Pack ice

Dissolution

Water soluble compounds diffused through 110 cm thick ice, but concentrations were low (6 ppb)

Faksness and Brandvik 2005

Predictive Modeling Algorithms

2004-2009

Canada, Ohmsett NJ

Alaskan crude oils

Cold water, Ice

 

Thickness of oil on calm water;

Spreading of oil; Equilibrium of oil thickness on ice; Oil Spreading on ice, on snow; Evaporation on ice, under snow, among drift ice

Buist et al. 2008; Buist et al. 2009

Weathering of Oil in Ice

2009

Svalbard

Statfjord crude

0, 30, 90% coverage 

Natural Weathering

Oil in 90% ice had > slick thickness,  reduced evaporation; wave dampening reduced emulsification; oil ignitable and chemically dispersible after 60 h

Brandvik and Faksness 2009

 

In general, the majority of the reviewed documents dealt with the feasibility of the response options under various conditions and the methods that were used or tested to optimize the efficiency of response techniques.  It is noted that very few papers focus on a comparison of effects between different environmental compartments as a result of the response action while this is one of the main challenges in NEBA.  Next to the potential impacts of oil and oil residues, NEBA should also look into the impacts of the response action itself, which may include access to vulnerable areas and impacts related to logistics.

4.1.1 Environmental Uniqueness of the Arctic Region in Relation to OSR

The Arctic region is characterized by the presence or absence of ice at different times of the year which generates a seasonal regime of ice coverage and ice thickness (AMAP 2007, SL Ross 2012b, Wang et al. 1999).  When spilled, the oil fate and behavior will depend on a combination of the ice growth stage and coverage area which together can reduce or stop the weathering processes and limit the spreading.  In some cases, the oil can be trapped by the ice and remain fresh and unweathered until the melting season. 

The extent and nature of ice coverage also greatly affects the OSR options that might be considered and their operational effectiveness (SL Ross 2011).  While shorelines and lagoons are not unique to the Arctic these are critical environmental compartments that have seasonal congregations of valuable ecosystem components that create significant risks to populations (refer to Sections 2, 7 and 8).  The importance of these seasonally used compartments and the logistical challenges of positioning manpower and equipment at remote locations that ultimately limit response feasibility and/or timing suggest that response options need to be selected to avoid contact with these environmental compartments during those seasonal uses (refer to Sections 7, 8 and 9).  Surface or subsurface oil releases will be influenced by the presence of ice (refer to Section 3) and will impact both environmental compartments and key species groups differently depending on location and density of the ice cover and seasonal use of those environmental compartments by VECs (refer to Section 2).   Depending on the season of the year and the site of released oil, an incident may result in different types or magnitudes of environmental impacts.  Generally speaking, “Arctic habitats are characterized by extreme seasonal change, which drives extensive migrations on land and at sea. The seasonal patterns of movement to, from, and within the Arctic determine to a large extent the vulnerability of Arctic ecosystems to oil spills. These patterns of seasonal activity and movement must be taken into account in selecting response strategies designed to reduce or avoid environmental impacts from oil and gas activities” (AMAP 2007). 

Arctic species are subject to potentially dense concentrations (e.g. for birds, mammals, and fish) according to migrations patterns and corresponding ice regime.  Ice edges are important locations for concentrating marine mammals and birds as a result of high biological activity. In addition, possibly due to the low ambient temperature, there is a low turnover for many species and consequently population recovery can be slow.  As an example, some Arctic fish spawn under ice in winter; their eggs incubate under ice and hatch when ice begins to melt and plankton blooms occur.  An oil spill in such spawning areas and during early life stages could severely reduce that year’s recruitment (AMAP 2007).  The presence and the movements of the different species in both localized and regional areas is generally not  known to the extent needed to make some of the important tradeoff decisions when selecting spill response options.  Baseline studies are often conducted as part of the Environmental Impact Assessment (EIA) process that is undertaken in areas where oil and gas production is envisaged or in the process of being developed. Mapping (spatial and temporal) the ecological vulnerability of Arctic habitats in these areas will contribute significantly to regional oil spill response planning. 

4.2 Knowledge Status - Impact of OSRs

4.2.1 Natural Attentuation

Several field tests with experimentally released oil have been completed in the Arctic (see SL Ross, 2012 for an overview).  However, except for the Baffin Island Oil Spill (BIOS) experiment these tests were devoted to studying the behavior and environmental fate of the oil in icy conditions and not the environmental impact at the spill location or other ECs.  Nonetheless, this information helps frame the physical and chemical factors that affect the nature of oil released in an Arctic environment and help to identify the challenges to be encountered in implementation of any type of OSR activity.  In fact, this sets the basic expectations for what might occur if natural attenuation were the only response option implemented. 

The specific challenge encountered in the Arctic during OSR is the presence of permanent or seasonal ice, which has many consequences (Potter et al. 2012).   Ice reduces the sea surface agitation which coupled with the low prevailing temperature, slows down the spreading of the oil slick reducing physical weathering and emulsification that occurs with more active surface water disruption.  Ice also limits oil spreading when it is between ice blocks, or when beneath or on top of the ice.  This helps keep the oil relatively concentrated reducing the rate of oil weathering.  Large quantities of oil can be trapped either in snow, or on or under the ice within spaces found in the unevenness of the ice surface.  When ice is forming, oil can be encapsulated in the new ice and thus kept unaltered during the winter season.  Oil trapped in ice can then be released to surrounding waters when the ice melts, possibly reappearing as fresh, unweathered oil.  Oil has been observed to migrate through the ice, at a rate that is dependent on oil viscosity.  Experimental studies have determined that oil components separate within the ice and undergo degradation during the winter in the ice brine channels, as discussed below and in Section 5.

4.2.1.1 Potential Environmental Impact of Untreated Oil

If spilled oil is not recovered or treated, heavier oils may persist on the surface of water or surface of ice and can affect biological communities that are utilizing these interfaces, especially birds or mammals with fur (due to the potential loss of thermal insulation) whereas lighter oils may naturally disperse into the water column.  A reduced rate of oil weathering may occur if oil is encapsulated by ice and the oil may not be biologically available until the spring thaw.    Another unique attribute of the Arctic is that oil can strand on the shore during the ice-free season, whereas at other times the shoreline may be protected by landfast ice, which prevents oil from coming ashore.  Oil stranding on shoreline substrates during the ice free period is subjected to the strong erosional forces of the ice on shoreline substrates during the next ice build-up and ice break-up seasons.  Only deeply buried oil that might occur in intertidal cobble fields would be sequestered for extended periods of time (e.g., decades).  Therefore in many cases the persistence of oil onshore will be governed by the physical erosional forces that occur in many shorelines which minimize its retention.  However, the weathering and recovery processes may be longer (e.g. occupy those area for more than one year) in cases where oil is sequestered in spaces among cobble or boulder fields or where oil may be trapped in isolated nearshore water bodies.  The practicalities of staging recovery operations in remote locations are also a consideration.

The effect of oil that is left to natural attenuation on the shoreline depends upon:

  • The environmental resources of concern that are present when the oiling occurs and during subsequent seasons when extended oil exposures could be possible.
  • The duration certain ECs of shoreline contamination, which may not persist for more than one year in while in others it may be extended for longer periods.
  • The resiliency of the populations of various species that were impacted as a result of the spill and treatment methods that were used.  The resiliency of these populations and communities of organisms is controlled by fecundity, immigration from unaffected areas, and the diversity of organisms present within the affected habitats.

McAuliffe et al. 1980 reported on the effects of oil on under-ice meiofauna as a part of the BIOS project (McAuliffe et al. 1980).  Effects of this experimental spill on ice algae are summarized in a report by SL Ross (2010).  In that study, the bottom 10 cm of ice had decreased density of meiofauna, no adverse effects were observed on the ice algal community, and under-ice invertebrates showed no mortality but did drift away from the oil impacted area for days following the spill.  All the findings of the BIOS Project are summarized in Li et al. (1992).

A separate study conducted on first-year sea ice off Svalbard showed that there is a migration of bioavailable water soluble components (WSC) from encapsulated oil through the ice to the underlying water (Dickins et al. 2006, Faksness et al. 2012).  The estimated toxicity of these dissolved oil components in the ice was calculated using toxic units and the findings indicated that the concentration of WSC in the brine channels might be acutely toxic to the ice fauna.   Results from another field study of an experimental release of 7000 L crude oil in the Barents Sea showed low concentrations of dissolved hydrocarbons (maximum concentrations were 4 ppb dissolved hydrocarbons and 32 ppb total hydrocarbons) in subsurface water.  Predicted toxicity to the exposed community in the upper layers of the water, expressed as toxic units, was 0.11 or less, indicating that the potential for acute toxicity was low in subsurface bulk water (Faksness et al. 2012).  However, the effects of surface oil on organisms using the surface layer, polynyas, ice-edges or adjacent shorelines where oil compounds can be re-concentrated was not assessed.

These studies indicate that there could be effects on the local ice biota if the oil is encapsulated in the ice or trapped underneath the ice. The organisms associated directly with the ice could be exposed to potentially toxic dissolved hydrocarbons over the course of several months, causing potentially toxic oil components to enter the Arctic marine food web. On the other hand, the measured concentrations of dissolved hydrocarbons in the water column, or underneath an untreated oil slick, have been lower than potentially toxic concentrations, perhaps indicating that severe effects to organisms residing in the water column would be negligible.  However, this does not take into account the reconcentration processes that occur at interfaces such as the surface of the water, at ice water interfaces, convergence zones and shorelines (refer to Sections 2 and 3).

4.2.1.2 Conclusions on Natural Attenuation

Oil remaining in Arctic habitats without treatment will behave similarly to non-Arctic situations, although chemical processes such as dissolution, volatilization, and biodegradation may occur at a slower rate resulting in increased persistence.  In non-ice periods oil spills on the sea surface will remain at the sea surface and be transported in slicks by winds and currents to shorelines, convergence zones, and offshore surface waters.  During that process some of the oil will dissolve into the water column or be physically dispersed into the water column as droplets, some will volatilize into the atmosphere, while the majority of the oil may remain on the surface where it will weather, biodegrade, emulsify and accumulate in zones of reconcentration.  Subsurface releases of untreated oil will generally rise towards the sea surface but during that transport it may also be rapidly biodegraded based on the increased surface area of oil droplets created by the turbulence of the release.  Oil that remains on the sea surface can be stranded on shorelines or concentrate in convergence zones but the oil may also be encapsulated by ice as it forms.  In order to facilitate the forecasting of the seasonal dynamics of oil in these compartments, it is important that data are available for NEBA evaluations. This will facilitate the decision making process regarding the most appropriate response option under various conditions. 

Such a NEBA process would evaluate the trade-offs of untreated oil containment by ice and treatment efficiency with decreased impact on pelagic environments by dispersant treated oil in non-ice environments.  In addition the increased effects of surfaced oil as it is captured and released by formation and melting of ice on seabirds, marine mammals, annual ice fauna and flora should be evaluated. Also the biological significance of overwintered oil and ice must be determined.

Many data that serves as a basis for such evaluations is already available, but improvement of the information base would result in further reduction of uncertainties. Suggested topics for such studies are: 

  1. Biodegradation
    • Measure the biodegradation of oil in ice and trapped within leads or under ice over a winter season.  Compare to biodegradation of oil in pelagic waters and surface layers during non-ice periods.
    • Does frazzle ice increase biodegradation of oil released from ice by physical grinding and disturbance of oil, creating larger surface area for microbes to degrade the oil?
  2. 2. Presence of VECs
    • Determine avoidance behavior for fish and invertebrate VEC’s associated with oil trapped with ice.  Indications are that they will move away from oil.
    • Evaluate the use of polynyas or leads by VEC fish, invertebrates, sea birds, and marine mammals and the potential for oil effects in these critical habitats. Compare oil within broken ice fields and open waters as an attraction to seabirds, marine mammals, fish and invertebrates.
    • Summarize the same types of information for seabirds, shorebirds, marine mammals. 
  3. Considering the uniqueness of Arctic shorelines influenced by landfast ice, it will be important to understand the dynamic processes controlling the fate and persistence of oil on such shorelines.  This will require an assessment of the potential for lingering oil releases and the assessment of the natural decontamination rate resulting from different responses

4.2.2 Mechanical Recovery and Containment

When oil is spilled on the surface of the water or rises from a deep water discharge and then accumulates on the surface it is possible to concentrate the oil by placement of booms in the pathway of the oil transport.  As the oil accumulates next to the booms it can be recovered by pumping the captured oil into collection containers.  Oil can also be collected e.g. using boats and surface water booms that accumulate the oil as the vessels travel through oil slicks.  The success of these processes depends on the encounter rate and efficacy of the mechanical collection techniques and the success of containment or accumulation.  Ice can act as a natural boom that allows oil to collect along its edges, within leads, under the ice in pockets, and within polynyas.  Ice can also hold the oil for extended periods of time, allowing mechanical recovery to occur over more extended periods of time from its formation to when it begins to melt.  Many of the tools used for mechanical recovery are not unique to application in the Arctic but for some adaptations have been made to collect oil mixed with ice (Broje and Keller 2007).   The tools used to mechanically recover oil after it is concentrated are described in more detail in the Artic response technology report by SL Ross (2012). 

4.2.2.1 Environmental impacts from Mechanical Recovery and Containment

Moving ice, either as ice floes or frazzle ice can interfere with containment and recovery equipment deployment and operations (Potter et al. 2012, EPPR 1998).  On the other hand, ice can also slow the spreading of oil on water, keeping a slick thicker during recovery, which increases the efficiency of this type of response activity.  Environmental impacts of mechanical recovery are usually considered in terms of emissions of response equipment, noise, and the impacts of the presence of large numbers of personnel.  However, it is important to consider that mechanical containment and recovery is a slow, tedious, and challenging response method. Mobilizing and supporting such activities in remote areas adds further inefficiencies and time constraints.  Impacts from a containment and recovery response effort are:  the impact from oil that is left behind (oil that escapes containment), and impacts from the activities necessary to reclaim or dispose of the recovered oil and associated oily debris.

The impacts considered with natural attenuation are also associated with the residual oil left behind from mechanical recovery.  Historically, mechanical recovery in open water spills is often reported as less than 15% of the spill volume and in most cases less than 5%, although specific performance can vary widely from incident to incident (EPPR 1998). Thus, for this report, impacts considered under MNR will also be a large part of any mechanical containment response scenario.  Our consideration of the ISB, dispersants and OMA and herder technologies will therefore compare tradeoffs using either mechanical response or naturally attenuation as their baselines.

4.2.2.2 Conclusions

Mechanical recovery in an Arctic spill situation may have marginal improvements in effectiveness due to the presences of some types of ice conditions, or may have additional inefficiencies brought on by different types of ice.  Impacts of residual oil left in the environment due to the low effectiveness of mechanical recovery can also serve as the baseline assessments for evaluating tradeoffs for ISB (with or without herders), chemical and physical dispersants. The areas of proposed work are the same as those included in the natural attenuation section of this report. 

4.2.3 In-Situ Burning and Chemical Herders

The in-situ burning (ISB), also referred to as controlled burning, has been the subject of extensive research, development, and testing over the past 30 years in temperate and sub-arctic water (Potter et al. 2012).  The basic premise for effective and efficient burns is to collect and/or concentrate the oil slick to a thickness greater than 2 mm and provide an ignition source that can start the burning of surfaced oil.  The oil must not be weathered or emulsified to such an extent that there are not enough lower molecular weight compounds (LMW) and available oil to sustain combustion (Fingas and Punt 2000).

Photo 4-6. In-situ Burning (Liv-Guri Faksness)
Photo 4-6. In-situ Burning (Liv-Guri Faksness)

ISB can be effective in rapidly removing large quantities of oil from the marine environment. Ideally, about 85 to 95% of the burned oil becomes carbon dioxide and water. The rest, 5 to 15% which is not burned efficiently is converted into particulates (soot) and a few percent is converted into organic compounds and combustion products that remain in the marine environment (Potter et al. 2012). The burn residue from a typical efficient ISB operation is in the order of less than 15% (SL Ross 2010).  ISB seems well suited to Arctic conditions and the presence of ice (Photo 4-5). The presence of significant ice formations can keep  oil from reaching water (burning of oil on ice) or limit the spreading of the oil on water (burning thick patches of oil on water contained among ice formations).

In-situ burning is an efficient process that removes ~80% of the oil (SL Ross, 2010).  However, the residuals of these burns include unburned volatile materials that are released into the air, soot particles that are also mobilized and transported into the air, and the modification of oil compounds into new products or the addition of agents (e.g., herders) and their burn products for release into the air or water.  These residues of the burning process (smoke, volatiles, soot particles, additives and unburnt oil) are the potential materials that can pose environmental and human health effects.  In addition there is a small risk of causing secondary fires that could threaten human life, property and natural resources; this risk is, however, easily manageable.  

Chemical herding agents are products used for thickening an oil slick and concentrating oil on the water surface in order to reverse the effects of spreading.  The increase in thickness may facilitate oil combustion during an in-situ burning operation.  Several herding agents are listed on the National Contingency Plan registry as approved for use during oil spills, including Thickslick 6535 and Siltech OP-40.  A third herder employed by the US Navy (USN herder) has been tested under arctic conditions.  The USN herders (65% Span-20 and 35% 2-ethyl 1-butanol) and silicon based herders have been used under Arctic conditions and have been shown to work in cold open water environments as well as in broken ice (Buist and Nedwed 2011). 

4.2.3.1 Potential environmental and human health effects of ISB residues and unburnt oil

Generally an efficient burn leaves 5 to 15% of the initial oil as residual or unburnt oil (SL Ross 2010).  The residual is mainly composed of high molecular weight (HMW) oil compounds which are similar to those in highly weathered heavy fuel oil.  The physical properties of burn residues depend on burn efficiency and type of oil.  Factors that determine whether residues float or will sink are:  water density, oil chemical properties, thickness of slick, and efficiency of the burn (Buist and Trudel 1995).  The residual ash may also settle on the surface of the surrounding ice or sea where it may come into contact with surface dwelling organisms. 

Tests have been carried out on the burn residue of Alberta Sweet Mixed Blend which had been used in the Newfoundland Offshore Burn Experiment (NOBE).  The water accommodated fraction (WAF) was prepared from the unburnt residue and tested on rainbow trout to assess the 96 h LC50 and on sea urchin for inhibition of fertilization (20 min contact). The maximum total petroleum hydrocarbon (TPH) concentration measured in the test solutions (WAF) was 1.1 mg/L.  All samples were not toxic to the tested species (Blenkinsopp et al. 1997).  In another study, Daykin et al. (1994) concluded the toxicity of the residue should be lower than that of the initial oil.  Other studies showed that the residue had very little or no acute toxicity to key indicator species in salt water and freshwater because an effective burn removes the lightest, most toxic components of crude oil (Blenkinsopp et al. 1997).

In a recent study by Faksness et al. (2010), the semi-volatile organic compounds (SVOCs) in a crude oil prior to and after ISB were analyzed.  No volatile analyses were performed, but a removal of approximately 60% of the SVOCs, mainly the decalines and naphthalenes, had occurred during the ISB. Acute toxicity tests with the marine copepod Calanus finmarchicus exposed to the underlying water after ISB indicated no increase in toxicity when compared to WAF generated with unburned weathered oil. These findings were in accordance with the results presented by Daykin et al. (1994) as a part of NOBE.  Concerning potential for exposures to the PAHs in burned oil residue, several studies have demonstrated that the concentrations of PAHs in the residue were lower than that in the initial oil.

While the toxicity from uptake of chemical components of the residue does not appear to be a concern to water column organisms, there is possible impact on surface dwelling species (by ingestion and/or direct smothering) and benthic species if the residues were to be stranded onshore or sink onto the sea bed. These risks are however considerably lower than when fresh oil remains on the surface or strands.  Concern for such an impact arose during oil spill incidents involving ISB, e.g. the Honan Jade spill in South Korea (1983) and the Haven spill in Italy (1991) where the NRC (2005) reported that the burn residue was typically a semisolid, tar-like layer.  The surface oils were burnt, removing the impacts in the surface waters but the burnt residues would sink where they affected the benthos.  Such sinking residues consisted of scattered chunks rather than as a continuous mat covering a broad area.   While this type of impact on the benthos is very hard to assess it must be considered as part of an overall assessment of potential benefits and impacts of ISB.

Oil combustion produces gas, smoke and soot into the atmosphere.  Typically the smoke plume is composed of CO2, steam, soot, CO, SO2, NOand VOCs including PAH and BTEX, dioxins and dibenzofuran (Tennyson 1994, Fingas et al. 1993).  Despite the fact that VOC concentrations in the plume are usually lower than the accepted threshold value for human health concerns (Buist et al. 1999), responders put an exclusion zone in place to ensure there is limited exposure of downwind communities or wildlife populations to these compounds.  Responders are also excluded from the immediate area of the fire and at more lengthy distances downwind of the fire. 

In one assessment of the NOBE study Ross et al. (1996) found that burning of 1 Kg of oil produced 40 µg of PAH in the soot/particles while the original oil had 9.5 g/kg of oil.  Therefore multiple authors have concluded the ISB residue would have a lower toxicity than the initial oil (Buist et al. 1999, Fraser et al. 1993, Garrett et al. 2000, Li et al. 1992, Lin et al. 2005).  The potential effects that would occur with species living at the air/water interface or that break through the surface (e.g. seabirds and marine mammals) were unaddressed by these studies. 

The production of smoke during an ISB, and the concentrations of smoke particles at ground or sea levels are usually of most concern to the public as they are highly visible from significant distances and can persist for several miles downwind of a burn.  Concerns include human health and wildlife inhalation risks from particulates carried in the smoke plume.  Particulate concentrations in the plume are greatest at the burn site, but they decline with increasing distance from the site, primarily through dilution, dispersion, and fallout, but also through washing out by rain and snow (API 2004). The species of greatest concern to atmospheric pollution or fallout of soot and particles will be downstream of the burn and include marine mammals that must breathe at the surface of the sea or species and life stages that live within the very surface of the water.

4.2.3.2 Environmental Impact of Herders

The literature dealing with herders is rather old as these products have not been considered or promoted until very recently.  Available data shows that most chemical herding agents are not soluble either in water or in oil (less than 1%; Hayward et al. 1995, MSRC 1993) and are used at low application rates; therefore, acute toxicity of these products to pelagic organisms is generally not considered to be an issue.  Additional considerations may be required under special use conditions such as very shallow waters with low flushing rates and organisms abundant in early life stages, but exposures are still likely to be at very low concentrations (Walker et al. 1999, MSRC 1993). The more recent documents deal with the efficiency or operations related to herder use (SL Ross 2010).  As with dispersants, the toxicity information is limited to water column organisms and there is little information available on the toxicity of these materials to surface dwelling organisms.

4.2.3.3 Conclusions on ISB and Herders

Most of the available information on ISB deals with the efficiency of the technique and operational guidance.  While there is some information regarding monitoring activities to ensure air pollution is not a human health or wildlife exposure issue, information on the environmental impact is more limited.  Information on fate and effects studies could be compiled, taking into account the specificity of Arctic environment (species with the possibility of large concentrations of juvenile life stages, especially within the surface microlayer) and additional studies could be conducted to evaluate the persistence of these products and their residues where necessary.

4.2.4 Improving Dispersion of Oil

Chemical dispersants are most effective when applied during or quickly after a spill or sub-sea release event, before dilution, weathering and emulsification of the oil reduces the effectiveness of the dispersants.  Modern dispersants are mixtures of solvents consisting of organic carbon chains that are oleophilic and surfactants that are hydrophilic.  The combination of oleophilic and hydrophilic components change the surface viscosity of the oils and create small droplets of oil that are released from the surface water and move into the water column or from deep water releases into adjacent deep pelagic environments.  These small oil droplets have greatly increased surface area that increases the rate of microbial degradation compared to the oil prior to dispersant application.  Dispersant mixtures have been evaluated by numerous organizations to determine their toxicity and efficiency of dispersion under many different environmental conditions.  By breaking up the oil and creating micron-sized droplets, chemical dispersion reduces the persistence of a surface slick or the potential for sub-sea discharges to reach the surface and thereby minimizes potential encounters by marine mammals and offshore bird populations. Application of chemical dispersant to sub-surface and to surface oil slicks reduces the amount of oil that becomes stranded on the shoreline and prevents oil from transforming into weathered oil-in-water emulsions that are resistant to further biodegradation (Lewis and Daling 2001). 

However, dispersing oil into the water column from surface slicks or deep water releases is most effective when the oil is fresh and unweathered.  Mitigating damage to the shoreline and to organisms that may encounter surface slicks means exposing the near surface and shallow or deep pelagic communities to elevated concentrations of dispersed oil for short periods of time. 

The literature reviewed focused on dispersant applied at the sea surface and it contains information on the toxicity of oil chemically dispersed into the water column and effects on those marine organisms from laboratory studies.  Information on the behavior of dispersants applied below the upper surface layer during blow-out scenarios were sparse during this review period.  However, it is expected that post-spill studies of the recent Macondo well blow-out and explosion that occurred in 2010 in the Gulf of Mexico will contribute greatly to our knowledge of subsea application of dispersants as well as natural dissolution and biodegradation processes that can occur in the deep ocean environment and at very cold temperatures that are similar to arctic temperatures.

There are limited holistic assessments that combine the toxicity of chemically dispersed oil data with information on specific assessments on toxic impacts associated with Arctic communities or the comparative damage resulting if oil persists on the surface or comes ashore.  Therefore, the rationale for application of chemical dispersants should be based on the comparison of the possible extent and duration of impacts to the organisms living in the water column (e.g. fish, shellfish, plankton, etc.) resulting from the use of dispersants, and the extent and duration of potential damage which would result if dispersants were not used, i.e. from a persistent surface slick (e.g. effects to birds, mammals, and fish and invertebrates that live in the very surface of the water) and from the stranding of the weathered oil on the shoreline (e.g. effects to coastal shoreline species and benthic organisms).  This type of information must be factored into the tradeoffs associated with Arctic dispersant use, and is considered in the section “Monitoring Natural Recovery (no active response)”.

4.2.4.1 Impact of Chemically Dispersed Oil

Most toxicity studies evaluate the impact of increasing the exposure of pelagic organisms to oil as a result of dispersing the material into the water column.  Considering the toxicity toward water column organisms, it is recognized that the observed toxicity effects from chemically dispersed oil is due to the effects of the increased quantity of dispersed oil into the water and are not caused by­­­­ the dispersant itself, as modern dispersant formulations are much less toxic than oils (Hemmer et al. 2011).

In assessing dispersed oil toxicity, determinants of adverse effects for a given species are exposure concentration and duration of exposure (see also a more detailed review of peer-reviewed literature presented in Section 6, Ecotoxicology of Oil and Treated Oil).  A review of field studies found that small-scale field tests have demonstrated that the concentration of dispersant in water falls to less than 1 mg/L within hours (NRC 2005).  The available data suggest that in general, maximum dispersed oil concentrations after a spill are less than 50 mg/L immediately after dispersion into the upper water column (top 3 m) and that dispersed oil concentrations dilute rapidly, dropping to 1 to 2 mg/L in less than 2 h throughout the water column (Cormack and Nichols 1977, Daling and Indrebo 1996, McAuliffe et al.1980).  These low concentrations are generally below estimated toxicity threshold concentrations derived from exposure experiments for most common water column organisms (McFarlin et al. 2011, Gardiner et al. 2013). 

The BIOS experiment conducted in sub-Arctic nearshore areas in the 1970s studied oil dispersion impact on nearshore environments and concluded that the results offer no compelling ecological reasons to prohibit the use of chemical dispersants on oil slicks in nearshore areas (Potter et al. 2012).  Secondly, the results provide no strong ecological reasons to undertake an intrusive effort to cleanup stranded oil (on certain shoreline types).

During an experimental oil spill in the Barents Sea in 2009, 2000 L of crude oil were dispersed six hours after release (Potter et al. 2012). Two hours later, measurements of oil in water were performed at depths of 1, 2 and 3 m. The maximum concentration of oil in water was measured to 5.5 ppm (at 2 m depth) with an oil droplet size smaller than 10 µm, 30 minutes after mixing energy was added by the ship thrusters.  The monitoring indicated background concentrations were restored shortly after these measurements, as the plume had most likely drifted and diluted with the currents (Merlin and Le Floch 2012).  After the Sea Empress incident, a major spill in nearshore waters at the port of Milford Haven, UK, dispersed oil concentrations were monitored and quantified in the field.  Results showed 10 ppm dispersed oil immediately after the dispersant application, decreasing to 1 ppm 2 days after, 0.5 ppm 1 week after and 2 ppb 1 month after (SEEEC 1998). 

Such a decrease can be modeled with the following relationship:

                                   C = C0 e-1.35                                                                                                 Equation 1

Where C equals oil concentration at time (t in hours);

C0 is the initial concentration;

e-1.35 represents an exponential decline in oil concentration 

Application of the equation yields a half-life of 12 h for the dispersed oil concentrations [every 12 hours the concentration is divided by 2 (Merlin and Le Floch 2012)]. This reflects a dilution rate for a sustained spill response implemented over several days in a deep, but nearshore environment.  In more recent toxicology studies carried out in several laboratories in North America (Aurand and Coelho 2005), the exposure duration was modeled after a single dispersant application to offshore, open water habitats establishing a half-life of 4 hours. These representations of ‘spiked’ exposures are more environmentally realistic (closer to real field conditions) than standard laboratory ‘constant’ exposures, and result in a reduced level of effects (Gardiner et al. 2013). 

The effect and toxicity of a water soluble fraction (WSF) versus chemically dispersed oil was studied by using realistic exposure concentrations based on the WSF concentrations  monitored during an offshore field experiment (i.e. initial TPAH concentration of less than 7 ppb; NRC 2005).  The Arctic amphipod Gammarus setosus was used as test species in a continuous flow experiment. Body burden measurements showed higher level of PAHs in the gammarids exposed to oil and dispersant for 12 days than in those exposed to oil alone, consistent with the higher concentrations of oil that would be present when dispersant are used.  Several biomarkers were monitored, and gammarids exposed to oil and dispersant also showed moderate signals of exposure after recovery in clean seawater.

In a recent study on adult and juvenile fish and bivalve species conducted at elevated concentrations (up to 70 mg/L), the observed effects were sublethal and temporary.   After 2 weeks, sublethal bioindicators did not show any differences between animals exposed to the chemically dispersed oil and mechanically dispersed oil (Merlin and LeFloch 2012).  This demonstrates that exposure to chemically dispersed oil is not more toxic than the physically dispersed oil.  However, the same research team reported that fish kept in a natural environment after exposure did show residual responses (persistent) in terms of growth (Merlin and LeFloch 2012).  In conjunction with the previous study, experiments conducted with herring embryos in a wave tank showed abnormalities after constant exposure to elevated concentrations (to 10 ppm), but no effect when a more realistic and rapid dilution exposure regime was generated (McIntosh et al. 2010). 

Dispersant toxicity research has been conducted recently on specific Arctic species of concern as part of a laboratory toxicity testing program conducted in Barrow, Alaska.  It was found that Arctic species that were tested have similar or greater tolerance to representative concentrations of dispersed oil compared to the numerous temperate species that have been tested (Word and Gardiner in prep.).  Also, the acute toxicity of exposures to dispersant alone only occurs at concentrations that are greater than concentrations proposed for application of dispersant products in OSR (McFarlin et al. 2011; Gardiner et al. 2013).  For most species that have been tested, dispersed-oil acute toxicity thresholds are on the order of 1 mg/L based on laboratory tests that expose test organisms for periods of 2 to 4 days.  Water column concentrations above toxicity thresholds in an actual spill are limited to the top few meters and exposures at potentially toxic concentrations are limited in duration due to rapid dilution kinetics. 

4.2.4.2 Conclusions on Chemical Dispersion

The available body of laboratory data, experimental field studies and monitoring following actual spills shows that dispersed oil may potentially cause environmental impacts but these will be limited to the organisms in the immediate vicinity of dispersed oil plume and in cases when the rate of dilution of the dispersed oil plume is slow.  This would be the case for sensitive areas with limited water exchange,e.g. close to the shore.  Even in such cases, these impacts would generally be limited to non-mobile organisms.  For example, monitoring following dispersant use at major oil spill incidents over the past 40 years has never reported significant losses of mature fish populations at sea following dispersant applications. 

Laboratory and field research as well as monitoring following actual incidents assessing the impact of use of dispersants in OSR, demonstrate that:

  • The toxicity of oil/dispersant mixtures is related to the oil in the mixture and not the dispersant.
  • The toxicity of the oil is directly related to the amount of oil that organisms are exposed to.  That is, when dispersants are applied to oil the increase in response of pelagic organisms is directly related to the exposure concentration and duration of exposure to the oil.
  • The toxicity of dispersed oil is relatively low and often not observable in real environment as long as there is no restriction to the rapid dilution process of the plume of dispersed oil (e.g. open-ocean).
  • There is no evidence that Arctic species are more or less sensitive than other temperate climate species that have been tested with dispersed oil.
4.2.4.3 Dispersing Oil using Oil Mineral Aggregates (OMA)

The use of fine mineral particle (such as clay minerals) is an alternative response method to dispersant used to break up an oil slick into small droplets and stabilize the oil dispersion in the water column.  When applied to physically dispersed oil, oil droplets aggregate readily with suspended particulate matter (SPM) such as clay minerals and organic matter to form [oil-SPM] aggregates called oil mineral aggregates (OMA; Le Floch et al. 2002).  It is important to distinguish the use of OMA from sinking agents.  Rather than bind to bulk oil as dense sediment and cause the oil droplets to sink, OMA will cause the oil to be suspended in the water as micron-sized droplets associated with a complex of mineral material in much the same result as chemical dispersants generate micron sized droplets (Khelifa 2005, Khelifa et al. 2005).  The simplest form of OMA consists of an oil droplet coated with micrometer-sized solid mineral particles that prevent the droplets from sticking to each other and reforming a slick.  When OMA forms, the dense mineral fines (small but 2.5 to 3.5 times denser than most oils) adhering to the oil droplets will reduce the overall buoyancy of the droplets, retarding their rise to the surface but keeping them somewhat buoyant so they do not sink.  This promotes oil droplet dispersion throughout the water column to low concentrations, and ultimately enhancing their biodegradation by natural bacteria (Lee et al. 2011).

Positive lab and basin tests of the concept led to a field test in 2008 (Lee et al. 2011).  The field test was designed to evaluate the concept of using an icebreaker’s propeller and application of mineral chalk fines with seawater to create OMA. Visual observations confirmed that the oil stayed physically dispersed in the upper water column and did not resurface (Potter et al. 2012).   Attempts were made to combine chemical dispersant use with fine mineral application.  The dispersant was added to promote the dispersion of micron-sized droplets into the water column while the addition of fines attempted to stabilize this dispersion.  The result was not especially convincing, as dispersant presence seemed to inhibit the formation of OMA complex with the oil droplets.

Preventing the re-surfacing of the droplets under the adjacent ice in the Arctic would be a significant environmental benefit since OMA also enhances natural biodegradation of spilled oil.  The application of fine minerals seems well adapted to ice infested conditions as the presence of ice reduces the sea surface agitation; chemically dispersed oil may tend to resurface over prolonged time periods if not stabilized by OMA formation.  It is also beneficial that the types of fine minerals needed for OMA dispersion are those that are commonly stockpiled in oil exploration facilities as drilling mud components; consequently, a source would be readily available in the event of a spill.

4.2.4.4 Environmental Impact of OMA formation

Most of the studies on this topic were devoted to the mechanism and efficiency of the technique to optimize the application conditions; very few considered the environmental impact of the use of OMA.  A lab study dealing with the use of dispersant in estuaries assessed the toxicity of dispersed oil with presence of Montmorillonite clay (Merlin and Le Floch 2012).  It was shown that the presence of this fine grained material reduced the observed impact on biota exposed to the dispersed oil plume to the level of impact commonly seen from oil that is mechanically dispersed (without chemical dispersant addition).  To date, sparse information has been identified on the environmental impacts and relative toxicity of OMA in the Arctic.

Laboratory and field research as well as monitoring following actual incidents assessing the impact of use of OMA in OSR, demonstrate whether:

  • The toxicity of oil/OMA mixtures is altered or if the oil in the mixture is the predominant cause of any toxicity observed with OMA use.
  • The association of oil and OMA may alter the toxicity of the oil by decreasing bioavailability due to the adsorptive process that occurs to the OMA.
4.2.4.5 Conclusions on OMA

The environmental advantages of using OMA to stabilize oil dispersion in the upper water column are similar to those expected from the use of chemical dispersant.   Mineral fines are nontoxic to marine life.  The main impact expected from addition of the mineral fines could be a temporary increase of the sea turbidity which should be similar to the level of turbidity promoted by chemical dispersion.  Other mechanisms of impact would be similar to the environmental/biological impacts discussed with chemically dispersed oil.  The description of the flocculation of fines to the outer surface of small oil droplets leads to the following questions prior to its acceptance as an OSR option for the Arctic.

  • What is the optimum rate of OMA application to oil to maximize its benefits (OMA may need a 1:1 ratio with oil to provide its benefits.
  • Does OMA surface coating of oil droplets reduce potential microbial degradation or use of the oil droplets?
  • Does OMA surface coating of oil droplets reduce the potential toxicity of the droplets or decrease the solubility and exposure of the more soluble/toxic components of the oil?
  • Are OMA coated oil droplets available to suspension feeding organisms in an unweathered, potentially more toxic form?
  • The remaining questions regarding the long term environmental fate of the OMA aggregates are: do they tend to sink progressively with time? What is the impact of settled OMA to the exposed area of bottom resources that could be large but at very diffuse concentrations?  Does the mineral separate from the oil droplet?   What is the impact of OMA or separated mineral exposures to dilute inorganic particulate matter and would it be any different than that endured from settling of ocean particulate matter? 

4.3 Future Research Considerations

The review of the main oil spill response strategies to be used in the Arctic described by the authors in this section led to suggestions of further research which can reduce remaining uncertainties.  The more generic suggestions compiled from this review are summarized below while recommendations that are important for improving Arctic NEBA are listed separately.  Recommendations for further study captured previously within each response option section are summarized below.

  1. Expand effect data for non-water column species.  Data are available on impacts of OSR technologies on species living in the upper pelagic environment.  Additional studies on species living in other ECs (e.g., deep sea, surface layers, ice/water interfaces, hard substrates) should be performed.  Noteworthy limitations of the available data are:
    1. Most of the toxicological information provided for different response strategies is in terms of sensitivity to organisms that are exposed in the water column (generally the upper 10 m) and do not include species that are using deep water, surface layer, ice/water and shoreline interfaces.
    2. Models have been developed to predict the toxicity of a constant concentration of oil and dispersed oil components to these species based on a narcosis model associated with uptake into lipids.  Other modes of action (fouling, epithelial damage, developmental abnormalities, and others) and exposure types (declining dose) need to be considered.
    3. Assimilated information on recent subsea processes is appearing in the peer-reviewed literature; these studies need to be reviewed and important trends summarized.
  2. Expand Assessment to Population-Level Effects, Resilience, and Recovery Potential.   Toxicity evaluations have been generally limited to studies on the acute impact on individuals of a single species and do not evaluate recovery times for the suite of species, at the population-level.  The resiliency of a population to a stressor is built around the population’s ability to recover from the stress, i.e. it is dependent on how widely the species or population is distributed, its reproductive potential, its mobility, and ability to avoid exposure to the stressor.  Knowledge of the length of recovery time for populations living in each environmental compartment is needed prior to evaluating the consequences of each response decision.  Critical summaries of the biological attributes that extend or shorten recovery times need to be developed for populations residing in the different EC that would be affected by response actions.
  3. Examine persistent effects.  Investigations have been conducted on oil and chemical toxicity, generally focusing only on the acute and short term toxicity.  In NEBA decision-making chronic sublethal effects are often treated as acute mortality levels.  An analysis should be performed to better understand the role of chronic sublethal effects in NEBA decision making. This includes the change in bioavailability, toxicity and biodegradation potential of oil and its components as physical weathering and biodegradation occurs after treatment by OSR actions. 
  4. Explore additional modes of toxic action.  There is little information on other types of effects of oil exposure other than toxicity due to water borne exposure and uptake into tissues; additional study of fouling and epithelial tissue disruption by oil or oil residues on organisms that reside or move through the air/water, ice/water or shoreline interfaces or breathe contaminated air above an oil slick is warranted.  Smothering/fouling related toxicity, atmospheric contamination and interference with air breathing mammals and birds has not been well investigated. 
  5. Evaluate potential for impact on the basis of habitat function.  The available information is often concentrated on individual species or groups of species rather than on habitat function.  However, it is the effects on population or habitat which are important in the overall assessments of the consequences of selecting alternative response actions.

4.3.1 Priority Recommendations for Enhanced NEBA Applications in the Arctic

The recommendations presented below indicate where increased knowledge of OSR processes and consequences would result in reducing existing uncertainties in NEBA assessments.  No prioritization has been made to the list; for some of the recommendations, surrogate data may be already available.

  1. Identify VECs and ECs that can be impacted by each OSR.  Develop structured overviews of all VECs that can potentially be impacted by each OSR technology with special focus on VECs at interfaces and on seasonality.
    1. Improve knowledge on the spatial distribution of the living resources and seasonal evolution in the Arctic because large movements of populations occur.
    2. The resources should be ranked in terms of vulnerability to the oil in its different forms:  surfaced oil (fresh, weathered and burnt), dispersed oil, and smoke and soot (in case of burning), oil on the shore.
  2. Collate available data.  Identify data needs for all potentially impacted VECs to score exposure potential, sensitivity, resilience, recovery preferably on a population level basis, using ARCAT approach.
    1. The oiling mechanism on Arctic shoreline and related impact on living resources taking into account the uniqueness of Arctic shoreline which are subject to strong icing periods.
    2. Short and long term effects of dispersed oil on Arctic living resources, especially those which are the most vulnerable and which are identified to congregate locally at certain times of the year.
    3. Data on air emissions should be included as well as residues for ISB (fate, exposure potential, effects, biodegradation).
  3. Highlight uncertainties in data.  Identify data uncertainties and evaluate needs for additional studies such as:
    1. Effects caused by surfaced oil (ingestion, fouling, smothering) to the different living resources of concern (birds, mammals, etc.)
    2. Increased knowledge of the effects produced by-products (smoke, soot and un-burnt residue) of ISB towards the Arctic living resources subject to direct or indirect ISB contact
    3. Effects of herding chemicals and OMA have not been fully evaluated to date.  Investigation of the intrinsic toxicity of chemical herding agents towards Arctic species especially related to the possible exposure to early life stages.

4.4 Further Information

Authors François Merlin and Dr. Stephane Le Floch (CEDRE), Dr. Jack Q Word (ENVIRON), 

Dr. James Clark (HDR/EM&A), Dr. Janne Fritt-Rasmussen (DCE/Department of Bioscience), Dr. Liv-Guri Faksness (SINTEF) 

References

API

2004

In-situ burning: The fate of burned oil.

Am Petroleum Institute; Publication 4735. 40 pp.

Blackall PJ and GA Sergy

1981

The Bios Project - Frontier oil spill countermeasures research

Proceedings, IOSC 1981(1):167-172

Broje V, AA Keller

2007

Effect of operational parameters on the recovery rate of an oleophilic drum skimmer

Jnl of Hazard Mat 148(1-2):136-143.

Buist I et al.

2006

Mid-scale test tank research on using oil herding surfactants to thicken oil slicks in pack ice : an update

AMOP 2006

Buist I, J McCourt, S Potter, S Ross, K Trudel

1999

In situ burning

Pure Appl Chem 71(1):43-65

Buist I, Portter, T Nedwed, J Mullin

2008

Herding agents thicken oil spills in drift ice to facilitate in situ burning: a new trick for an old dog

Proceedings, IOSC 2008.

Buist I, T Nedwed

2011

Using Herders for Rapid In Situ Burning Of Oil Spills on Open Water

Proceeding, 2011 International Oil Spill Conference

Buist I, Trudel K, Morrison J, Aurand D

1995

Laboratory studies of the physical properties of in situ burn residues

Proceedings,18th Arctic Marine Oil Spill Programme (AMOP) technical seminar, Environment Canada

Buist IA, DF Dickins

1987

Experimental spills of crude oil in pack ice

Proceedings, IOSC 1987 Vol 1:373-381

Buist IA, DF Dickins

1983

Fate and behavior of water-in-oil emulsions in ice.

In Canadian Offshore Oilspill Research Association. Report CS 11. Dome Petroleum Ltd, Calgary, Canada

Buist IA, R Belore, A Guarino, D Hackenberg, DF Dickins, Z Wang

2009

Emperical weathering properties of oil in ice and snow

Proceedings, 32nd AMOP Technical Seminar; Environment Canada, Vancouver, BC. Vol 1, p 67-107

Cormack D, Nichols JA

1977

The Concentration of Oil in Sea Water Resulting from Natural and Chemically Induced Dispersion of Oil Slicks

Proceedings, 1977 IOSC ; API, Washington DC

Daling PS, Indrebo G

1996

Recent Improvements in optimizing use of dispersants as a cost-effective oil spill countermeasure technique

International Conference on Health and Safety and Environment, New Orleans

Daykin M, G Sergy, D Aurand, G Shignaka, Z Wang, A Tang

1994

Aquatic toxicity reslting from in situ burning of oil-on-ice

Proceedings, Seventeenth Arctic and Marine Oilspill Program�Technical Seminar, Environment Canada, Ottawa, Ontario, pp. 1165

Dickins D, PJ Brandvik, J Bradford, L-G Faksness, L Liberty, 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

Proceedings, IOSC 2008 Vol 1:681-688

Dickins DF, IA Buist, WM Pistruzak

1981

Dome’s petroleum study of oil and gas under sea ice

International Oil Spill Conference, IOSC. Pp 183-189

EPPR (Emergency Prevention, Preparedness and Response)

1998

Field guide for Oil Spill Response in Arctic Waters, 1998

Environment Canada, Yellowknife, NT Canada, 348 pp (Contributing authors EH Owens, LB Solsberg, MR West and M McGrath

Faksness L-G, Borseth JF, Baussant T, Hansen BH, Altin D, Tandberg AH, Ingvarsdottir A, Aarab N, Nordtug T

2011

The effects of different oil spill cleanup technologies on body burden and biomarkers in Arctic marine organisms - a laboratory study

Proceedings, 34th AMOP; Banff, Canada. October 4-6, 2011.

Faksness L-G, Daae RL, Brandvik PJ, Leirvik F, Borseth JF

2010

Oil distribution and bioavailability. Field experiment - FEX 2009.

SINTEF report A16584. SINTEF; Trondheim Norway. Www.sintef.no/projectweb/JIP-oil-in-ice/publications/

Fingas MF and M Punt

2000

In-situ burning: A cleanup technique for oil spills on water

Environment Canada Special Publication, Ottawa, Ontario, 214 p.

Fingas MF, Halley G, Ackerman F, Vanderkooy N, Nelson R, Bissonnette MC, Laroche N, Lambert P, Jokuty P, Li K, Halley

1994

The Newfoundland Offshore Burn Experiment - NOBE. Experimental design and overview

Proceedings, 17th AMOP Technical Seminar

Fingas MF, Li K, Ackerman F, Campagna PR, Turpin RD, Getty SJ, Soleki M, Trespalacios MJ, Pare J, Bissonnette MC, Ten

1993

Emissions from mesoscale in situ oil fires: The mobile 1991 and 1992 tests

Proceedings,16th Arctic and Marine Oil Spill Program (AMOP) Technical Seminar, Environment Canada

Fingas, MF, G Halley, F Ackerman, R Nelson, MC Bissonnette, N Laroche, Z Wang, P Lambert, K Li, P Jokuty, G Sergy, W

1995

The Newfoundland Offshore Burn Experiment - NOBE (2)

Proceedings of the 1995 International Oil Spill Conference, American Petroleum Institute, Washington, D.C., pp. 123-132, 1995.

Fraser J et al.

1997

A review of the literature on soot production during in situ burning of the oil, AMOP 1997

AMOP 1997

Fraser JP, EJ Tennyson

1993

Oil spills of opportunity-an appropriate place for response research

Proceedings, 16th Arctic and Marine Oil Spill Technical Seminar, pp 255-269, Calgary, Alberta, Canada

Fritt-Rasmussen J, Ascanius BE, Brandvik PJ, Villumsen A, Stenby EH

2013

Composition of in situ burn residu as a function of weathering conditions

Mar Pollut Bull 67(1-2):75-81

Gallaway B

2012

Estimated impact of hypothetical oil spill in the Eastern Alaskan Beaufort Sea on Arctic cod

Presentation; NewFields/UAF workshop; Anchorage, AK. June 2012

Garrett RM, Guenette CC, Haith CE, Prince RC

2000

Pyrogenic polycyclic aromatic hydrocarbons in oil burn residues

Environ Sci Tech 34(10):1934-1937

Garrett RM, Guenette CC, Haith CE, Prince RC

2000b

In-situ burning of crude oil and emulsions IV: Uncontained and contained in broken ice.

SINTEF Applied Chemistry; report no. STF21 A95026

Hayward A, Kucklick JH, Steen AE, Fritz D

1995

Oil spill chemicals in freshwater environments: Technical issues

Proceedings, IOSC 1995

Hemmer MJ, MG Barron, RM Greene

2011

Comparative toxicity of eitht oil dispersants, Louisiana sweet crude oil (LSC), and chemically dispersed LSC to two aquatic test species

Environ Toxicol Chem 30(10):2244-2252

Ivanov B, A Bezgreshnov, N Kubyshkin, A Kursheva

2005

Spreading of oil products in sea ice and their influence on the radiation properties of the snow-ice cover.

Proceedings, 18th Intnl Port and Oceans Engineering under Arctic Conditions. Vol 2:853-862

Jezequel R

2011

Le brulage des nappes d’hydrocarbures . etat de l’art – technique de mise en oeuvre

Cedre – R/11/26/ /3231

Karlsson J, C Petich, H Eicken

2011

Oil entrainment and migration in laboratory grown saltwater ice

Proceedings, 21st International Conference on Port and Ocean Engineering under Arctic Conditions, July 10-14. Montreal, Canada.

Kenneth R et al.

1997

Oil containment firebooms

AMOP 1997

Khelifa A

2005

Validation de la formation d’aggregats dans une eau saumatre et froide

Technical report n° FJMP3-05RTI; Submitted to the Canadian Coast Guard

Khelifa A,Ajijolaiya LO, MacPherson P, Lee K,PS Hill, Gharbi S, Blouin M

2005

Validation of OMA formation in cold brackish sea waters

Proceedings, 28th AMOP Technical Seminar, Environment Canada, Ottawa, Ontario. Pp. 527-538

Le Floch S, Guyomarch J, Merlin F, Stoffyn-Egli P, Dixon J, Lee K

2002

The influence of salinity on oil-mineral aggregate formation

Spill Sci Tech Bull 8(1):65-71

Lee K, Li Z, Robinson B, Kepkay PE, Blouin M, Doyon B

2011

Field trials of in-situ oil spill countermeasures in ice-infested waters

Proceedings, IOSC 2011

Lewis A and P Daling

2001

Oil spill dispersants

Report: Alun Lewis and SINTEF, 27 pp

Li K, Caron T, Landriault JR, Paré J, Fingas M

1992

Measurement of volatiles, semi-volatiles and heavy metals in an oil burn test

Proceedings,15th Arctic and Marine Oilspill Program (AMOP) Technical Seminar, Environment Canada. pp. 561-573

Lin Q, Mendelssohn IA, Carney K, Miles SM, Bryner NP, Walton WD

2005

In-situ burning of oil in coastal marshes. 2. Oil spill cleanup efficiency as a function of oil type, marsh type, and water depth

Environ Sci Technol 39(6):1855-1860

Mabile N

2012

The coming of age of controlled in-situ burning: Transition from alternative technology to a conventional offshore spill response option

Interspill Conference; London, UK; March 13-15, 2012

McAuliffe CD, JC Johnson, SH Greene, GP Canevari, and TD Seral

1980

Dispersion and Weathering of Chemically Treated Crude Oils on the Ocean

Environmental Science Technology. 14: 1509-1518.

McFarlin K, Perkins RA, Gardiner W, Word JD, Word JQ. 2001.

2011

Toxicity of physically and chemically dispersed oil to selected Arctic species

Proceedings, 2011 IOSC; Portland

McIntosh S, T King, D Wu, PV Hodson

2010

Toxicity of dispersed weathered crude oil to early life stages of Atlantic herring (Clupea harengus)

Environ Toxicol Chem 29(5):1160-1167

Merlin F, Le Floch S

2012

Evaluation de l’impact des produits dispersants utilisés pour lutter contre les pollutions pétrolières maritimes en zones côtières ou estuariennes

Cedre report R.12.12.b.C./3258

Miller MC, Stout JR

1986

Effects of a controlled under-ice oil spill on invertebrates of an arctic and a subarctic stream

Environ Pollut Ser A, Ecol & Biol 42(2):99-132

MSRC

1993

Chemical oil spill treating agents

MSRC - Technical Report Series – 93-015

National Research Council (NRC)

2005

Oil Spill Dispersants: Efficacy and Effects 2

The National Academy Press, Washington, DC.

NERI

2002

Potential environmental impacts of oil spills in Greenland

NERI Technical Report N) 415

Ohtsuka N, K Ohshima, K Maida, N Usami, H Saieki

1999

Weathering of oil spills in icy seawater and characteristics of crude oil which trapped into ice floe

Proceedings, 15th Intnl Port and Oceans Engineering under Arctic Conditions. Vol 2: 828-837

Ohtsuka N, K Origwara, S Takahashi, K Maida, Y Watanabe, H Saeki

2001

Experimental study on spreaading of oil under ice floes.

Proceedings, 16th Intnl Port and Oceans Engineering under Arctic Conditions. Vol 3: 1333-1342

Ross JL, RJ Ferek, PV Hobbs

1996

Particle and Gas Emission from an In Situ Burn of Crude Oil on the Ocean

Jnl Air Waste Manag Assoc 46 251-259

SEEEC

1998

The Environmental Impact of the Sea Empress Oil Spill

The Stationery Office, London

Sergy GA, Blackall PJ

1987

Design and Conclusions of the Baffin Island Oil Spill Project.

Arctic 40(supp1):1-9

Serova I

1992

Behaviour of oil in ice and water at low temperatures. Combatting marine oil spills

In: Ice and Cold Climates, HELLCOM Seminar, Helsinki, Finland

SL Ross

2010

A two year research program on employing chemical herders to improve marine oil spill response operations

SL Ross Environmental Research Ltd August 2010

SL Ross

2012

Mid-scale test tank research on using oil herding surfactants to thicken oil slicks in broken ice

In. Potter S, uist I, Trudel K et al/ - Spill Response in !rctic Offshore – prepared for the !merican Petroleum Institute and the Joint

SL Ross

2011

Spill response gap study for the Canadian Beaufort Sea and the Canadian Davis Strait

SL Ross Environ Res Ltd; Ottawa, Ontario

SL Ross

2012b

Research on using oil spill herding agents for rapid response In situ burning of oil slick on open water

Report prepared for US Dept Interior; OSSR program. SL Ross Environmental Research Ltd August 2012; Ottawa, ON. 95 p.

Tennyson EJ

1994

In situ burning overview

National Institute of Standards and Technology and Minerals Management Service, NIST

Walkert AHA, Kucllick J, Michel J

1999

Effectiveness and environmental considerations for non-dispersant chemical countermeasures

Pure Applied Chemistry 71 (1) 1999

Walton WD, Jason NH

1998

In situ burning of oil spills workshop proceedings

NIST special publication

Wang Z, Fingas M, Shu YY, Sigouin L, Landriault M, Lambert P

1999

Quantitative Characterization of PAHs in Burn Residue and Soot Samples and Differentiation of Pyrogenic PAHs from Petrogenic PAHs-The 1994 Mobile Burn Study

Environ Sci Technol 33:3100-3109

Aurand DV and GM 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

Faksness LG, Hansen BH, Altin D, Brandvik PJ

2012

Chemical composition and acute toxicity in the water after in situ burning – ! laboratory experiment

Mar Poll ull 64.49–55

McFarlin KM, RC Prince, R Perkins, MB Leigh

2014

Biodegreadation of dispersed oil in Arctic seawater at -1C

PLOS ONE; www. plosone.org

Word JQ, WW Gardiner, M Barron, A Bejarano

review

Relative Sensitivity of Arctic and Non-Arctic Marine Species to Physically and Chemically Dispersed Crude Oil

In preparation

AMAP

2007

Arctic Oil and Gas 2007

Paper presented at the Arctic Monitoring and Assessment; Programme, Oslo.

Blenkinsopp S, Sergy G, et al.

1997

Evaluation of the toxicity of the weathered crude oil used at the Newfoundland Offshore Burn Experiment (NOBE) and the resultant burn residue

AMOP 1997

Cross WE

1987

Effects of Oil and Chemically TreatedOil on Primary Productivity of High Arctic Ice Algae Studied in situ

Arctic 40(suppl 1):266-276

IMO

2012

Guidelines for the use of dispersants for combatting oil pollution at sea; 2012 draft version

International Maritime Organization; oil pollution response committee

5.0 BIODEGRADATION

Executive Summary

Microbial communities rapidly adapt to increasing abundance of hydrocarbon utilizers in all aquatic environments that have been studied.

Carbon utilization by these microbial populations mineralize complex mixtures of carbon compounds to CO2, degrade complex compounds to smaller molecules, and directly place these carbon atoms into the tissues of the microbial populations.

Microbial communities in the Arctic are adapted to life in this extreme environment and rapidly respond to carbon rich but nitrogen poor petroleum resources. Many of the organisms are unique to this environment while others are similar to those species that also respond to petroleum in other parts of the world.

The speed of microbial utilization of oil is primarily related to the amount of surface area exposed to aerobic processes. Thick layers of oil, physically weathered oil, or oil isolated to less aerobic environments undergo biodegradation more slowly than dispersed oil.

The cold waters and harsh environmental conditions in the Arctic have led some investigators to conclude that spilled oil will be more recalcitrant and remain in the environment for longer periods of time than oil spilled in less harsh environments. Other investigators have concluded that the microorganisms in the Arctic are adapted to those conditions and rapidly respond to spilled oil. The objective of this document is to examine experimental evidence from peer-reviewed literature that addresses these conclusions and identify areas of additional research. 

This chapter focuses on biodegradation of oil spilled into the marine Arctic environment. The state-of-knowledge will be summarized, and main uncertainties that need further addressing will be highlighted. In this summary we also describe some of the specific conditions that organisms involved in biodegradation processes encounter in the Arctic environment but not at lower latitudes. These differentiating conditions, which may affect biodegradation rates and capabilities in the Arctic, include generally low near-surface temperatures, ice coverage and (in open water) extreme fluctuations in the photoperiod during the Arctic summer and winter. 

The risk of oil pollution in the Arctic is relevant to both marine and terrestrial environments. The Arctic marine environment may be influenced by recently discovered oil and gas deposits, together with the potential for increased traffic attributable to shipping lanes made possible by historically low summer ice cover. The terrestrial environments may be exposed to pollution risk, for instance, from oil transport pipeline systems and other land-based activities. In this review we will focus on the marine environments, although references will also be made to terrestrial studies, when relevant. 

Oil compounds can be degraded by a variety of prokaryotic and eukaryotic organisms. After uptake in eukaryotic macro-organisms, provided they are at sub-lethal or sub-inhibitory concentrations, the compounds may be transformed by the metabolic apparatus, bio-accumulated, or removed through fecal excretion. Despite the range of organism types that can act on oil, however, the microbial communities are regarded as the main source for biological degradation of marine contamination. This perspective is based on the large concentrations of microorganisms, large surface-to-volume ratios and their rapid responses to the situation. This section will therefore focus on the microbial populations and their responses to possible oil discharges in the Arctic marine environment. 

Based on sequencing studies of the 16S rRNA gene, in the mid-1970s Carl Woese and colleagues separated living organisms into the domains Eukarya, Archaea and Bacteria (Woese et al. 1990). Microorganisms are present in all three domains. The bacteria and archaea are all microorganisms, though some (e.g. cyanobacteria) sometimes aggregate into visible mats or filaments. Among the eukaryotes are unicellular microorganisms often referred to as protists, which include the photosynthetic phytoplankton as well as some heterotrophic organisms (e.g. amoebae). Within this diverse collective of microorganisms, biodegradation of oil compounds is primarily associated with bacterial processes. In the marine environment the archaeal responses to oil pollution have mainly been considered low when compared to bacterial responses (e.g. Röling et al. 2004). However, archaea are important members of communities in petroleum reservoirs and are involved in the mineralization of petroleum hydrocarbons, for instance by methanogenesis under anoxic conditions. Recent studies have also implied the contribution of halophilic archaea to the biodegradation of hydrocarbons in the marine environment (Al-Mailem et al. 2010; Liu et al. 2009). Specific to the Arctic marine environment, essentially nothing is known of the archaeal contributions to oil degradation. 

Among eukaryotes, degradation of oil is mainly associated with fungi (Prince 2005). Only a few studies have been conducted to characterize fungal communities in Arctic marine environments (e.g. Gunde-Cimerman et al. 2003; Butinar et al. 2011). Among the microorganisms, our understanding of the protist contribution to oil biodegradation is also limited. Heterotrophic protists may affect petroleum biodegradation indirectly through their role as bacterial predators. Efficient grazing by protists can maintain the bacterial community in the log-growth phase, resulting in increased metabolic rates and thus increased biodegradation rates (Ballestero and Magdol 2011). 

In summary, most evidence for microbial degradation in the Arctic is connected to bacteria and bacterial communities. Our focus in this section will be on bacterial communities and the bacterial processes involved in oil biodegradation in the Arctic environment. However, the need for greater knowledge related to the contribution of archaeal and protistan organisms to oil biodegradation is an incentive to develop a better understanding of their contribution to oil biodegradation and microbial ecology in the Arctic. 

5.1 Introduction

The cold waters and harsh environmental conditions in the Arctic have led some investigators to conclude that spilled oil will be more recalcitrant and remain in the environment for longer periods of time than oil spilled in less harsh environments. Other investigators have concluded that the microorganisms in the Arctic are adapted to those conditions and rapidly respond to spilled oil. The objective of this document is to examine experimental evidence from peer-reviewed literature that addresses these conclusions and identify areas of additional research. 

This chapter focuses on biodegradation of oil spilled into the marine Arctic environment. The state-of-knowledge will be summarized, and main uncertainties that need further addressing will be highlighted. In this summary we also describe some of the specific conditions that organisms involved in biodegradation processes encounter in the Arctic environment but not at lower latitudes. These differentiating conditions, which may affect biodegradation rates and capabilities in the Arctic, include generally low near-surface temperatures, ice coverage and (in open water) extreme fluctuations in the photoperiod during the Arctic summer and winter. 

The risk of oil pollution in the Arctic is relevant to both marine and terrestrial environments. The Arctic marine environment may be influenced by recently discovered oil and gas deposits, together with the potential for increased traffic attributable to shipping lanes made possible by historically low summer ice cover. The terrestrial environments may be exposed to pollution risk, for instance, from oil transport pipeline systems and other land-based activities. In this review we will focus on the marine environments, although references will also be made to terrestrial studies, when relevant. 

Oil compounds can be degraded by a variety of prokaryotic and eukaryotic organisms. After uptake in eukaryotic macro-organisms, provided they are at sub-lethal or sub-inhibitory concentrations, the compounds may be transformed by the metabolic apparatus, bio-accumulated, or removed through fecal excretion. Despite the range of organism types that can act on oil, however, the microbial communities are regarded as the main source for biological degradation of marine contamination. This perspective is based on the large concentrations of microorganisms, large surface-to-volume ratios and their rapid responses to the situation. This section will therefore focus on the microbial populations and their responses to possible oil discharges in the Arctic marine environment. 

Based on sequencing studies of the 16S rRNA gene, in the mid-1970s Carl Woese and colleagues separated living organisms into the domains Eukarya, Archaea and Bacteria (Woese et al. 1990). Microorganisms are present in all three domains. The bacteria and archaea are all microorganisms, though some (e.g. cyanobacteria) sometimes aggregate into visible mats or filaments. Among the eukaryotes are unicellular microorganisms often referred to as protists, which include the photosynthetic phytoplankton as well as some heterotrophic organisms (e.g. amoebae). Within this diverse collective of microorganisms, biodegradation of oil compounds is primarily associated with bacterial processes. In the marine environment the archaeal responses to oil pollution have mainly been considered low when compared to bacterial responses (e.g. Röling et al. 2004). However, archaea are important members of communities in petroleum reservoirs and are involved in the mineralization of petroleum hydrocarbons, for instance by methanogenesis under anoxic conditions. Recent studies have also implied the contribution of halophilic archaea to the biodegradation of hydrocarbons in the marine environment (Al-Mailem et al. 2010; Liu et al. 2009). Specific to the Arctic marine environment, essentially nothing is known of the archaeal contributions to oil degradation. 

Among eukaryotes, degradation of oil is mainly associated with fungi (Prince 2005). Only a few studies have been conducted to characterize fungal communities in Arctic marine environments (e.g. Gunde-Cimerman et al. 2003; Butinar et al. 2011). Among the microorganisms, our understanding of the protist contribution to oil biodegradation is also limited. Heterotrophic protists may affect petroleum biodegradation indirectly through their role as bacterial predators. Efficient grazing by protists can maintain the bacterial community in the log-growth phase, resulting in increased metabolic rates and thus increased biodegradation rates (Ballestero and Magdol 2011). 

In summary, most evidence for microbial degradation in the Arctic is connected to bacteria and bacterial communities. Our focus in this section will be on bacterial communities and the bacterial processes involved in oil biodegradation in the Arctic environment. However, the need for greater knowledge related to the contribution of archaeal and protistan organisms to oil biodegradation is an incentive to develop a better understanding of their contribution to oil biodegradation and microbial ecology in the Arctic. 

5.1.1 The Microbiology of the Arctic Oceans

5.1.1.1 Transport routes

The Arctic Ocean represents a complex system of currents with influx and efflux of water (see Figure 5-1).  Cold and relatively less saline water enters the Arctic Ocean through the narrow Bering Strait between Alaska and Siberia, while warmer, more saline surface waters from the Atlantic penetrate the Arctic Ocean and are cooled as they move through the Greenland Sea and the Norwegian Sea. Water reaching the Arctic Ocean basin is swept into a huge circular current — driven by strong winds — the Beaufort Gyre. Siberian and Canadian rivers drain into the circular marine current to create a captured internal reservoir of relatively fresh water. Periodically, the circular current weakens, allowing large volumes of fresh water to leak out and cross the Arctic in the Transpolar Current. The water exits the Arctic Ocean via several “gateways.” It can flow through the Fram Strait, between northeast Greenland and Svalbard Island, and then branch around either side of Iceland. It can flow around the west side of Greenland through Baffin Bay and out Davis Strait. It can also flow through a maze of Canadian islands and out Hudson Strait. These mixes of salty and fresh water also generate Arctic haloclines, a vertical effect in which the cold fresh water lies atop warmer saltier water.

Figure 5-1. Prevailing currents in the Arctic Ocean (Source: Woods Hole Oceanographic Institution; http://www.divediscover.whoi.edu
Figure 5-1. Prevailing currents in the Arctic Ocean (Source: Woods Hole Oceanographic Institution; http://www.divediscover.whoi.edu

The transport routes carry microorganisms into the Arctic from a complex mixture of seawater and river sources, with the North-Atlantic current (number 6 in Figure 5-1) as the single most important influx transport route. This results in cosmopolitanism of microorganisms in the Arctic Ocean and the presence of sometimes unexpected microorganisms in this cold marine environment (Hubert et al. 2009).

Significant to the discussion of currents, water masses, and effects on microbial population structures, a considerable part of the Arctic consists of one- or multi-year ice coverage. Since marine ice is frozen seawater the microbes frozen into the ice are transported with ice movement. This movement essentially follows two routes; the Beaufort Gyre circulating clockwise around the North Pole and the Transpolar Drift Stream, where ice moves from the Siberian coast of Russia across the Arctic basin, exiting into the North Atlantic off the east coast of Greenland (number 5 in Figure 1).  During ice growth, brine pockets are generated in the ice, since ice crystal generation is almost devoid of impurities (Petrich and Eicken 2010).  Brine accumulations are first generated as pockets, and later as chimney-like tubes termed brine channels. Microbes require a fluid environment for active metabolism, and these brine channels represent a fluid microenvironment within the ice mass. These channels therefore become important ecological niches for microbes which are able to survive in the high salinity and subzero temperatures in these systems. Eventually, contact with seawater will result in nutrient transport and reduced salinity in the channels. If channels are wide enough (approximately 2 mm in diameter), they are able to support counter-flow, i.e. simultaneous upward and downward movement inside the same channel (Lake and Lewis 1970).  Despite the potential for flow and mixing with bulk marine water, brine salinity tends to increase with distance to the seawater.   These salinity gradients in brine channels may result in gradients of microbial population density.  In winter, the apparent effect is higher bacterial and algal concentrations in the lower than the upper parts of the ice (Gradinger and Zhang 1997). However, in late spring and summer with rising air temperatures, the opposite situation is observed. The ice is warmed up from the top, becoming more porous, and melt water penetrates the porous ice and generates a linear bulk salinity profile that is commonly observed in the upper 20 to 50 cm of the ice during melt. The organisms in the brine channels are therefore exposed to fluctuating and physiologically challenging salinity conditions during the periods of ice freezing and melting (Faksness et al. 2011). Since much of the brines are drained off the ice during the first year after freezing, multi-year ice tends to have lower salinity and density than first-year ice (Tang et al. 2007), resulting in fewer brine pockets and also fewer possibilities for active microbial metabolism in the ice.

5.1.1.2 Microbial populations in the Arctic Ocean

The population structures of bacteria in Arctic seawater are comparable to those in seawater from temperate regions, with the predominance of Alphaproteobacteria, Flavobacteria/Bacteroidetes,Gammaproteobacteria and Verrucomicrobia constituting more than 90% of the communities (Comeau et al. 2011; Teske et al. 2011; Ghiglione et al. 2012). It is also relevant to compare the microbial communities in the Arctic and Southern Oceans, since the two geographical areas have similar climatic conditions. Although some comparative studies of the microbial communities in Arctic and Antarctic marine environments indicated a high degree of resemblance (e.g. Bano et al. 2004; Brinkmeyer et al. 2003), a recent pole-to-pole study of surface and deep marine bacterial communities revealed significant differences, with 78% of operational taxonomic units (OTUs) unique to the Southern Ocean and 70% unique to the Arctic Ocean. Despite these dissimilarities, polar ocean bacterial communities were more similar to each other than to lower latitude pelagic communities (Ghiglione et al. 2012).

At a more localized scale, sea ice microbial communities may differ from those in the surrounding seawater. A study of first-year Arctic ice showed differences between ice and source seawater populations that were characterized by an increased abundance of Gammaproteobacteria and a lower abundance of Bacteroidetes in sea ice relative to seawater.  Winter sea ice communities were more similar to autumn water communities; species-specific die off processes were not observed in sea ice.  The implication was that winter sea ice communities are stable, whereas seawater communities undergo seasonal succession. The Alphaproteobacteria dominated in both environments (Collins et al. 2010). A study of multi-year ice revealed overall dominance of Gammaproteobacteria, as well as an increase ofBacteroidetes in the sea ice when compared to source seawater; the abundance of Alphaproteobacteria was greatly reduced in sea ice relative to the abundance in surrounding seawater (Bowman et al. 2012). A number of studies have also investigated the microbial communities in Arctic sediments, which may be regarded as pristine environments compared to most other sediments investigated. Studies from an Arctic fjord (Svalbard) showed that the Alphaproteobacteria predominated in the overlying seawater, while Gammaproteobacteria were more abundant in surface sediments. Deeper anoxic sediment layers were dominated by Deltaproteobacteria, which include common genera of strictly anaerobic bacteria associated with marine sediments (Ravenschlag et al. 2001; Teske et al. 2011).

5.1.2 Microbial Adaptation to Arctic Conditions

The primary climatic condition associated with the Arctic is low temperature, and therefore most reviews and fundamental studies of Arctic microbiology focus on low temperature and psychrophilic (cold-loving) microbes. However, other conditions are also specific for this environment, including extreme light/dark conditions in the Arctic summer and winter, respectively. The Arctic area also includes coverage, with 9-12 million km2 of pack ice with average ice thicknesses of 3-4 m, with ridges up to 20 m thick. As described above the microbes in this ice are not only challenged by the subzero temperature, but also by high salt concentrations in the brine channels.

5.1.2.1 Low temperature and microbial adaptions

The temperature of the Arctic Ocean normally does not exceed 4-5°C, with cold (close to 0°C) low-salinity surface water and warmer high-salinity water (up to 5°C) below 50 m depth. The marine organisms in the Arctic are therefore consistently exposed to low temperatures. Numerous studies of microbial adaptions to low temperature have been published (e.g.  Brakstad et al. 2008; Helmke and Weyland, 2004; Knoblauch et al. 1999; Robador et al. 2009). As the temperature decreases with concomitant decrease in the thermal energy of a system (enthalpy), increased stability and rigidity of biomolecules occur. Proteins become less flexible, while the secondary structures of DNA and RNA become more stable, and the processes of DNA replication, transcription of RNA and translation of proteins are inhibited (Bakermans 2012). However, psychrophilic microorganisms have adapted to these low-temperature conditions in a number of ways:

  • Enzymes in bacteria isolated from sea ice may be activated by the cold, with detectable catalytic activity well below the freezing point of seawater (Groudieva et al. 2004). The enzymes of psychrophilic microbes are characterized by increased flexibility of active sites resulting from structural modifications to address freezing temperatures.  These modifications may include less enzyme core hydrophobicity, increased surface polarity and fewer interactions between amino acids (hydrogen bonds, salt and disulphide bridges). These characteristics facilitate retention of enzymatic activities at low temperature (Bakersman 2012).
  • Nucleic acids show adaption to low temperature at the sequence level in psychrophiles, as exemplified by increased uracil content in ribosomal RNA and dihydrouridine content in transfer RNA as a result of decreased temperature (Khachane et al. 2005; Dalluge et al. 1997).
  • Microbial membrane structures may crystallize at low temperatures, impairing function. However, in psychrophiles, membrane fluidity is retained by increasing the proportion of unsaturated fatty acids and by desaturation of membrane lipids, resulting in unsaturated fatty acyl chains and less tightly packed lipid chains (Russel 2008).
  • Microorganisms can be challenged in ice by reduced water availability due to the formation of ice crystals. However, psychrophiles can combat the effects of reduced water activity by the synthesis of compatible solutes such as sugars, amino acids, alcohols or cryoprotective proteins. Ice crystal formation can also be prevented by production of anti-nucleating materials (Bakersman 2012).
  • Psychrophilic bacteria in brine channels are challenged by both low temperature and high salinity.  These may produce extracellular polymeric substances (exopolysaccharides), which have a cryoprotective role in the sea ice brine channels, as well as binding essential cationic trace metals to enhance halotolerance (Nichols et al. 2005).
  • Microbes exposed to large drops in temperature may produce cold shock and cold acclimation proteins (Berger et al. 1996). Cold shock proteins are supposed to be involved in protein translation regulation, while cold acclimation proteins show high catalytic activity at low temperatures and rapid inactivation at moderate temperatures (Fukunaga et al. 1999).
5.1.2.2 Light and microbial phototrophs

In Arctic regions, with a long polar summer and a correspondingly long polar winter and associated widespread ice coverage, phototrophic organisms must tackle a number of challenges.  For example, phytoplankton communities responded rapidly and increased in numbers by an order of magnitude in the Beaufort Sea with the increased solar irradiation in May, despite the presence of sea ice. However, eukaryotic phototrophs somehow also persisted throughout the winter darkness (Terrado et al. 2008). Among phototrophic bacteria isolated from the Arctic Ocean were photoheterotrophic microbes, which are capable of both utilizing dissolved organic materials and harvesting light energy.  Photoheterotrophic taxa included coccoid cyanobacteria (Synechococcus and Prochlorococcus), aerobic anoxygenic phototrophic (AAP) bacteria, and proteorhodopsin (PR)-containing bacteria. In a study of photoheterotrophic microbes in the Arctic Ocean in the summer and winter it was observed that some of these bacteria (AAP and PR containing bacteria) decreased from summer to winter, in parallel with a threefold decrease in the total prokaryotic community, while the abundance of Synechococcus did not decrease in winter (Cottrell and Kirchman 2009). These data demonstrated the ability of phototrophs to switch from photosynthesis to heterotrophic activities during the winter.  Characterization of green light-absorbing PR from bacterial isolates from the Arctic Ocean showed a slower photocycle, relatively larger change in activation energy of the transition between the oxygen intermediate and the ground state compared to PR from bacterial phototrophs isolated from temperate seawater (Jung et al. 2008).

Photosynthesizing microbes have also been observed within sea ice. Marine algae seem to be more sensitive to brine salinity than bacteria and therefore are detected mainly in the lower parts of sea ice (lower salinity in winter associated with contact with seawater; Gradinger and Zhang 1997). Recent studies have also identified marine autotrophic bacteria in marine ice indicating the potential ability for ongoing use of dissolved organic materials (Díez et al. 2012).

5.1.2.3 Marine ice and microbial survival and metabolism

Conditions in marine ice vary considerably from those in seawater. The temperature is lower in the ice relative to fluid water, and may drop as low as the air temperature during the polar winter. Transmission of light through the ice is severely reduced compared to seawater, and light may be entirely absent when the ice is covered by a snow cap. In first-year ice, brine inclusion networks are generated by salting-out processes, as described in previous sections. The brine channels of the ice provide liquid saline niches at subzero temperatures which enable microbial motility and metabolic activity (Junge et al. 2002; 2003; 2004; 2006). The brine channels also act as a matrix for transport of soluble organic matter and other nutrients.  In the brine channels the ice temperatures can be as low as -20°C and salinity conditions can be as high as 200 Practical Salinity Units (Salinity in normal seawater is 35 PSU). Though challenging to microbial life, laboratory experiments have shown that bacteria (Colwellia psychroerythraea) were able to incorporate amino acids into proteins at temperatures as low as -20°C in brines (Junge et al. 2006). This metabolic activity was associated with particles or with the surfaces of the brine channels (Junge et al. 2003; 2004). In addition, the ability to grow at subzero temperatures has been demonstrated in the bacterium Psychromonas ingrahamii, which was cultured at temperatures as low as ‑12°C (Breezee et al. 2004). The brine channels may become depleted in oxygen, for instance if microbial processes occur without oxygen renewal (seawater or photosynthesis), and anaerobic denitrification processes have been measured in marine ice (Rysgaard and Glud 2004). These studies therefore demonstrate that bacterial activity can take place in marine ice, both with respect to motility within the ice and active metabolism. Recently, Mykytzuk et al. (2013) presented a detailed characterization of the physiological adaptations that enable a halotolerant psychrophilic permafrost bacterium to grow and metabolize at temperatures as low as -15°C.  The implication is that halotolerant psychrophilic species may also exist in the ice.

5.2 Knowledge Status

Since Arctic environments are increasingly exposed to petroleum exploration, production and transport, the studies of microbes and processes involved in cold-environment biodegradation is essential both for the industries operating in these areas and the governmental bodies responsible for environmental stewardship.  As described, microbes can function within most Arctic environments, including ice at sub-zero temperatures.

Numerous studies have shown that populations of known oil-degrading bacteria are present in Arctic environments (Zinger et al. 2011, Ghiglione et al. 2012; Sul et al. 2013).  However, few studies have directly assessed microbial community composition with the ability to clearly distinguish all of those taxa that can be defined as oil-degrading bacteria (e.g. heterotrophic and other non-designated microbes that respond positively to oil). 

The physico-chemical characteristics and weathering conditions of different oils at low temperatures and which are treated chemically may vary, having impacts on biodegradation efficiencies and should be an issue for further research, for instance, in the relationship between biodegradation and oil appearance (e.g. viscosity, dispersibility, resurfacing) after a spill.  A variety of processes in the Arctic are season-variable (e.g. photooxidation, different ice-conditions).

Marine ice poses a particular challenge to the indigenous microbes, which require the ability to survive and be metabolically active at sub-zero temperatures and at high salinity.  Microbial metabolism has been demonstrated in marine ice, and populations have been stimulated by oil in the ice.  However, the extent of hydrocarbon biodegradation in Arctic ice is likely to be low but requires more attention.

Few studies have attempted to use modeling approaches to predict oil biodegradation in cold marine environments.  Modeling tools are used today by oil companies and authorities to predict the fate of oil spills, but these models are often less well calibrated for the Arctic environment.    In that respect temperature-related biodegradation data, which often are based on Q10-approaches may show erroneous results at very low temperatures, probably due to physical changes in the oil.  Therefore, more detailed temperature-related biodegradation studies will be required to improve fate models, which often rely on incomplete data sets for cold climate spills. 

5.2.1 Biodegradation of Oil in Cold Marine Environments

Typical oil discharge scenarios include process losses or accidental releases from exploration, production, processing or transport related activities. However, the risk of accidental oil discharges will always be present in an environment with oil-related activities. In addition, the diminishing Arctic ice zones will likely increase ship activity through the Northwest and Northeast passages, activities which may contribute to an increased risk of oil discharges in Arctic marine environments. Biodegradation of oil compounds is regarded as the most complete process, as it is able to completely remove oil compounds, through the mineralization to carbon dioxide and water. In particular, two large oil spills in marine water have drawn public attention to oil biodegradation issues in general, the Exxon Valdez grounding in Prince William Sound in 1989 and the recent Deepwater Horizon blowout in 2010, both recently reviewed by Atlas and Hazen (2011). Media attention on the deep water plume from the Deepwater Horizonblowout has resulted in increased focus on the potential effects of oil spills in cold water environments. 

5.2.1.1 Types of Crude Oils

Physico-chemical properties of the spilled oil will affect biodegradation. Crude oils may be separated into 4 main types; paraffinic, asphalthenic, naphthenic and wax-rich (Moldestad et al. 2003). For instance, paraffinic oils will have a distinct chromatographic n-alkane pattern and a high content of light compounds like BTEX and naphthalenes. These oils therefore contain a high degree of easily biodegradable compounds. However, naphthenic oils are often biodegraded in-reservoir (Head et al. 2003), and in these oils most of the easily biodegradable compounds may have been consumed before oil production.  Asphalthenic oils are rich in poorly biodegradable asphaltenes and resins, while wax-rich oils will often have a high pour point and may be solidified in cold water, a physical factor that will affect their availability for biodegradation.  Marine fuel oils may also be spilled.

5.2.1.2 Surface oil spills

When oil is discharged to the marine environment a number of weathering processes occur:

5.2.1.2.1 Evaporation

Volatile compounds with low boiling points (e.g. saturates up to nC11, mono- and some diaromatic hydrocarbons) are rapidly evaporated after surface spills. These are compounds that are normally rapidly biodegradable in the water column, but evaporation is normally more rapid than biodegradation after a surface spill. Evaporation is slower in cold than in temperate seawater (Brandvik et al. 2005), and this may result in temporarily higher concentrations of volatile toxic compounds (e.g. BTEX) in the seawater. At high concentrations these compounds may prolong microbial lag-phases and delay the onset of biodegradation (Atlas and Bartha 1972; Hokstad et al. 1999), although it is not known if this will have an impact on biodegradation under field conditions. In subsurface releases these volatiles are rapidly dissolved from dispersed oil and we suspect may be biodegraded rather than evaporated. Evaporation also results in increased viscosity of the residual oil (Faksness 2008), which will negatively affect the ability of oil to disperse, thereby slowing biodegradation.

5.2.1.2.2 Water solubility

Components dissolved from the oil phase are available for biodegrading microbes in the water column. In cold seawater the dissolution of oil compounds is decreased compared to temperate water (Faksness, 2008). The typical oil compounds in a water-soluble fraction (WSF) from fresh oils include phenols, naphthalenes and 2-3 ring PAHs. In addition, the WSF contains considerable amounts of highly polar compounds with nitrogen, sulphur, and oxygen atoms in their structures (so-called NSO compounds), often present as a chromatographic "hump", termed the "unresolved complex mixture" (UCM). In a study of WSF from an in-reservoir biodegraded oil (Troll) approximately 70 % of the WSF was separated by preparative high-pressure liquid chromatography, into a polar fraction (Melbye et al. 2009).  The non-polar compounds of the WSF are often considered to be rapidly biodegraded in the marine environment (Brakstad and Faksness 2000), and biodegradation of these compounds may result in a significant increase in the UCM  concentration relative to other crude oil components (e.g. Meredith et al. 2000), highlighting their persistence (Han et al. 2008).

5.2.1.2.3 Photooxidation

Photooxidation is an important process in degrading and transforming crude oil compounds after release to the environment. The polar region exhibits vast seasonal differences in light conditions, and as a result photooxidation varies significantly between the polar summer and winter. UV-irradiation of crude oils has shown that aliphatic compounds are mainly resistant to photodegradation, while aromatic compounds appear particularly sensitive to this process (Maki et al. 2001). In contrast to biodegradation, increased size and alkyl substitution result in increased sensitivity of aromatic hydrocarbons to photochemical oxidation. As photooxidation leads to the inclusion of oxygen atoms in the structures of these compounds, photooxidized products appear mainly in the polar resin fraction of the oil (Maki et al. 2001; Garrett et al. 1998; Prince et al. 2003). Additionally, the average molecular weight of oil compounds is reduced and the oxygen content increased.  Studies have also shown that the photooxidized compounds subsequently exhibit increased susceptibility to biodegradation (Dutta and Harayama 2000; Maki et al. 2001; Ni'matuzahroh et al. 1999). Consistent with the physico-chemical properties of the photooxidized compounds, both the dissolved organic carbon concentration and acute toxicity of the water-soluble fraction of oil increased during the irradiation period (Maki et al. 2001). Thus, photooxidation results in a greater proportion of oxidized compounds that exhibit increased water-solubility and subsequently more significant impacts on toxicity and biodegradation. However, investigation of the relationship between photooxidation, biodegradation and toxicity would be of interest as part of the fate-determination of different oil compounds during the Arctic summer.

5.2.1.2.4 Sedimentation

In shallow seawater and at higher levels of suspended sediments (e.g. after a storm), sediment particles may adhere to the oil and sink to the subtidal seabed sediments. Oil spills may also drift to shore and be mixed into the intertidal sediments. Oils mixed into the sediments will be subject to microbial processes in the sediment (Refer to Section 3). If seawater replenishment is poor, aerobic processes may consume most of the oxygen, resulting in anoxic conditions. Anaerobic biodegradation of hydrocarbons, by several alternative mechanisms, will occur in the absence of oxygen, as reviewed by Heider (2007).  Genes associated with anaerobic hydrocarbon degradation (e.g. benzyl- and alkylsuccinate synthase genes) have been detected in hydrocarbon-contaminated sediments (Callaghan et al. 2010).  

5.2.1.2.5 Water-in-oil emulsification

Water uptake into the spilled oil may cause the formation of viscous and often stable water-in-oil emulsions. Emulsions have been shown to be poorly biodegradable (Brakstad et al. 2011; Cook et al., 2011). Water taken up in the emulsions may contain oil-degrading microbes, but if water is trapped in the emulsions this will not promote biodegradation on the bulk oil as the emulsions may be depleted in essential nutrients. 

5.2.1.2.6 Natural dispersion

With sufficient energy from wave action the oil may break up into droplets in the water column. If oils are easily dispersed, small droplets are generated. The rising or settling rate of the droplets is related to the size and specific gravity of particles. As an example droplets 100 µm in diameter and a lower specific gravity that the surrounding seawater have been observed to rise with a velocity of approximately 1.5 m/h. Thus, larger droplets will resurface rapidly and thin oil films (sheens) may be formed. Oil dispersion is important for biodegradation. Dispersible oils will generate relatively small droplets, resulting in large surface areas for bacterial attachment. For instance, fresh Louisiana Sweet Crude oil can be made to generate dispersions with a median droplet size of 50-150 µm under continuous breaking wave conditions in an oil-on-seawater flume experiment (Brakstad et al. 2011). Several biodegradation studies of dispersed oil in cold seawater (5-8°C) have shown bacterial colonization of oil-droplets and biodegradation of dispersible oils (e.g. Lindstrom and Braddock 2002; MacNaughton et al. 2003; Venosa and Holder 2007; Prince et al. 2012). This colonization may tend to generate flocs of oil and biomass (MacNaughton et al. 2003; Bælum et al. 2012). However, for waxy oils with high pour points, evaporation, dilution and dispersion may be reduced in cold seawater, since precipitated wax may form a matrix which limits internal mixing and acts as a diffusion barrier between the oil and the water (Faksness 2008).

5.2.1.2.7 Oil films

As described above, surface and resurfaced oil may generate thin films on the sea surface. In a series of studies with thin oil films immobilized on hydrophobic adsorbents, n-alkanes in these films were rapidly biodegraded in temperate and cold seawater (0-13°C), while aromatic compounds were subject to mixed dissolution and biodegradation (e.g. Brakstad and Bonaunet 2006; Brakstad et al. 2004). In experiments over a period of 112 days with different oil thicknesses of a wax-rich oil it was apparent that a thickness limit for measurable biodegradation (nC17/Pristane and nC18/Phytane) was between 0.1 and 1.0 mm in cold (6-10°C) seawater (Brandvik et al. 2006). 

5.2.1.3 Microbial Oil-Degrading Populations in Cold Water Environments

In the aftermath of the Deepwater Horizon incident, a large body of new information has been collected and integrated with our already existing understanding of the microbial response to oil spilled in the marine environment (Hazen et al. 2010; Mason et al. 2012; Valentine et al. 2012). In general, in situ sampling and analysis revealed unexpectedly rapid disappearance of released oil in the Gulf of Mexico environment, which is characterized by a temperate climate (Hazen et al. 2010). This rapid disappearance was affected by the prevalence of water-soluble constituents in the crude oil (Reddy et al. 2012), injection of subsea dispersant into the erupting oil flow (Kujawinski et al. 2011), and presence of indigenous oil-degrading microorganisms in this area that is well known for natural seeps of crude oil from reservoirs (Lu et al. 2012). Such indigenous oil-degrading microorganisms are the topic of this section.

Following the Deepwater Horizon incident, extensive analysis of microbial responses was done both in situ and in laboratory microcosms. These analyses support, in general, a paradigm of successive blooms of taxonomically distinct indigenous microbial populations as the oil weathers and labile components are sequentially degraded leaving less-readily degraded components to feed subsequent blooms (Hazen et al. 2010; Valentine et al. 2010; Kostka et al. 2011; Baelum et al. 2012; Beazley et al. 2012; Lu et al. 2012; Mason et al. 2012; Valentine et al. 2012).

Conditions are very different in high latitude marine environments.  As described in previous sections, the Arctic and Antarctic marine environments are characterized by seasonal extremes of photoperiod, spatial variability in salinity and temperature, as well as generally colder surface temperatures compared to the temperate latitudes. These differences may result in different expectations about the rate of oil degradation, as described in previous sections.  They also result in different expectations about the indigenous populations of oil degrading microorganisms.

As mentioned in previous sections, microbial responses to oil in marine environments generally are dominated by bacteria rather than archaea (Roling et al. 2004). Although fungi are known to degrade petroleum compounds in some marine settings (Zinjarde and Pant 2002), few surveys of fungal abundance in high latitude marine environments have been done (Butinar et al. 2011) and thus far none have addressed oil degradation by fungi in high latitude environments. For these reasons, this section focuses on the bacterial component of the marine microbiological community. 

5.2.1.3.1 Indigenous Microorganism Populations

Among the bacterial taxa catalogued in high latitude marine environments, many appear to be specific to that environment (Ghiglione et al. 2012; Sul et al. 2013).  This apparent specificity may be due to truly unique populations, or it may be a function of the limit of detection.  Community members that thrive in the high latitude marine environment grow to relatively high cell densities and are therefore more easily detected.  Various investigations have found that microbial species richness curves are not saturated with typical levels of effort.  This finding has led to the hypothesis that there is an under sampled “rare biosphere” of organisms with low population density (Sogin et al. 2006) that, despite low population levels, can respond to changes in environment and energy source.  This phenomenon may be typified by the explosion of Oceanospiralles and Colwellia populations in the presence of different partitions of spilled oil during the Deepwater Horizon incident (Hazen et al. 2010; Bælum et al. 2012).

Marine ice represents an extreme biosphere with below-zero-centigrade temperatures and high salt concentrations. It has been demonstrated by field studies that bacterial populations in Arctic marine ice are affected by oil pollution, stimulating species of a few genera like Colwellia, Marinomonas and Glaciecola (Brakstad et al. 2008). 

Not all of the microorganisms found in the Arctic oceans are adapted to that environment.  The various currents carry viable microorganisms from diverse locations to the Arctic (Rosnes et al. 1991; Hubert et al. 2009; Hubert et al. 2010); thus, there is an expectation of cosmopolitanism among the free-living microorganisms.  This is not to say that the population structure is homogenous as if the Arctic were a giant mixing bowl.  In fact, there is documented variability in population structures, with different communities associated with water masses of different origins (Galand et al. 2010; Sul et al. 2013).   The presence of non-adapted microorganisms such as thermophiles does, however, indicate that microbial populations adapted to the consumption of natural or human-induced oil releases might be transported to and be present in areas that are not commonly exposed to oil.

5.2.1.3.2 Population Effects on Oil Degradation

Crude mineral oil is degradable by indigenous microorganism populations in the Arctic marine environment, even at near-freezing temperatures (Brakstad and Bonaunet 2006), although at slower rates compared to higher temperatures (Margesin et al. 2003; Michaud et al. 2004).  Nevertheless, over a time course on the order of weeks substantial biodegradation can be observed in nutrient-enriched cold Arctic seawater (Brakstad and Bonaunet 2006). Community analysis of oil-degrading Arctic microbial consortia indicated that several taxa of bacteria are involved in biodegradation in this environment, including genera related to Pseudoalteromonas, Pseudomonas, Shewanella, Marinobacter, Psychrobacter, and Agreia (Deppe et al. 2005).  Of interest, these are different organisms from those directly associated with degradation in the Deepwater Horizon spill in the Gulf of Mexico, specifically bacteria of the orders Oceanospiralles (Hazen et al. 2010; Kostka et al. 2011) and Alteromonadales (Bælum et al. 2012), among others.

Linear alkanes often are characterized as an easily accessible carbon source, either through degradation or direct incorporation into microbial biomass, in the marine environment (Harayama et al. 1999).  The metabolic pathways for linear, branched, and cyclic alkanes have been studied and described since the 1960s (Jobson et al. 1972, Coates et al. 1997, Feng et al. 2007, Rojo 2009, Gray et al. 2011).  Preferential degradation of short-chain alkanes (represented by C15) over long-chain alkanes (represented by C26) was observed in situ in a deep plume (circa 1,400 m) in the Gulf of Mexico under aerobic conditions (Hazen et al. 2010).  Furthermore, during weathering in subsurface petroleum reservoirs, alkyl chains on substituted soluble PAHs such as alkane-substituted naphthalenes may be transformed even more rapidly than linear alkanes (Jones et al. 2008).  Whether this phenomenon, observed in anaerobic subsurface reservoirs, would occur in the presence of petroleum hydrocarbons released into the deep sea, remains unknown. 

The specific bacteria known to accomplish alkane degradation are numerous (Whyte et al. 1997; Rabus et al. 1999, Hara et al. 2003, van Beilen et al. 2004, Throne-Holst et al. 2006, Feng et al. 2007, Throne-Holst et al. 2007, Wentzel et al. 2007, Rojo 2009, Teramoto et al. 2009, Wasmund et al. 2009, Tapilatu et al. 2010, Alonso-Gutierrez et al. 2011, Teramoto et al. 2011).  Among these, many were characterized from high-latitude marine environments.  Specifically, the Pseudomonas strains isolated by Whyte et al. (1997) from Arctic soils may be transported to the marine environment via runoff.  Alcanivorax species are known to be widespread in marine environments exposed to oil (Hara et al. 2003, van Beilen et al. 2004) and, if not prevalent in the Arctic environment, might be expected to be present because of currents.  Thus, bacteria capable of alkane degradation are expected to be present in the Arctic oceans.

The ability to degrade aromatic hydrocarbons and, in particular, polynuclear aromatic hydrocarbons typically is considered to be less widespread than the ability to degrade alkanes.  For example, some organisms have diverse pathways that confer the ability to degrade polynuclear aromatic hydrocarbons, e.g., Mycobacterium vanbaalenii (Kweon et al. 2011) and various Pseudomonas spp. (Whyte et al.1997). The distribution of these genes among bacteria in Arctic marine environments remains unknown.       

5.2.1.4 Hydrocarbon biodegradation in cold marine environments

In general, biodegradation of oil compounds is expected to follow the order n-alkanes > branched alkanes > low molecular weight aromatics > cyclic alkanes (Perry 1984). In cold seawater the same order is expected, although degradation will be highly influenced by the physico-chemical characteristics of the oil. The low temperature affects both dissolution from the non-aqueous (crude oil) to the aqueous phase (Schluep et al. 2001), and evaporation of volatile compounds, as described above.  

5.2.1.4.1 Seawater

At temperatures above the freezing point of seawater (approximately -1.8°C) biodegradation of crude oil hydrocarbons is well documented. This is exemplified in Figure 5-2, showing the mineralization curves of 14C-labelled naphthalene, phenanthrene and hexadecane in seawater at 0°C when the compounds were spiked into crude paraffinic oil. Degradation of the n-alkane (hexadecane) was faster than for the aromatic compounds, and a smaller aromatic (naphthalene; 2-ring) degraded faster than larger aromatics (phenanthrene; 3-ring). This pattern followed the generally accepted order of crude oil compound biodegradation described above.

Figure 5-2. Mineralization in seawater at 0°C of 14C-labelled hydrocarbons spiked into crude oil (from Brakstad and Bonaunet 2006). No mineralization was measured in sterile controls (not shown).
Figure 5-2. Mineralization in seawater at 0°C of 14C-labelled hydrocarbons spiked into crude oil (from Brakstad and Bonaunet 2006). No mineralization was measured in sterile controls (not shown).

One of the first attempts to study oil biodegradation in Arctic seawater at low temperatures (2-11°C) showed that shifts in microbial populations towards more oil-degrading bacteria, that abiotic oil losses were lower than expected, and that various classes of hydrocarbons (saturates, mono-, di- and polyaromatics) were subject to biodegradation (Horowitz and Atlas 1977). Several studies have compared oil biodegradation in seawater or with bacterial cultures at different temperatures, and results from some of these including temperatures relevant for the Arctic are summarized in Table 5-1.

In summary, these and most other relevant studies (e.g. MacNaughton et al. 2003) show slower biodegradation by lowering of the temperature, but the results also show that biodegradation at low seawater temperature is considerable. In a recent study with low concentrations (2.5 mg/L) of Alaska North Slope oil with Atlantic seawater, 80% was biodegraded (saturates, 2- to 4-ring aromatics) after 60 days at 8°C (Prince et al. 2012). While laboratory studies indicate that biodegradation in Arctic seawater may be slower than in temperate seawater, these results have not been confirmed by field studies. Seasonal biodegradation data and comparison of oil biodegradation from different geographic areas with the same oils and analytical procedures may be necessary to test these assumptions. Oil characteristics should also be addressed in more detail, for instance, by comparison of dispersed oil biodegradation of different oil types and weathering degrees at several seawater temperatures.  The physical properties of oil may decrease bioavailability of oil (e.g. larger droplets at lower temperatures would increase the surface area-to-volume ratio).

Table 5-1. Summary of selected biodegradation studies performed at different seawater temperatures 

OilsInoculaTime (days)ComponentsTemp. (°C)ResultsReferences

Fresh Prudhoe Bay crude (dispersions)

Mixed consortium

28

nC10-nC35 alkanes and 2-4 ring aromatics

20

A)K1=0.13-0.23 (t½=3-5 days)

Venosa and Holder 1997

5

A)K1=0.052-0.093 (t½=7-13 days

Weathered Alaska North Slope (dispersions)

Mixed consortium

90

GC-MS detectable

20

61.5 % biodegradation

Garrett et al. 2003

6

48 % biodegradation

Diesel oil (dispersions)

Two Antarctic strains

60

GC-FID detectable

20

75-86 % biodegradation

Michaud et al. 2004

4

55-58 % biodegradation

Fresh Statfjord oil (immobilized films)

Natural seawater

56

nC10-nC36 alkanes

5

95 % biodegradation

Brakstad et al. 2006

0

32 % biodegradation

Arabian light crude oil (dispersion)

Natural Antarctic seawater

50

nC17/Pristane ratio

20

B)40 % reduction

Delille et al. 2009

10

B)47 % reduction

4

B)20 % reduction

A) k1 is first-order rate coefficient; t½ is half-life (0.69/k1) B) Reduction determined by comparison to sterile controls

5.2.1.4.2 Sediments and soils

Several biodegradation studies of oil in Arctic sediments have been conducted, most of these to investigate the potential for bioremediation of stranded oil in the Arctic (see later chapter). Studies on oil pollution of Arctic and Antarctic beaches has demonstrated the presence of indigenous hydrocarbon-degrading bacteria in these pristine environments (e.g. Grossman et al. 2000; Delille and Delille 2000; Powell et al. 2005). Oil removal from beach sediments may be attributed to several processes, including physical removal, photooxidation and biodegradation. For instance, significant depletion of total hydrocarbon concentrations and mineralization of radiolabelled hexadecane have been measured in Canadian Arctic soils at 4°C (Greer 2008). Anaerobic biodegradation has also been measured in the Arctic. Low-temperature degradation of PAH-compounds was reported from Arctic soils under anoxic and nitrate-reducing conditions at 7°C (Eriksson et al. 2003). In the Arctic winter the upper parts of marine sediments become frozen.  Whether biodegradation stops or continues at very slow rates under these conditions is not known, although microbial activity at subzero temperatures has been demonstrated (Doyle et al. 2012).  Several studies with oil-contaminated freeze-thaw cycled soil or permafrost have shown that microbial respiration takes place even at subzero temperatures and hydrocarbon degradation was observed (Rike et al. 2003; Børresen et al. 2007; Chang et al. 2011).  These studies therefore demonstrate that the lower limit for biodegradation can be below the freezing point.

5.2.1.4.3 Sea ice

If oil spills reach the marginal ice zone, the ice may become oil-infested. Once trapped within the ice, ocean currents can transport the oil over large distances.  A secondary discharge situation occurs during the spring melt season and, if the ice has been transported from the original spill site, this can result in contamination of new locations. In the spring and summer seasons, chemical alteration of the crude oil through photooxidation may also become an important process (Refer to Section 3).  Although the immediate impact of oil spills in ice has been studied (e.g. Fingas and Hollebone 2003) and is fairly well understood, little is known about the long-term fate and effects of such pollutants on ecosystems in polar environments.  To date, few studies have attempted to determine the transport and fate of individual water-soluble oil components in sea ice.  However, data from some recent studies have shown that the more water-soluble compounds (mainly naphthalenes, phenanthrenes and dibenzothiophenes) migrate through the brine channels in the ice (Figure 5-3).  As a result, such compounds come into contact with sea ice microbes in the brine and the underlying water (Faksness and Brandvik 2008a; Faksness and Brandvik 2008b).

In line with the results from studies with Arctic soils one should expect that biodegradation may also take place in marine ice at subzero temperatures. As described earlier, microbial metabolism and motility have been measured in the brine channels of marine ice (Breezee et al. 2004; June et al. 2002; Junge et al. 2003; Junge et al. 2004; Junge et al. 2006; Mykytzuk et al. 2013).  However, biodegradation of oil in marine ice has not yet been fully investigated. In a winter field study (February to June) performed on Svalbard with crude oil frozen into fjord ice, a slow reduction in the ratio between naphthalene and phenanthrene was measured in the parts of the ice with downward migration of soluble compounds, while no significant change in n-C17/Pristane was measured, as shown in Figure 5-3 (Brakstad et al. 2008). However, the bulk oil stimulated bacterial biomass, including a few bacterial genera expected to be oil-degraders (Brakstad et al. 2008). The results from another field study performed at Svalbard showed that no significant degradation of oil hydrocarbons occurred in the ice at subzero temperatures, but at 0°C melt pool oil samples fertilized with inorganic nutrients showed a significant change in bacterial diversity (Gerdes and Dieckmann 2006). Marine ice represents an extreme environment for life.  The combination of low temperature and high salt content in the brine channels require that microbes be both halo- and psychro-tolerant.  Extremely halophilic or halotolerant microbes able to degrade oil have been reported (e.g. Diaz et al. 2002, Al-Mailem et al. 2010), but not so far in cold environments.  However, as described above, it has been demonstrated that oil pollution in marine ice may stimulate the growth of a few specific bacteria (Brakstad et al. 2008), but the ability to degrade oil compounds needs to be clarified. In addition, most oils will also be solidified under these conditions, but the migrating water-soluble compounds may be relevant target compounds for oil-degrading bacteria in this environment. If this is true, bacteria able to degrade small aromatics may be more relevant than alkane-degrading bacteria. 

Figure 5-3. Oil migration and degradation in marine ice. [The middle core shows the migration of different oil components through an ice core with the chromatograms of the components in the left panel and the relative ratios of specific components in the right panel (from Brakstad et al. 2008)].
Figure 5-3. Oil migration and degradation in marine ice. [The middle core shows the migration of different oil components through an ice core with the chromatograms of the components in the left panel and the relative ratios of specific components in the right panel (from Brakstad et al. 2008)].

5.2.1.5 Modeling of biodegradation
5.2.1.5.1 Biodegradation in oil spill models

Several oil spill models have been developed during the last decades.  Most of these are physical models which can be separated into oil weathering models, trajectory models (predicts the route of an oil spill), or stochastic models (describing an impact area of an oil spill).  Examples of well-known models are the Oil Spill Contingency and Response (OSCAR) and the OILMAP models (www.sintef.no/Materialer-og-kjemi/Marin-miljoteknologi/Miljomodellering/Modellverktoy/OSCAR-Oil-Spill-Contingency-And-Response/; http://www.asascience.com/software/oilmap/index.shtml). Models have also been presented for predictions of oil behavior in ice-infested water (Drozdowski et al. 2011). Most of these models are physical models, but the OSCAR model also incorporates biodegradation of 25 pseudo oil compound groups, in addition to descriptions of the physical environment, physical-chemical fate processes and ecotoxicity (Aamo et al. 1997; Reed et al. 2000). In the OSCAR model, which is an industry standard in Norway, biodegradation is one of the fate processes together with physico-chemical processes like advection, spreading, evaporation, dispersion, dissolution, particle adsorption/dissolution, volatilization from water column, and seabed contamination. An example of vertical oil concentrations and mass balance after a simulated 60-day blowout is shown in Figure 5-4. However, biodegradation as part of the mass balance may be overestimated in the model, since degradation is determined on the bases of biotransformation, not complete biodegradation, and only compounds determined by gas chromatography-mass spectrometry (GC-MS) analyses are included.  

Figure 5-4. Simulation of a deep water blowout by the OSCAR model. [The left figure shows the oil concentration vertically in the water masses during a simulated 60-d blowout from 1600 m depth with a light paraffinic oil. The right figure shows the mass balance between different fate processes during the blowout period (from Brakstad <em>et al.</em> 2011).]
Figure 5-4. Simulation of a deep water blowout by the OSCAR model. [The left figure shows the oil concentration vertically in the water masses during a simulated 60-d blowout from 1600 m depth with a light paraffinic oil. The right figure shows the mass balance between different fate processes during the blowout period (from Brakstad et al. 2011).]

5.2.1.5.2 Biodegradation modeling and temperature

In the oil spill models biodegradation must be predicted at different temperatures. Oil biodegradation data in Arctic environments with cold seawater are limited, since most published studies have been performed at higher temperatures than relevant for these environments. In order to transform degradation data between different temperatures, plots have been used to transpose results of bacterial metabolism and growth at different temperatures. Temperature-related bacterial growth rates may be estimated by using modified Arrhenius plots (Arrhenius, 1889).  Ideally, Arrhenius plots should show temperature-related linearity.  However, in a study with psychrotrophic toluene-degrading strains of Pseudomonas putida grown on toluene or benzoate, growth rates had to be fitted using two linear segments at a temperature range of 4-30°C: one segment above and one below 17-20°C (Chablain et al. 1997).  When using Arrhenius plots, the temperature range should therefore not be too broad.  For water-soluble compounds, temperature-dependent biodegradation has been suggested to follow a Q10-value, which is a relationship describing the degradation rate increases when temperatures are raised by 10°C increments.

 Some description

Equation 1

Where R is the general gas constant (8,314·10-3 kJ/mol·K), Ea is the activation energy (kJ/mol), T1 is the reference temperature in Kelvin and T2 is the actual temperature in Kelvin.  According to this approach, the degradation rates should double for every 10°C increase, resulting in an ideal Q10 of 2.0.  The Q10–values for the biodegradation of oil hydrocarbons in seawater were determined with a heavy fuel oil (Bunker C), and with winter or summer water samples from the North Sea.  When incubation temperatures of 4-18°C were used, Q10–values of 2.4 and 2.1 were determined for waters in winter and summer, respectively, where biodegradation was measured as biological oxygen demand (Minas and Gunkel 1995).  Calculations of Q10-values from a variety of studies have shown that the rule-of-thumb value (Q10 = 2.0) is a fairly good approximation in a temperature range of 5–27°C (Andrea Bagi, personal communication).  However, for the narrow range and freezing temperatures of the Arctic the expectation that calculated Q10 would remain close to 2 may not be valid.  For instance, a calculation of Q10 in immobilized oil films (Statfjord B oil) based on data for 5 and 0°C (Brakstad and Bonaunet 2006) showed a value of 16.2 (Andrea Bagi, personal communication).  This may be caused by changes in the physico-chemical characteristics of this oil at these temperatures.  Thus, changes in oil characteristics at low seawater temperatures may affect the biodegradation models, and therefore predictions of oil degradation rates in Arctic seawater will require closer examination. 

5.2.1.6 Determination of Biodegradation
5.2.1.6.1 Analytical methods for oil compound analyses

Since oil consists of thousands of different compounds (Marshall and Rogers 2003) measurements of individual compounds is a challenge.  Bulk oil biodegradation may be determined by traditional gravimetric analyses (e.g. Horowitz and Atlas 1977), while broader groups of oil components (saturates, aromatics, resins and asphaltenes = SARA) may be determined by Iatroscan thin-layer chromatography with flame ionization detection (TLC-FID; Stevens 2004).  Using this method, crude oil components are determined according to their polarity.  The saturate fraction consists of nonpolar material including linear, branched, and cyclic saturated hydrocarbons (paraffins). Aromatics, which contain one or more aromatic rings, are slightly more polarizable.  The remaining two fractions, resins and asphaltenes, have polar substituents.  Additional bulk oil analytical methods include Fourier Transform Infrared (FTIR) spectroscopy and Nuclear Magnetic Resonance (NMR) spectroscopy.  FTIR is an absorption technique that uses infrared (IR) electromagnetic radiation to examine the identity of chemical bonds within the substance of interest.  As microbial degradation of the oil is expected to result in the addition of oxygen atoms into the structure of oil compounds this method may be a method for measuring bulk changes in composition, although the resolution and sensitivity is poor compared to other methods.  NMR is a nondestructive technique that is well-suited for identifying and quantifying different hydrocarbon classes and can provide information on the relative content of aliphatic, olefinic, and aromatic components.  Studies have shown that NMR spectra in conjunction with multivariate statistical analysis can be correlated to a number of physicochemical properties and standard distillation cut yields (Molina et al. 2007).  Mass spectrometry (MS) has become one of the most important detection principles in modern analytical chemistry.  The principle behind MS is that molecules can be identified through their molecular weight and fragmentation patterns.  MS is very often connected to a separation step, usually gas (GC) or liquid (LC) chromatography.  These methods may be used to identify and quantify targeted oil compounds or for fingerprinting of complex chemical mixtures.  To separate between different oil compound groups gas chromatographic analyses (GC-FID and GC-MS) are the standards today, but these methods favor detection of nonpolar compounds.  The common use of these methods therefore limits our knowledge of oil biodegradation, mainly to some compound groups, such as the C10-C40 saturates, cyclic saturates (decalines), BTEX, phenols, 2-6 ring PAHs, and a variety of biomarkers.  LC-MS analyses may therefore be an important supplement to the gas chromatographic analyses for more polar compound groups.  In addition, biodegradation studies of compounds like naphthenic acids have been of interest in specific areas like Canada.  Several high-resolution instruments, like time-of-flight mass spectrometers (ToF-MS) coupled to GCxGC systems (GCxGC-ToF-MS)(e.g. Tran et al. 2010) and Fourier transform ion cyclotron resonance (FT-ICR) mass spectrometer (e.g. Hughey et al. 2008) provide powerful techniques for the analytical separation of complex mixtures combined with methods for characterizing the resolved compounds. Minor components hidden in the large background can be detected by these instruments, and both resolution and sensitivity allow for searching of spectra from very narrow peaks.  For instance FT-ICR MS can separate masses of <0.002 Dalton of compounds that contain heteroatoms such as N, O, S and other elements, identifying oil compounds by mass and molecular formula at high resolution.

5.2.1.6.2 Experimental apparatus

Advances in microbial sampling capabilities, in particular sampling of the ocean in drilling areas, came with advances in drilling technology.  The Ocean Drilling Project (ODP) and subsequent Integrated Deep Ocean Drilling Program (IODP 2003-2013) and planned International Ocean Discovery Program (IODP 2013-2023) provide a framework for these activities (Edwards et al. 2012).  Each of the named programs includes or will include a sampling component for microbial ecology research.  In particular, the 2003 IODP included extensive evaluations of seafloor and sub-seafloor microbial communities (Cyranoski 2003).  In conjunction with new microbiological techniques, these samples provided new perspectives on deep ocean microbial community composition and function (D'Hondt et al. 2004; Schippers et al. 2005; Inagaki et al. 2006; Biddle et al. 2008; Kobayashi et al. 2008; Forschner et al. 2009; Lomstein et al. 2012).  Understanding native populations in Arctic drilling fields requires sampling such as has been carried out in these programs.

Much of the sampling that is associated with drilling activities focuses on the deep subsea floor while water column and sediment samples can be collected with remote samplers or can also  be collected by autonomous underwater vehicles (AUVs).  The Chemosynthetic Ecosystem Science (ChEss) project of the Census of Marine Life (2002-2010) was one such project that generated a substantial amount of new information about marine microbial communities.  Much of the success of the ChEss project was attributed to the development of improved deep-ocean AUVs (German et al. 2011) that allowed systematic exploration of previously understudied areas, including cold seeps.  Modern AUVs are capable of rapid deployment and operation at a range of depths.  They have been effectively deployed to sample in response to events such as the Deepwater Horizon spill of 2010 (Camilli et al. 2010).  These vehicles contribute to the ability to observe natural processes and conduct in situ experiments, particularly at depth in harsh marine environments.

Another strategy is to employ microbial observatories in marine environments that incorporate real-time sensors, time-lapse cameras, and other experimental devices.  These observatories, along with autonomous and cabled sensors, allow direct measurement of microbial processes in the deep ocean.  In particular, beginning some 20 years ago, circulation obviation retrofit kits (CORKs) came into use to study connectivity of hydraulics and biogeochemistry at the interface of the ocean bottom and open water (www.corkobservatories.org; Cowen et al. 2003) rather than relying on extrapolation from controlled laboratory experiments (e.g. Bartlett 2002; Tapilatu et al. 2010) or inference from population composition (Simonato et al. 2006) as is more commonly done.  Another type of observatory, the Microbial Methane Observatory for Seafloor Analysis (MIMOSA), is an autosampler that collects and archives microbial material for later recovery and analysis.  Two of these devices recently were deployed in the Gulf of Mexico to evaluate petroleum seeps and spills as they affect microbial population structure (Balinski 2012).  This type of observatory may be useful to implement in situ experiments to monitor biodegradation rates and processes and further advance knowledge of petroleum hydrocarbon degradation processes in the deep ocean environment.

Finally, another tool that is directly relevant to petroleum biodegradation and carbon utilization is the so-called “bug trap,” in which hydrophobic beads or woven matrix is dosed with petroleum hydrocarbons to evaluate in situ degradation potential and analyzed to characterize degrading community composition.  Because petroleum-degrading microorganisms can be chemotaxic to suitable substrates, these experimental devices can be used to attract and study degraders in the laboratory (Brakstad and Bonaunet 2006) and in situ experiments (Raloff 2010; DeAngelis et al. 2011).

5.2.1.6.3 Biodegradation data processing

In standard laboratory studies, oil degradation is usually determined by comparison of depletion in normal seawater or cultures to depletion in sterile (killed) controls. In this way processes like evaporation, wall effects, dissolution of compounds from the oil phase etc. may be accounted for and separated from the biodegradation process. However, in field and meso-/large-scale studies biodegradation is determined by normalization of degradable compounds to less degradable (recalcitrant) compounds. Common compounds for this internal normalization are pentacyclic triterpane biomarkers (e.g.C3017α(H),21β(H)-hopane) and the isoprenoids pristane and phytane (Prince et al. 1994; Douglas et al. 1996; Page et al. 1996). The isoprenoids have proven to be biodegradable themselves, although at slower rates than their corresponding n-alkanes (e.g. Douglas et al. 1996). Hopanes also have limitations if used to determine biodegradation of compounds with low boiling points, since it may be difficult to separate biodegradation from evaporation. In addition, determination of biodegradation as ratios between biodegradable and more persistent compounds has also been suggested using other compounds, like 2-methylphenanthrene/1-methylphenanthrene, and C3-phenanthrene/C3-dibenzothiophene (Fedorak and Westlake 1981; Christensen and Larsen 1993; Wang et al. 1998; Lamberts et al. 2008).

5.2.1.7 Persistent Oil Compounds

Most environmental studies of petroleum-derived chemicals have focused on effects related to specific and easily identified hydrocarbons such as n-alkanes, BTEX and PAHs. However, in the case of environmentally weathered samples, most oil compounds appear as an unresolved complex mixture (UCM in gas chromatograms, and are often referred to as the “hump” (see Figure 5-5). Having undergone a variety of weathering processes (e.g. evaporation, biodegradation and photooxidation), this residual UCM is comprised of thousands of environmentally persistent compounds (Gough and Rowland 1990, Killops and Al-Jaboori 1990). In fact, it has been established that natural biodegradation of spilled crude oil leads to a significant increase in the UCM  concentration relative to other crude oil components (e.g. Meredith et al. 2000), highlighting the persistence of these compounds. Fractionation and subsequent characterization studies have shown that both non-polar (e.g. aliphatic and aromatic) and polar (resin and asphaltene) compounds contribute to crude oil UCMs.  Both the aromatic hydrocarbon and polar UCMs have been shown to bioaccumulate in marine organisms and elicit ecotoxicological responses and impaired health (e.g. Farrington et al. 1982; Widdows et al. 1995; Barron et al. 1999; Smith et al. 2001; Rowland et al. 2001; Donkin et al. 2003). The polar UCM fractions comprise compounds containing highly polar N, S, and O atoms in their structures (so-called NSO compounds).  Many of these compounds (e.g. phenols and naphthenic acids) are thought to be homologous in structure to compounds present in the non-polar fraction of the UCM, hence their resistance to biodegradation.  Due to their persistence these compounds may reach other locations. 

Figure 5-5. Gas chromatogram of the aromatic hydrocarbon UCM fraction extracted from mussels (Mytilus edulis) from a petroleum impacted site (Southend) on the UK coast (From Booth et al., 2007a).
Figure 5-5. Gas chromatogram of the aromatic hydrocarbon UCM fraction extracted from mussels (Mytilus edulis) from a petroleum impacted site (Southend) on the UK coast (From Booth et al., 2007a).

The biodegradation potential of these seemingly persistent UCM compounds is further complicated by the environmental conditions prevalent in Arctic regions.  Lower ambient temperatures will result in reduced biodegradation rates, and currently nothing is known about the abilities of psychrophilic or psychrotrophic bacteria to degrade these compounds. In temperate regions, microbial communities from previously oil-impacted sites have been shown to partially degrade model UCM compounds, such as alkylcyclohexyltetralins, alkylcyclohexylnaphthalene and naphthenic acids (Scott et al. 2005; Booth et al. 2007b; Frenzel 2008).  Recent studies using naphthenic acids showed that it was the molecular structure rather than the number of carbon atoms that was important for determining biodegradation.  Specifically, the most recalcitrant compounds included those with relatively high degrees of alkyl branching (Han et al. 2008).  During future degradation studies of oil compounds in cold environments it is therefore of major importance to consider these environmentally persistent and toxic UCM-related compounds.  

5.2.2 Accelerated Biodegradation

Naturally occurring hydrocarbon-degrading bacteria are found in all environments.  Although these bacteria are capable of initiating the biodegradation of spilled crude oil, attempts have been made to increase removal efficiency through bioremediation strategies.  Most bioremediation attempts have focused on developing good biostimulation strategies, typically by applying degradation rate-limiting nutrients, or the combination of these and other treatments, to accelerate the natural biodegradation processes.  Bioremediation processes, if successful, are cost-effective and reduce the environmental impacts of marine oil spills (Prince 1993; Swannell et al. 1996; Prince and Clark 2004; Prince 2005; Prince and Atlas 2005).  An alternative to biostimulation is bioaugmentation which involves the inoculation of indigenous or exogenous microbial cultures with high biodegradation potentials for contaminants.  Bioaugmentation approaches have been reported to improve biodegradation of hydrocarbons from oil spills in cold soil or marine sediments, and can be used in combination with fertilizers (Margesin and Schinner 1997; Ruberto et al. 2003).  Bacterial mats from marine oil-contaminated sites have also been suggested for use in the degradation of coastal oil spills, although these are of greater relevance for spills in temperate areas (Cohen 2002).  Various methods and strategies for bioremediation have been reviewed (e.g. Lee and Merlin 1999; Prince 2010).

To date, most experimental oil bioremediation studies in Arctic or Antarctic environments have been conducted on stranded oils, employing the application of fertilizers to stimulate the indigenous flora, and often in combination with mechanical treatments which improve oxygen and nutrient availability (e.g. Sveum and Ladousse 1989; Prince et al. 2003a; Obbard et al. 2004; Pelletier et al. 2004). Bioremediation strategies have also been applied to real oil spill situations such as the Exxon Valdez accident in 1989, where it formed part of a beach cleaning strategy (Bragg et al. 1994).

Also the use of chemical dispersants is regarded as an effective way of stimulating biodegradation of oil.  The dispersants work as surfactants, changing the surface characteristics of the oil, and reducing the droplet size.  Chemical dispersants have been used primarily on surface oil spills, but were also injected directly (Corexit 9500) at the wellhead during the Deep Water Horizon (DWH) spill (Atlas and Hazen 2011).

5.2.2.1 Biostimulation

Biostimulation includes the addition of nutrients or other methods to enhance the capability of the indigenous microbial communities to degrade pollution components.  Biostimulation has been regarded as a cost-effective strategy for secondary cleanup mainly of stranded oil pollution.  However, other methods may also be regarded as biostimulatory actions, for instance the use of chemical dispersants for oil spills in seawater, since this may aid in oil biodegradation by increasing the oil surface area accessible to the oil-degrading microbes.

5.2.2.1.1 Shoreline sediments

Most biostimulation activities have focused on stranded oil, with application of fertilizers to increase natural degradation by the indigenous microbial flora.  Biostimulation treatment is often combined with mechanical treatment to improve oxygen and nutrient availability.

In marine environments some growth- and biomass-stimulating factors are essential for oil biodegradation, especially nitrogen and phosphorous and the addition of these nutrients are common practice in bioremediation.  Balanced nutrient availability is important for biodegradation and the composition of hydrocarbon-degrading communities, since nutrient amendments, in some instances, can inhibit microbial activity (Braddock et al. 1997).  It is therefore important to avoid excess nutrients, which can cause detrimental effects, such as eutrophication.  During biostimulation, molar carbon/nitrogen/phosphorous ratios of 100/10/1 have often been used (e.g. Bouchez et al. 1995; Obbard et al. 2004).  However, results from laboratory studies have also shown that certain microbial populations may require different N/P ratios for optimal degradation of different hydrocarbons (Smith et al. 1998). Nutrient products are available as briquettes, granules or liquid fertilizers.  Liquid inorganic fertilizers have proven effective but require frequent application, and therefore oleophilic slow-release nutrient formulations have been developed, which promote hydrocarbon degraders at the oil-water interface.

For improved results bioremediation may be combined with other clean-up procedures.  Surf washing and the use of surfactants may increase the surface area of the oil and hence increase oil degradation. Ex situ technologies like land farming (spreading the polluted sediments over a larger area for better oxygenation), composting and biopiling may be used for treating oily waste during spill treatment (Lynch and Moffat 2005) although these approaches have seen limited application in polar environments.

Several field biostimulation trials have been conducted on Arctic beaches at Spitzbergen, either in Ny Ålesund (78° 55' N, 11° 56' E) or in the Van Mijen fjord close to the small mining society Svea (77°56’N, 16°43’E).  Experiments performed by SINTEF and the oil company ELF in the 1980s in Ny Ålesund with the slow-release oleophilic fertilizer Inipol EA22 indicated that application of the fertilizer to oil in beach sediments resulted in increased biodegradation in coarse sediments, but not on oil in finer sediments (Sveum and Ladousse 1989).  During a full-scale trial at the ITOSS (In Situ Treatment of Sediment) program in 1997 several remediation processes were tested on intermediate fuel oil (IF-30) artificially stranded on mixed (sand and pebble) intertidal shorelines.  The remediation methods included sediment relocation (surf washing), mixing (tilling) and bioremediation (Guenette et al. 2003; Sergy et al. 2003).  The fertilizers used included both soluble (prilled ammonium nitrate and superphosphate [Ca(H2PO4)2]) and commercial slow-release (Inipol SP1) chemicals, applied to the top of the sediments during the first two months of the experiment.  The introduction of fertilizers resulted in elevated levels of bioavailable nitrogen and phosphorous in oiled sediments.  The biodegradation rates were approximately doubled over a period of one year in the oiled sediments that received fertilizers when compared to non-treated oiled sediments, and no acute toxicity was associated with the bioremediation treatment (Prince et al. 2003).  Mixing/tilling also seemed to result in increased microbial activity for limited periods by increasing sediment permeability (Owens et al. 2003).

Biostimulation field experiments have also been conducted in Antarctic environments. An Arabian crude oil was added to several 1 m2 enclosures on intertidal sandy beaches on the main island of the Kerguelen Archipelago (49°19’S, 69°42.5’E).  Different fertilizers were added to the top of the oil, including the slow-release Inipol EAP 22 and various experimental mixtures consisting of dry fish compost, with or without supplements of urea, phosphate and charged or neutral surfactants (Pelletier et al. 2004).  During a 300-day experiment the oil was eventually depleted in both untreated and treated sediments in this cold environment (seawater temperatures 3-4°C), but the various fertilizers accelerated the biodegradation rates.  It was also observed that a fertilizer with a neutral surfactant reduced the toxicity of the oil during the last three months of the experiment (Delille et al. 2002).

Bioremediation was used as an oiled-beach cleaning technology on a full-scale oil spill in Arctic environments during the Exxon Valdez accident in March 1989.  This spill in Prince William Sound, Alaska, resulted in the release of 41 million liters of Alaskan North Slope crude oil.  Bioremediation was used extensively, employing the fertilizers Inipol EAP 22 and Customblen (slow-release granulated fertilizer).  Approximately 50,000 kg nitrogen and 5,000 kg phosphorous were applied to the shorelines over the summers of 1989-1992 (Bragg et al. 1994).  For a low-energy beach containing both surface and subsurface oil and treated with both fertilizers, it was estimated that the fertilizers enhanced oil biodegradation by 5.5 times over non-treated controls (Bragg et al. 1994).

Results from this bioremediation field experiment, supplemented with several laboratory experiments have been summarized by Pritchard et al. (1992):

  • Fertilizers with slow release nutrients were recommended in tidal zones to avoid rapid wash away of the fertilizer components.  Solid granulated fertilizers were easier to apply over large areas, and adhered well to the oiled beach material. These fertilizers persisted over periods of 2-3 weeks after application.  Also fertilizer briquettes were tested with outcomes similar to the granules.  Burial of fertilizer material parallel to the water line has also been suggested.
  • Components of the nitrogen and phosphorous from liquid Inipol EAP 22 seemed to be released rapidly when submerged in seawater.  However, residual nutrients remained and were available for oil-degrading bacteria.  Laboratory experiments indicated that the nitrogen part of the oleophilic fertilizer was more important than the phosphorous compounds for biodegradation.
  • Inspection of areas treated with Inipol EAP 22 showed faster visual disappearance of oil from cobble surfaces than untreated surfaces, and laboratory experiments indicated biodegradation as the cause for disappearance rather than washing.  However, oil under the cobbles remained. It was also suggested that Inipol could work as a surfactant by chemically removing aliphatic hydrocarbons from samples.
  • Measurements of bioremediation effectiveness should include analyses of several hydrocarbon groups, in addition to the aliphatic hydrocarbons.  Measuring compositional changes in the aromatic fraction by GCMS was therefore added as a further dimension to biodegradation.  Bioremediation should also result in removal of oil residues, measured as a reduction in total mass of the oil. 
  • The use of common indicators of biodegradation like nC18/Phytane was problematic since phytane in some instances disappeared at the same rate as nC18 alkane (phytane-degrading microbes were actually isolated from beach material).  For aromatics a number of methyl-substituted homologues close in mass number to their parent structures were selected, and these were normalized to hopane, which are resistant to biodegradation.
  • Results from oil compound analyses showed that reduction in nC18/Phytane ratios resulted in corresponding changes in aromatic and heterocyclic hydrocarbons.  Samples with reduced nC18/Phytane ratios also showed increased concentrations of compounds associated with asphaltenes and polar materials, which may be partially biodegraded material.  Further, there was a positive correlation between nC18/Phytane reduction and removal of residual oil, determined by mass.  It was therefore suggested that reduction in oil residues could be used as a measure of biodegradation.

Interactions between stranded oil and mineral fines have been used to reduce oil adhesion to solid surfaces and generate stable oil droplets released to the water column.  In this way the increased surface area of oil became more accessible to nutrients, oxygen and bacteria, resulting in increased microbial activity and oil biodegradation (Lee et al. 1996).

5.2.2.1.2 Seawater

Efforts to stimulate crude oil biodegradation in seawater and ice (see next section) have not been investigated to the same extent as for stranded oil.  Most remediation strategies in oiled seawater have focused on mechanical removal methods such as the use of oil booms and skimmers.  In addition to mechanical methods, chemical dispersants have also been widely used as an alternative treatment. Dispersants are mixtures of surface-active chemicals that reduce the surface tension of the oil, resulting in the formation of oil droplets that are smaller than those initially generated by mechanical wave action (Brandvik 1997). These chemicals are primarily used to disperse oil spilled on the seawater surface into the water column.  This approach aims to reduce the impacts of oil spills on seabird and mammal populations in the vicinity of the spill and help prevent the oil reaching the coastline.  In addition, this process increases the oil-water interface and generates more bioavailable surface area for microorganisms.

The efficiencies of chemical dispersants on crude oil degradation at low seawater temperatures have shown conflicting results. In a seawater mesocosm experiment (3.5 m3 flow-through tanks with natural seawater from the St. Lawrence Estuary, Quebec) with temperatures of ‑1.8 to 5.5°C, Forties and Western Sweet Blend crude oils were treated with the dispersant Corexit® 9527 or different surfactant mixtures.  Over a 63-day period at water temperatures >0°C, chemical dispersal was found to result in higher biodegradation rates than in untreated oil samples (Siron et al. 1995).  In microcosm studies with a seawater temperature of 8°C, MacNaughton et al. (2003) added the dispersant Corexit 9500 to Alaskan North Slope (ANS) crude oil.  The dispersant resulted in rapid colonization of oil droplets by bacteria, and heterotrophic and oil-degrading microbes proliferated in the microcosms.  However, when the total hydrocarbons, C11 to C35 n-alkanes, or the sum of selected aromatics were assessed over a degradation period of 35 days, the addition of dispersant resulted in only slow or negligible biodegradation of the oil when compared to naturally dispersed oil.  In another biodegradation experiment, Lindstrom and Braddock (2002) exposed cultures of oil-degrading microbes to ANS (fresh or evaporated and spiked with radiolabelled hydrocarbons) dispersed with Corexit® 9500 over a 2 month period at 8°C.  Respirometric analyses were conducted to determine 14CO2 mineralization of 14C-labelled dodecane, hexadecane, 2-methyl-naphthalene and phenanthrene. The dispersant was found to inhibit degradation of some of the hydrocarbons (hexadecane and phenanthrene), while others (dodecane and 2-methyl-naphthalene) were unaffected when compared to mineralization of the oil without dispersant.  It was suggested that carbon mineralization, at least initially, was the result of dispersant mineralization rather than degradation of the oil compounds. In a biodegradation study comparing two dispersants, Corexit® 9500 and JD2000, fresh Prudhoe Bay crude oil was mixed with the dispersants and a microbial culture originating from the shorelines contaminated during the Exxon Valdez accident.  Biodegradation of n-alkanes and PAHs was measured over a period of 46 days at 5°C.  The first-order degradation rates of most n-alkanes and PAHs were found to be higher with dispersants than without dispersants, although these data were not statistically significant.  When the same experiment was performed at 20°C the degradation rates of n-alkanes and PAHs were more rapid than at 5°C, showing temperature-related biodegradation rates both in non-dispersed and dispersed oil. Furthermore, the influence of the two dispersants on biodegradation differed between the two degradation temperatures (Venosa and Holder 2007).  Correlation studies between the droplet surface area of dispersed oil and the resulting degradation indicated that both dispersed area and dispersant chemistry controlled the degradation and that the surfactant blend hydrophile-lipophile balance and treatment levels were also significant controlling factors (Varadaraj et al. 1995).  Recently, Prince and Butler (2013) suggested that the variability in results may result from low oil concentrations in a physically dynamic experimental system which would act to form only slightly larger oil droplets than dispersed oil treatment.  Additionally, the authors indicate that an accelerated rate of biodegradation occurs when a dispersed oil treatment is compared to biodegradation of a surfaced oil, principally due to relatively small droplet size.

A variety of bacteria and yeasts also produce biosurfactants like rhamonolipids, sophorolipids and surfactin.  Biosurfactants consists of fatty acid hydrophobic parts and carbohydrate, amino acid, cyclic peptide, phosphate, carboxylic acid or alcohol as hydrophilic part (Mulligan 2005). Most biosurfactants are produced from hydrocarbon substrates (Syldatk and Wagner 1987).  It has been suggested that rhamnolipis addition can enhance biodegradation of hydrocarbon mixtures in liquid systems and soil (Maier and Soberon-Chevez 2000).  Two mechanisms for enhanced biodegradation were proposed; enhanced substrate solubility and interactions with the cell surface to increase the hydrophobicity of the cell surface, allowing improving association of hydrophobic hydrocarbons (Mulligan 2005).  Combined use of rhamnolipids and slow-release fertilizers (Inipol EAp-22) also enhanced biodegradation of aromatic and aliphatic hydrocarbons in liquid phase and soil (Churchill et al. 1995).  Studies withRhodococcus sp. Q15 grown on hexadecane or diesel fuel at 5°C showed production of biosurfactants at low temperature (5°C), indicating that the cell surfaces became more hydrophobic  (Whyte et al.1999).

In addition to chemical dispersants use of fertilizers to increase oil compound biodegradation has been investigated to some extent.  As early as in the 1970s enhanced biodegradation of oil spills by lipophilic slow-release fertilizers was investigated (Olivieri et al. 1975).  In a study with hydrocarbon-degrading bacteria (Alcanivorax sp.) and an open seawater-based system at 30°C guano was used as fertilizer.  It was demonstrated that commercial guano was an effective source of nitrogen and phosphorus for the growth of bacteria on crude oil, and that the guano resulted in extensive biodegradation of crude oil (n-alkanes (C10- C36) and polyaromatics at (Knezevich et al. 2007).

5.2.2.1.3 Marine ice

Bioremediation of oil in ice is an intriguing prospect. If biodegradation of crude oil could be stimulated in ice, especially for the most toxic compounds migrating out of the ice through the brine channels, this would be of benefit for organisms inhabiting the polluted ice or nearby areas.   Studies have shown that fertilizers can stimulate biodegradation of crude oils in cold seawater under controlled experimental conditions (Delille et al. 1998). The slow-release oleophilic fertilizer Inipol EAP 22 was added to Antarctic seawater contaminated with “Arabian light” crude oil in a mesocosm study.  The experiment was completed over 5 weeks during the Austral summers of 1992/1993 and 1993/1994. In both ice-covered and ice-free seawater, the addition of the fertilizer enhanced both the concentrations of heterotrophic and hydrocarbon-degrading bacteria and increased the rate of biodegradation during the experiments, measured as n-C17/Pristane and n-C18/Phytane ratios.

A winter field experiment was conducted at Svalbard in 2004 as part of the Arctic Operational Platform (ARCOP) program.  Crude Statfjord oil with and without fertilizers (mixture of Inipol EAP 22 and fish meal) was placed in fjord ice (Van Mijen Fjord, Svea) for a period of 6 months (December 2004-June 2005).  At sub-zero temperatures no significant degradation of oil hydrocarbons occurred with the addition of nutrients, but at 0°C melt pool samples fertilized with inorganic nutrients showed a significant change in bacterial diversity (Gerdes and Dieckmann 2006).  Importantly, many of the available slow-release fertilizers are not suitable for use in Arctic regions as they will solidify if used in ice at very low temperatures. For example, the pour point of Inipol EAP 22 is 11°C which makes it difficult to use effectively under Arctic conditions.  As a result slow-release fertilizers will require reformulation or new products will need to be developed specifically for use at very low temperatures.

A few studies have also been conducted to determine the impacts on ice protist communities after oil contamination and subsequent fertilizer treatment.  One of these studies formed part of the ARCOP field trial on Svalbard.  In oil-contaminated ice (no fertilizers) the protist communities were destroyed through complete ice coverage.  Upon addition of fertilizers a less pronounced decline of organisms in the ice interior was observed. Thus, the use of fertilizers (Inipol and fish meal) helped to maintain higher diversity and biomass of protists in the ice. In a separate study, heterotrophic flagellates appeared to escape or avoid the oil contamination by downward migration (Ikävelko et al. 2005).  In an Antarctic field experiment conducted during the Austral winter of 1993, land-fast ice on the continental shelf of Terre Adélie was contaminated with crude oil (Arabian light) or diesel fuel, and negative effects on the ice microalgae were determined by chlorophyll A measurements.  In crude oil-contaminated ice, negative effects were induced which lasted throughout the ice coverage period.  The diesel contamination studies were found to cause an even more rapid effect on the algae than the crude oil. However, the addition of the fertilizer Inipol EAP 22 resulted in clearly favorable effects on the sea ice microalgae (Fiala and Delille 1999).

5.2.2.2 Bioaugmentation

Bioaugmentation has been proposed as a bioremediation method for soil and sediments, often as a supplement to biostimulation treatments. Introduction of exogenous hydrocarbonoclastic bacteria for detoxification of hydrocarbon-polluted cold environments has been reported, with variable success. A number of commercial products exist, which include microbial inocula.  These products are often lacking essential information about the bacterial content.  National authorities may also be skeptical about using these products without proper product information.

In a study of diesel oil-contaminated Alpine soil, a psychrophilic diesel oil-degrading inoculum was added to the contaminated soil, but biostimulation with fertilizers proved more efficient than the bioaugmentation for improved biodegradation activity (Margesin and Schinner 1997).  In microcosm experiments performed in Antarctic gas-oil polluted soil (Jubany Station, King George Island, South Shetland Islands) inoculation of the psychrotolerant strain B-2-2 resulted in 75% hydrocarbon removal, whereas 35% hydrocarbon removal was observed by biostimulation methods when compared to abiotic controls (Ruberto et al. 2003).

Several bioaugmentation studies from marine environments have been reported, although none of these are from cold waters.  A laboratory biodegradation and toxicity study of 12 commercially available bioaugmentation products applied to weathered oil (Alaska North Slope) in seawater at 20°C showed that 3 of the products enhanced biodegradation more than nutrient-amended controls, but only one product resulted in reduced toxicity (Aldrett et al. 1997).  In a marine sediment microcosm study the aromatic-degrading bacterial strain Cycloclasticus sp. E2 was shown to play an important role during degradation of naphthalene in combination with biostimulation treatment (Miyasaka et al. 2006).  Interestingly, bacteria from this genus were also abundant during bioremediation treatment of Arctic oiled beaches at Svalbard (Grossman et al. 2000).

Bioaugmentation has often proved inferior to biostimulation.  One plausible explanation for this may be that the introduced bacteria will have an immediate effect due to the biomass added, but these exogenous microbes may gradually be outcompeted by the indigenous microbes adapted to the local environment.   In a study with small- and mesoscale systems, addition of inorganic nutrients was more efficient at enhancing oil biodegradation in sediments than a commercial product consisting of nutrients and bacterial inocula.  The use of the bioaugmentation product suppressed both the rate and the extent of oil loss by tidal activity and biodegradation, when compared to the periodic addition of inorganic nutrients (Lee et al. 1997).

5.2.2.3 Understanding Processes in Accelerated Biodegradation

Purposeful acceleration of biodegradation through, for example, ecological engineering requires an understanding of the microorganisms that are present in the local environment and their responses to stimuli such as those outlined in the previous sections.  Advances have been made in this area, including development of techniques and submersible vehicles that allow evaluation of biodegradation in situ, improvements in laboratory equipment to simulate the natural environment, maturation of molecular methods to track population dynamics during the degradation process, and development of techniques that allow better understanding of functions carried out by those populations.  These topics were dealt with in more detail in the analytical methods subsection of this section entitle Determining Biodegradation section.  The –omics approaches are of particular importance to accelerated biodegradation.  The effects of management practices on microbial populations and functions, with resulting effects on the components of degraded oil and rates of degradation, can begin to be understood through the emerging fields of metagenomics, metaproteomics, and metabolomics.  These techniques promise to provide much greater understanding of microbial responses, which in turn will allow evaluation and management of oil degradation processes in the Arctic environment. 

5.3 Future Research Considerations

The review of the biodegradation of oil in the Arctic described by the authors in this section led to suggestions of further research which can reduce remaining uncertainties. The more generic suggestions compiled from this review are summarized below while recommendations that are important for improving Arctic NEBA are listed separately. 

  1. Determine effects oil properties may have on biodegradation under Arctic conditions. The physico-chemical characteristics and weathering conditions of different oils at low temperatures may vary considerably, having impacts on biodegradation efficiencies and should be considered for further research, for instance, in the relationship between biodegradation and oil appearance (e.g. viscosity, dispersibility, resurfacing) after a spill. Temperature-related biodegradation data used in models, often based on Q10-approaches, may show erroneous results at very low temperatures, probably due to physical changes in the oil. 
  2. Further document the rates of weathering of stranded oil. Oil residuals from treated or untreated oil undergo different rates of weathering when transferred to sandy beaches or cobble beaches, under or within annual or multi-year ice and within lagoons or near estuaries. Stranded oil has a different potential for release of oil from those compartments and thus longer or shorter periods of time where localized re-oiling of adjacent locations occurs. 
  3. Explore options of enhanced bioremediation with biostimulation and/or bioaugmentation. Biostimulation at very low temperatures where hydrocarbon biodegradation is apparent should be a research issue. Accelerated biodegradation has mainly been investigated by adding fertilizer formulations to improve the oil biodegradation capabilities of natural microbial populations. Biostimulation is regarded as a relevant tool for removal of stranded oil residues as an extension of mechanical treatment. Several formulations of fertilizers have been tested experimentally and under real spill situations. Oleophilic fertilizers have proven more efficient than inorganic nutrients, but available formulations seem to have reduced effects at very low temperatures, such as in marine ice (Gerdes and Dieckmann 2006). Bioaugmentation seems to be less efficient than biostimulation, although some experimental studies have combined the two remediation options. However, several oil cleanup products have been marketed in recent years, including bacteria. Further development of techniques for the enrichment of site-specific oil-degrading communities may therefore be of relevance to combine with other remediation techniques. 

5.3.1 Priority Recommendations for Enhanced NEBA Applications in the Arctic

The recommendations presented below indicate where increased knowledge of biodegradation processes would result in reducing existing uncertainties in NEBA assessments.  No prioritization has been made to the list; for some of the recommendations, surrogate data may be already available.

  1. Investigate potential for recalcitrant effects of oil and consequences of oil spill response strategies (OSRs).    There is a need to better understand what happens to the unrecovered oil once released and distributed to different Environmental Compartments (ECs) in Arctic environments, such as surface slicks, and under ice or onshore stranding.  In order to meet these objectives, we recommend conducting a series of laboratory experiments to determine the persistence, rate and extent of biodegradation in relevant Arctic systems, including surface water, ice and sediment under Arctic marine conditions to further our understanding of the biodegradable and recalcitrant oil dynamics in these compartments.
  2. Examine the influence of ice on biodegradation rates.  Marine ice poses a particular challenge to the indigenous microbes, which require the ability to survive and be metabolically active at sub-zero temperatures and at high salt concentrations.  Microbial metabolism has been demonstrated in marine ice, and populations have been stimulated by oil in the ice.  However, the extent of hydrocarbon biodegradation in Arctic ice requires more attention.
  3. Improve knowledge of degradation rates of dispersed oil and oil in the water column and in other Arctic environmental compartments.  More information is needed on how employment of different oil spill response strategies would affect biodegradation rates and oil movement to different compartments.  For investigations of OSR strategies, we suggest studies be conducted with chemically dispersed oil. These studies include experiments relevant for the following compartments and conditions:
    1. The seawater surface with droplets entering the surface creating a surface film/sheen
    2. The water column, creating physically or chemically generated dispersions and maintaining these in the water column
    3. Subtidal seabed, examining the biodegradation of oil remaining on the seabed after particle integration
    4. Oil in marine ice, with dissolution of water-soluble components to the brine channels in the ice
    5. Shoreline sediments, examining the biodegradation of oil stranded in areas that are ‘primed’ by prior exposure to oil compared with biodegradation in areas that are ‘naïve’ or previously unexposed to oil

5.4 Further Information

Authors Dr. Odd Brakstad (SINTEF), Dr. Donald Stoeckel (Battelle), 

Dr. Charles Greer (NRC Canada) 

References

Aamo OM,M Reed, K Downing

1997

Oil spill Contingency And Response (OSCAR) model system: Sensitivity studies

Proc. Int. Oil Spill Conference, paper #008

Aldrett S, Bonner JS, Mills MA, Autenrieth RL, Stephens FL

1997

Microbial degradation of crude oil in marine environments tested in a flask experiment

Water Res. 31: 2840-2848

Al-Mailem DM, Sorkhoh NA, Al-Awadhi H, Eliyas M, Radwan SS

2010

Biodegradation of crude oil and pure hydrocarbons by extreme halophilic archaea from hypersaline coasts of the Arabian Gulf

Extremophiles 14.321–328

Alonso-Gutierrez J, Teramoto M, Yamazoe A, Harayama S, Figueras A & Novoa B

2011

Alkane-degrading properties of Dietzia sp. H0B, a key player in the Prestige oil spill biodegradation (NW Spain)

J Appl Microbiol 111: 800-810.

Amaral-Zettler L, Artigas LF, Baross J, et al.

2010

A Global Census of Marine Microbes

Life in the World's Oceans,(McIntyre A,) Wiley-Blackwell.

Arrhenius S

1889

Über die Reaktionsgeschwindigkeit bei der Inversion von Rohrzucker durch Säuren.

Z. Phys. Chem. 139:79-103

Atlas RM, Bartha R

1972

Biodegradation of petroleum in seawater at low temperatures

Can. J. Microbiol 18:1851-1855

Atlas RM, Hazen T

2011

Oil biodegradation and bioremediation. A tale of the two worst spills in U.S. history

Environ. Sci. Technol. 45:6709-6715.

Baek KH, Yoon BD, Cho DH, Kim BH, Oh HM & Kim HS

2009

Monitoring bacterial population dynamics using real-time PCR during the bioremediation of crude-oil-contaminated soil

J Microbiol Biotechnol 19: 339-345.

Bælum J, Borglin S, Chakraborty R, et al.

2012

Deep-sea bacteria enriched by oil and dispersant from the Deepwater Horizon spill

Environ Microbiol 14: 2405-2416

Bakermans C

2012

Psychrophiles: Life in the cold

In: Anitori, R.P. (Ed.). Extremophiles. Microbiology and biotechnology. Caister Acadamic Press, Norfolk, UK, pp. 53-785.

Balinski A

2012

Geochemist Laura Lapham studies effect of oil spills on microbes in Gulf of Mexico

Available at http://www.umces.edu/cbl/cbl-geochemist-studies-effect-oil-spills-microbes-gulf-mexico (02/11/2013)

Ballestero H, Magdol Z

2011

Biodegradation of Polycyclic Aromatic Hydrocarbons in Simulated Arctic Sea Ice Brine Channels and Protistan Predation

2011 Oil Spill Conference, Savanna Georgia, May 5-7.

Bano N, Ruffin S, Ransom B, Hollibaugh JT

2004

Phylogenetic composition of Arctic Ocean archaeal assemblages and comparison with Antarctic assemblages

Appl. Environ. Microbiol. 70: 781-789

Barron MG, Podrabsky T, Ogle S, Ricker RW

1999

Are aromatic hydrocarbons the primary determinant of petroleum toxicity to aquatic organisms?

Aquatic Toxicology 46: 253-268

Bartlett DH

2002

Pressure effects on in vivo microbial processes

Biochim Biophys Acta 1595: 367-381

Beazley MJ, Martinez RJ, Rajan S, et al.

2012

Microbial community analysis of a coastal salt marsh affected by the Deepwater Horizon oil spill

PLoS One 7: e41305

Berger F, Morellet N, Corrieu G

1996

Cold shock and cold acclimation proteins in the psychrophilic bacterium Arthrobacter globiformis SI55

J. Bacteriol. 178:2999-3007

Biddle JF, Fitz-Gibbon S, Schuster SC, Brenchley JE & House CH

2008

Metagenomic signatures of the Peru Margin subseafloor biosphere show a genetically distinct environment

Proc Natl Acad Sci U S A 105: 10583-10588

Booth AM, Aitken C, Jones DM, Lewis CA, Rowland SJ

2007

Resistance of toxic alkylcyclohexyltetralins to biodegradation by aerobic bacteria

Organic Geochemistry 38: 540-550

Booth AM, PA Sutton, CA Lewis, AC Lewis, A Scarlett, W Chau, J Widdows, SJ Rowland

2007b

Unresolved complex mixtures of aromatic hydrocarbons: Thousands of overlooked persistent, bioaccumulative, and toxic contaminants in mussels

Environ Sci Technol 41(2):457-464

Børresen MH, Barnes DL, Rike AG

2007

Repeated freeze-thaw cycles and their effects on mineralization of hexadecane and phenanthrene in cold climate soils

Cold Regions Sci.Technol. 49: 215-225

Bouchez, M, Blanchet D, Vandecasteele JP

1995

Degradation of polycyclic aromatic hydrocarbons by pure strains and defined strain associations – inhibition phenomena and catabolism

Appl Microbiol Biotechnol 43:156-164

Bowman JP, McCammon SA, Gibson JA, Robertson L & Nichols PD

2003

Prokaryotic metabolic activity and community structure in Antarctic continental shelf sediments

Appl Environ Microbiol 69: 2448-2462

Bowman JS, Rasmussen S, Blom N, Deming JW, Rysgaard S, Sicheritz-Ponten T

2012

Microbial community structure of Arctic multiyear sea ice and surface seawater by 454 sequencing of the 16S RNA gene

ISME J. 6: 11-20

Braddock JF, Ruth ML, Catterall PH, Walworth JL, McCarthy KA

1997

Enhancement and inhibition of microbial activity in hydrocarbon-contaminated Arctic soils: Implications for nutrient-amended bioremediation

Environ Sci Technol 31:2078-2084

Bragg JR, Prince RC, Harner EJ, Atlas RM

1994

Effectiveness of bioremediation for the Exxon Valdez oil spill

Nature 368:413-418

Brakstad OG, Bonaunet K

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 OG, Bonaunet K, Nordtug T, Johansen Ø

2004

Biotransformation of petroleum hydrocarbons in natural flowing seawater at low temperature

Biodegradation 15:337-346

Brakstad OG, Daling PS, Faksness L-G, Almås IK, Vang S-H, Leirvik F

2011

Surface Dispersion and Depletion of the Macondo MC252 Crude Oil

32nd SETAC meeting in Boston 13-17 November 2011

Brakstad OG, Lodeng AG

2005

Microbial diversity during biodegradation of crude oil in seawater from the North Sea

Microb Ecol 49: 94-103

Brakstad OG, S Ramstad, L-G Faksness, A Booth, P Rønningen, K Zahlsen

2011

Biodegradation and bioremediation of petroleum in cold marine environments

Presentation, 4th International Conference on Polar and Alpine Microbiology; Ljubljiana, Slovenia; SINTEF

Brakstad, O.G. and Faksness. L.-G.

2000

Biodegradation of water-accommodated fractions and dispersed oil in the seawater column

Proceedings for the International Conference on Health, Safety and Environment in Oil and Gas Exploration and Production, Stava

Brandvik PJ

1997

Optimisation of oil spill dispersants on weathered oils

PhD diss., Norwegian University of Science and Technology, Trondheim, Norway

Brandvik PJ, Resby JLM, Leirvik F, Johansen Ø, Brakstad OG

2006

Analysis of oil samples to determine the source of surface oil sheens around the Terra Nova FPSO in July -August 2005. Updated after fullfilment of a Terra Nova cold water biodegradation study

SINTEF Report STF80MK F06200. 69 pp (confidential)

Breezee J, Cady N, Staley JT

2004

Subfreezing growth of the sea ice bacterium “Psychromonas ingrahamii”

Microb.l Ecol. 47:300-304

Brinkmeyer R, Knittel K, Jürgens J, Weyland H, Amann R, Helmke E

2003

Diversity and structure of bacterial communities in Arctic versus Antarctic pack ice

Appl. Environ. Microbiol. 69:6610-6619

Butinar L, Strmole T, Gunde-Cimerman N

2011

Relative Incidence of Ascomycetous Yeasts in Arctic Coastal Environments

Microb/ Ecol/ 61.832–843

Callaghan AV, Davidova AV, Savage-Ashlock K, Parisi VA, Gieg LM, Suflita JM, Kukor JJ, Wawrik B

2010

Diversity of Benzyl-and Alkylsuccinate Synthase Genes in Hydrocarbon-Impacted Environments and Enrichment Cultures

Environ/ Sci/ Technol/ 44. 7287–7294

Camilli R, Reddy CM, Yoerger DR, et al.

2010

Tracking hydrocarbon plume transport and biodegradation at Deepwater Horizon

Science 330: 201-204

Chablain PA, Philippe G, Groboillot A, Truffaut N, Guespin-Michel JF

1997

Isolation of soil psychrotrophic toluene-degrading Pseudomonas strain: influence of temperature on the growth characteristics on different substrates

Res Microbiol 148:153-161

Chang W, Klemm S, Beaulieu C, Hawari J, Whyte L, Ghoshal S

2011

Petroleum hydrocarbon biodegradation under seasonal freeze-thaw soil temperature regimes in contaminated soils from a sub-Arctic site

Environ. Sci. Technol. 45: 1061-1066

Christensen C-H; Larsen TH

1993

Method for determining the age of diesel oil spills in the soil

Groundwater Monit. Rev. 13:142-149

Churchill SA, RA Griffin, LP Jones and PF Churchill

1995

Biodegradation rate enhancement of hydrocarbons by an oleophilic fertilizer and a rhamnolipid biosurfactant

J Environ Qual 24(1):19-28

Coates JD, Woodward J, Allen J, Philp P, Lovley DR

1997

Anaerobic degradation of polycyclic aromatic hydrocarbons and alkanes in petroleum-contaminated marine harbor sediments

Appl Environ Microbiol 63: 3589-3593

Cohen Y

2002

Bioremediation of oil by marine microbial mats

International Microbiology 5. 189–193

Collins RE, Rocap G, Deming JW

2010

Persistence of bacerial and archaeal communities in sea ice through an Arctic winter

Environ. Microbiol. 12: 1828-1841

Comeau AM, Li WKW, Tremblay J-E, Carmack EC, Lovejoy C

2011

Arctic Ocean microbial community structure before and after the 2007 record ice minimum

PLoSONE 6: e27492 (Open access)

Cook LL, Boehm P, Barrick R

2011

Degradation patterns of select PAHs provide evidence of photodegradation in surface MC252 crude oil

Paper presented at SETAC North America 32nd Annual Meeting, Boston, Ma, 13-17 October 2011

Cottrell MT, Kirchman DL

2009

Photoheterotrophic microbes in the Arctic Ocean in summer and winter

Appl. Environ. Microbiol. 75: 4958-4966

Cowen JP, Giovannoni SJ, Kenig F, et al.

2003

Fluids from aging ocean crust that support microbial life

Science 299: 120-123

Cyranoski D

2003

Ocean drilling: digging in

Nature 426: 492-494

Dalluge JJ, Hamamoto T, Horikoshi K, Morita RY, Stetter KO, McCloskey JA

1997

Posttranscriptional modification of tRNA in psychrophilic bacteria

J. Bacteriol. 179: 1918-1923

DeAngelis KM, Allgaier M, Chavarria Y, et al.

2011

Characterization of trapped lignin-degrading microbes in tropical forest soil

PLoS One 6: e19306

Delille D, B Delille, E Pelletier

2002

Effectiveness of crude oil bioremediation in Sub-Antarctic intertidal sediments: the bacterial response

Microbial Ecology 44:118-126

Delille D, Delille B

2000

Field observations on the variability of crude oil impact on indigenous hydrocarbon-degrading bacteria from sub-Antarctic intertidal sediments

Mar. Environ. Res. 49:403-417

Delille, D., Basseres, A., and Dessommes, A

1998

Effectiveness of bioremediation for oil-polluted Antarctic seawater

Polar Biology 19: 237-241

Delille, D., Pelletier, E., Rodriguez-Blanco, A. and Ghiglione, J.-F

2009

Effects of nutrient and temperature on degradation of petroleum hydrocarbons in sub-Antarctic coastal seawater

Polar Biol 32.1521–1528

Deppe U, Richnow HH, Michaelis W, Antranikian G

2005

Degradation of crude oil by an arctic microbial consortium

Extremophiles 9: 461-470

D'Hondt S, Jorgensen BB, Miller DJ, et al.

2004

Distributions of microbial activities in deep subseafloor sediments

Science 306: 2216-2221

Diaz MP, Boyd KG, Grigson SJW, Burgess JG

2002

Biodegradation of crude oil acrosss a wide range of salinities by an extremely halotolerant bacterial consortium MPB-N, immobilized onto polypropylene fibers

Biotechnol. Bioengineer. 79: 145-153

Díez B, Bergman B, Pedrós-Alió C, Antó M, Snoeijs P

2012

High cyanobacterial nifH gene diversity in Arctic seawater and sea ice brine

Environ. Microbiol. Rep. 4. 360–366

Donkin P, Smith EL, Rowland SJ

2003

Toxic effects of unresolved complex mixtures of aromatic hydrocarbons accumulated by mussels, Mytilus edulis, from contaminated field sites

Environ Sci & Techol 37: 4825-4830

Douglas GS, Bence AE, Prince RC, McMillen SJ, Butler EL

1996

Environmental stability of selected petroleum hydrocarbon source and weathering ratios

Environ. Sci. Technol. 30. 2332–2339

Doyle S, Dieser M, Broemsen EE, Christner B

2012

General characteristics of cold-adapted microorganisms

In: R.V. Miller and L.G. Whyte (eds.)Polar Microbiology: Life in a deep freeze, ASM Press, Washington, D.C. pp. 103-125

Drozdowski, A., Nudds, S., Hannah, C.G., Niu, H., Peterson, I.K. and Perrie, W.A.

2011

Review of oil spill trajectory modelling in the presence of ice

Canadian Technical Report of Hydrography and Ocean Sciences 274. 84 pages.

Dutta TK, Harayama S

2000

Fate of crude oil by the combination of photooxidation and biodegradation

Environ. Sci. Technol. 34: 1500-1505

Edwards KJ, Fisher AT & Wheat CG

2012

The deep subsurface biosphere in igneous ocean crust: frontier habitats for microbiological exploration

Front Microbiol 3: 8

Eriksson M, Sodersten E, Yu Z, Dalhammar G, Mohn WW

2003

Degradation of polycyclic aromatic hydrocarbons at low temperature under aerobic and nitrate-reducing conditions in enrichment cultures from northern soils

Appl. Environ. Microbiol. 69:275-284

Farrington JW, Davis AC, Frew NM, Rabin KS

1982

No. 2 fuel oil compounds in Mytilus edulis: Retention and release after an oil spill

Marine Biology 66: 15-26

Fedorak PM, Westlake DWS

1981

Degradation of aromatics and saturates in crude oil by soil enrichments

Water Air Soil Poll. 16: 367-375

Feng L, Wang W, Cheng J, et al.

2007

Genome and proteome of long-chain alkane degrading Geobacillus thermodenitrificans NG80-2 isolated from a deep-subsurface oil reservoir

Proc Natl Acad Sci U S A 104: 5602-5607

Fiala M, Delille D

1999

Annual changes of microalgae biomass in Antarctic sea ice contaminated by crude oil and diesel fuel

Polar Biology 21: 391-396

Forschner SR, Sheffer R, Rowley DC, Smith DC

2009

Microbial diversity in Cenozoic sediments recovered from the Lomonosov Ridge in the Central Arctic basin

Environ Microbiol 11: 630-639

Frenzel M

2008

Persistence, bioremediation and fate of unresolved complex mixtures (UCMs) of hydrocarbons in the environment

PhD diss., University of Exeter, U.K.

Fukunaga N, Sahara T, Takada Y

1999

Bacterial adaptation to low temperature: Implications for cold-inducible genes

J. Plant Res. 112: 263-272

Galand PE, Potvin M, Casamayor EO, Lovejoy C

2010

Hydrography shapes bacterial biogeography of the deep Arctic Ocean

ISME J 4: 564-576

Garrett RM, Pickering IJ, Haith CE, Prince RC

1998

Photooxidation of crude oils

Environ Sci Tech 32: 3719-3723

Garrett RM, Rothenburger SJ, Prince RC

2003

Biodegradation of Fuel Oil Under Laboratory and Arctic Marine Conditions

Spill Sci. Technol. Bull. 8: 297-302

Gerdes B, Dieckmann G

2006

Biological degradation of crude oil in sea ice

Abstract to the report “Growth Project GRD-2000-30112 ARCOP Technology and Environment WP 6 Workshop Activities, 16/02/20

German CR, Ramirez-Llodra E, Baker MC, Tyler PA & ChEss Scientific Steering C

2011

Deep-water chemosynthetic ecosystem research during the census of marine life decade and beyond: a proposed deep-ocean road map

PLoS One 6: e23259

Ghiglione J-F, Galand PE, Pommier T, Pedros-Alio C,Maas EW,Bakker K, Bertilson S, Kirchman DL, Lovejoy C, Yager P, M

2012

Pole-to-pole biogeography of surface and deep marine bacterial communities

PNAS 109: 17633-17638

Gough, M.A., and Rowland, S.J.

1990

Characterisation of unresolved complex mixtures of hydrocarbons in petroleum

Nature 344: 648-650

Gradinger R, Zhang Q

1997

Vertical distribution of bacteria in Arctic sea ice from the Barents and Laptev Seas

Polar Biol. 17: 448-454

Gray ND, Sherry A, Grant RJ, et al.

2011

The quantitative significance of Syntrophaceae and syntrophic partnerships in methanogenic degradation of crude oil alkanes

Environ Microbiol 13: 2957-2975

Greer C

2008

Bioremediation of contaminated sites in the Canadian Arctic: Monitoring performance and the effects of biostimulation using molecular methods

In: Bej, A.K., Aislabie, J., Atlas, R.M. (Eds.). Polar Microbiology. The ecology, biodiversity and bioremediation potential of microorganisms

Grossman M, Prince R, Garrett R, Garrett K, Bare R, Lee K, Sergy G, Owens E, Guenette C

2000

Microbial diversity in oiled and un-oiled shoreline sediments in the Norwegian Arctic

In: Bell CR, Brylinsky M, Johnson-Green P (eds) Microbial biosystems: new frontiers. Proceedings of the 8th International Symposium

Groudieva T, Kambourova M, Yosuf H, Royter M, Grote R, Trinks H, Antranikian G

2004

Diversity and cold-active hydrolytic enzymes of culturable bacteria associated with Arctic sea ice, Spitzbergen

Extremophiles 8:475-488

Guenette CC, Sergy GA, Owens EH, Prince RC, Lee K

2003

Experimental design of the Svalbard shoreline field trial

Spill Sci Technol Bull 8:245-256

Gunde-Cimerman G, Sonja S, Zalar P, Frisvad JC, Diderichsen B, Plemenita A

2003

Extremophilic fungi in Arctic ice: a relationship between adaptation to low temperature and water activity

Phys. Chem. Earth 28. 1273–1278

Han XM, Scott AC, Fedorak PM, Bataineh M, Martin JW

2008

Influence of molecular structure on the biodegradability of naphthenic acids

Environ. Sc.Technol. 42: 1290-1295

Hara A, Syutsubo K, Harayama S

2003

Alcanivorax which prevails in oil-contaminated seawater exhibits broad substrate specificity for alkane degradation

Environ Microbiol 5: 746-753

Harayama S, Kishira H, Kasai Y & Shutsubo K

1999

Petroleum biodegradation in marine environments

J Mol Microbiol Biotechnol 1: 63-70.

Hazen TC, Dubinsky EA, DeSantis TZ, et al.

2010

Deep-sea oil plume enriches indigenous oil-degrading bacteria

Science 330: 204-208

Head IM, Jones DM, Larter SR

2003

Biological activity in the deep subsurface and the origin of heavy oil

Nature 426: 344-352

Heider J

2007

Adding handles to unhandy substrates: anaerobic hydrocarbon activation mechanisms

Curr. Opin. Chem. Biol 11.188–194

Helmke, E. and Weyland, H

2004

Psychrophilic versus psychrotolerant bacteria. Occurrence and significance in polar and temperate marine habitats

Cell. Mol. Biol. 50: 553-561

Hokstad JN, Daling PS, Buffagni M, Johnsen S

1999

Chemical and ecotoxicological characterization of oil-water systems

Spill Sc.Technol. Bull. 5:75-80

Horowitz A, Atlas RM

1977

Continuous flow-through system as a model for oil degradation in the Arctic Ocean

Appl. Environ. Microbiol. 33: 647-653

Hubert C, Arnosti C, Bruchert V, Loy A, Vandieken V,Jorgensen BB

2010

Thermophilic anaerobes in Arctic marine sediments induced to mineralize complex organic matter at high temperature

Environ Microbiol 12: 1089-1104

Hubert C, Loy A, Nickel M, Arnosti C, Baranyi C, Bruechert V, Ferdelman T, Finster K, Christensen FM, de Rezende JR, Va

2009

A constant flux of diverse thermophilic bacteria into the cold Arctic seabed

Science 325: 1541-1544

Hughey CA, Minardi CS, Galasso-Roth SA, Paspalof GB, Mapolelo MM, Rodgers RP, Marshall AG, Ruderman DL

2008

Naphthenic acids as indicators of crude oil biodegradation in soil, based on semi-quantitative electrospray ionization Fourier transform ion cyclotron resonance mass spectrometry

Rapid Comm. Mass Spectrometry 22: 3968-3976

Ikävelko J, Gerdes B, Dieckmann G

2005

An Experimental Study on the Effects of Statfjord Crude Oil, and Application of Inipol and Fish Meal on the Sea Ice Biota in Svalbard in February - April 2004

Proceedings, 28th Arctic and Marine Oilspill Program Technical Seminar, Calgary, Canada, pp. 993-1003

Inagaki F, Nunoura T, Nakagawa S, et al.

2006

Biogeographical distribution and diversity of microbes in methane hydrate-bearing deep marine sediments on the Pacific Ocean Margin

Proc Natl Acad Sci U S A 103: 2815-2820

Jobson A, Cook FD, Westlake DW

1972

Microbial utilization of crude oil

Appl Microbiol 23: 1082-1089

Jones DM, Head IM, Gray ND, et al.

2008

Crude-oil biodegradation via methanogenesis in subsurface petroleum reservoirs

Nature 451: 176-180

Jung JY, Choi AR, Lee YK, Lee HK, Jung K-H

2008

Spectroscopic and photochemical analysis of proteorhodopsin variants from the surface of the Arctic Ocean

Spectroscopic and photochemical analysis of proteorhodopsin variants from the surface of the Arctic Ocean

Junge K, Eicken H, Deming JW

2003

Motility of Colwellia psychrerythraea strain 34H at subzero temperatures

Applied and Environmental Microbiology 69: 4282-4284

Junge K, Eicken H, Deming JW

2004

Bacterial activity at -2 and -20°C in Arctic wintertime sea ice

Applied and Environmental Microbiology 70: 550-557

Junge K, Eicken H, Swanson BD, Deming JW

2006

Bacterial incorporation of leucine into protein down to -20°C with evidence for potential activity in sub-eutectic saline formations

Cryobiology 52: 417-429

Junge K, Imhoff F, Staley T, Deming JW

2002

Phylogenetic diversity of numerically important Arctic sea-ice bacteria cultured at subzero temperature

Microbial Ecology 43:315-328

Kerkhof LJ, Goodman RM

2009

Ocean microbial metagenomics

Deep Sea Research Part II: Topical Studies in Oceanography 56: 1824-1829

essler JD, Valentine DL, Redmond MC, Du M, Chan EW, Mendes SD, Quiroz CJ, Villanueva CV, Shusta SS, Werra LM, Yv

2011

Persistant Oxygen Anomaly Reveals the Fate of Spilled Methane in the Deep Gulf of Mexico

Science 331: 312-315

Khachane AN, Timmis KN, dos Santos VAPM

2005

Uracil content of 16S rRNA of thermophilic and psychrophilic prokaryotes correlates inversely with their optimal growth temperatures

Nucl. Acids Res. 33: 4016-4022

Killops SD, Al-Juboori MAHA

1990

Characterisations of unresolved complex mixtures in the gas chromatograms of biodegraded petroleums

Organic Geochemistry 15: 147-160

Kim SJ, Kweon O, Jones RC, Edmondson RD & Cerniglia CE

2008

Genomic analysis of polycyclic aromatic hydrocarbon degradation in Mycobacterium vanbaalenii PYR-1

Biodegradation 19: 859-881

Kleikemper J, Schroth MH, Sigler WV, Schmucki M, Bernasconi SM & Zeyer J

2002

Activity and diversity of sulfate-reducing bacteria in a petroleum hydrocarbon-contaminated aquifer

Appl Environ Microbiol 68: 1516-1523

Knezevich V, O Koren O, EZ Ron, E Rosenberg

2007

Petroleum bioremediation in seawater using guano as the fertilizer

Bioremediation J. 10: 83-91

Knoblauch C, Jørgensen BB,Harder J

1999

Community size and metabolic rates of psychrophilic sulfate-reducing bacteria in Arctic sediments

Appl. Environ. Microbiol. 65: 4230-4233

Kobayashi T, Koide O, Mori K, et al.

2008

Phylogenetic and enzymatic diversity of deep subseafloor aerobic microorganisms in organics- and methane-rich sediments off Shimokita Peninsula

Extremophiles 12: 519-527

Kostka JE, Prakash O, Overholt WA, et al.

2011

Hydrocarbon-degrading bacteria and the bacterial community response in gulf of Mexico beach sands impacted by the deepwater horizon oil spill

Appl Environ Microbiol 77: 7962-7974

Kujawinski EB, Kido Soule MC, Valentine DL, Boysen AK, Longnecker K, Redmond MC

2011

Fate of Dispersants Associated with the Deepwater Horizon Oil Spill

Environ Sci Technol. 45. 1298–1306

Kweon O, Kim SJ, Holland RD, et al.

2011

Polycyclic aromatic hydrocarbon metabolic network in Mycobacterium vanbaalenii PYR-1

J Bacteriol 193: 4326-4337

Lake RA, Lewis EL

1970

Salt rejection by sea ice during growth

J. Geophys. Res. 75. 583–597

Lamberts R, Christensen JH, Mayer P, Andersen O, Johnsen AR

2008

Isomer-specific biodegradation of methylphenanthrenes by soil bacteria

Environ. Sci. Technol. 42: 4790-4796

Lee K, Merlin FX

1999

Bioremediation of oil on shoreline environments: development of techniques and guidelines

Pure Appl. Chem. 71. 161–171

Lindstrom JE, Braddock JF

2002

Biodegradation of petroleum hydrocarbons at low temperature in the presence of the dispersant Corexit 9500

Mar Pollut Bull 44: 739-747

Liu R, Zhang Y, Ding R, Li D, GaoY, Yang M

2009

Comparison of archaeal and bacterial community structures in heavily oil-contaminated and pristine soils

J. Biosci. Bioeng. 108: 400–407

Lomstein BA, Langerhuus AT, D'Hondt S, Jorgensen BB, Spivack AJ

2012

Endospore abundance, microbial growth and necromass turnover in deep sub-seafloor sediment

Nature 484: 101-104

Lu Z, Deng Y, Van Nostrand JD, et al.

2012

Microbial gene functions enriched in the Deepwater Horizon deep-sea oil plume

ISME J 6: 451-460

Lynch JM, Moffat AJ

2005

Bioremediation - prospects for the future application of innovative applied biological research

Annals Appl Biol 146:217-221

MacNaughton SJ, Stephen JR, Venosa AD, Davis GA, Chang YJ & White DC

1999

Microbial population changes during bioremediation of an experimental oil spill

Appl Environ Microbiol 65: 3566-3574

MacNaughton SJ, Swannell R, Daniel F, Bristow L

2003

Biodegradation of dispersed Forties crude and Alaskan North Slope oils in microcosms under stimulated marine conditions

Spill Sci. Technol. 8:179-186

Maier RM, G Soberon-Chavez

2000

Pseudomonas aeruginosa rhamnolipids: biosynthesis and potential applications

Appl. Microbiol. Biotechnol. 54. 625–633.

Maki H, Sasaki T, Harayama S

2001

Photo-oxidation of biodegraded crude oil and toxicity of the photo-oxidized products

Chemosphere 44: 1145-1151

Margesin R, Gander S, Zacke G, Gounot AM & Schinner F

2003

Hydrocarbon degradation and enzyme activities of cold-adapted bacteria and yeasts

Extremophiles 7: 451-458

Margesin R, Schinner F

1997

Efficiency of indigenous and inoculated cold-adapted soil microorganisms for biodegradation of diesel oil in Alpine soils

Applied and Environmental Microbiology 63:2660-2664

Marshall AG, Rogers RP

2003

Petroleomics: the next grand challenge for chemical analysis

!cc hem Res 37. 53–59

Mason OU, Hazen TC, Borglin S, et al.

2012

Metagenome, metatranscriptome and single-cell sequencing reveal microbial response to Deepwater Horizon oil spill

ISME J 6: 1715-1727

Meredith W, Kelland SJ, Jones DM

2000

Influence of biodegradation on crude oil acidity and carboxylic acid composition

Organic Geochem. 31: 1059-1073

Michaud L, Lo Giudice A, Saitta M, De Domenico M, Bruni V

2004

The biodegradation efficiency on diesel oil by two psychrotrophic Antarctic marine bacteria during a two-month-long experiment

Mar Pollut Bull 49: 405-409

Minard-Smith AB, Arthur M, Tilstone A, Metzger B, Faith S

2012

Use of "Omic" techlologies to study microbial response to North Slope crude oil. Anchorage, AK.

Presentation

Minas W, Gunkel W

1995

Oil pollution in the North Sea – a microbiological point of view

Helgoländer Meeresunters 49:143-158

Miyasaka T, Asami H, Watanabe K

2006

Impacts of bioremediation schemes on bacterial population in naphthalene-contaminated marine sediments

Biodegradation 17:227-235

Moldestad MØ, Leirvik F, Melbye AG, Ditlevsen MK, Wang UM

2003

Goliat – weathering properties, appearance code, water solubility and toxicity

SINTEF Report F03104, 136 p.

Molina DV, UN Uribe, J Murgich

2007

Partial Least-Squares (PLS) Correlation between Refined Product Yields and Physicochemical Properties with the 1H Nuclear Magnetic Resonance (NMR) Spectra of Colombian Crude Oils

Energy and Fuels 21:1674-1680

Mulligan CN

2005

Environmental applications for biosurfactants (Review)

Env. Poll. 133: 183-198.

Mykytzuk NCS, Foote SJ, Omelon CR, Southam G, Greer CW, LG Whyte

2013

Bacterial growth at -15°C; molecular insights from the permafrost bacterium Planococcus halocryophilus Or1

ISME J (in press)

Nichols CM, Bowman JP, Guezennec J

2005

Effects of incubation temperature on growth and production of exopolysaccharides by an Antarctic sea ice bacterium grown in batch culture

Appl. Environ. Microbiol. 71:3519-3523

Ni'matuzahroh M, Gilewicz M, Guiliano M, Bertrand JC

1999

In-vitro study of interaction between photooxidation and biodegradation of 2-methylphenanthrene by Sphingomonas sp. 2MPII

Chemosphere 38: 2501-2507

Obbard JP, Ng KL, Xu R

2004

Bioremediation of petroleum contaminated beach sediments: Use of crude palm oil and fatty acids to enhance indigenous biodegradation

Water Air and Soil Pollution 157:149-161

Olivieri R, P Bacchin, A Robertielli N Oddo, L Degen, Atonolo

1975

Microbial degradation of oil spills by a slow-release fertilizer

App. Environ. Microbiol. 31: 629-634

Page, D.S., Boehm, P.D., Douglas, G.S., Bence, A.E

1996

Identification of hydrocarbons in the benthic sediments of Prince Williams Sound and the Gulf of Alaska following the Exxon Valdez oil spill

In: Exxon Valdez oil spill: Fate and effects in Alaskan waters; Wells, P.G., Butler, J.N., Hughes, J.S., Eds.; American Society for Testing

Pelletier E, Delille D, Delille B.

2004

Crude oil bioremediation in sub-Antarctic intertidal sediments: chemistry and toxicity of oiled residue

Marine Environmental Research 57:311-327

Perry JJ

1984

Microbial metabolism of cyclic alkanes

In: Petroleum Microbiology, ed. R.M. Atlas, 61-98. Macmillan Publ. Co., New York

Powell SM, I Snape, JP Bowman, BAW Thompson et al.

2005

A comparison of the short term effects of diesel fuel and lubricant oils on Anarctic benthic microbial community

J Exp Mar Biol Ecol 322:53-65

Prince RC

2010

Can we improve bioremediation ?

In: Timmis, K.N. (Ed.). Handbook of Hydrocarbon and Lipid Microbiology. Springer-Verlag Berlin Heidelberg, pp. 3351-3355

Prince RC

1993

Petroleum spill bioremediation in marine environments

Critical Reviews in Microbiology 19:217-242

Prince RC

2005

The microbiology of marine oil spill bioremediation

In: Petroleum Microbiology, ed. B. Ollivier, and M. Magot, 317-335. ASM Press, Washington DC.

Prince RC and JD Butler

2013

A protocol for assessing the effectiveness of oil spill dispersants in stimulating the biodegradation of oil

Environ Sci Pollut Res DOI 10.1007-s11356-013-2053-7

Prince RC, Atlas RM

2005

Bioremediation of marine oil spills (3)

In: Bioremediation: applied microbial solutions for real-world environmental cleanup, ed. R.M., and J. Philp, 269-292. ASM Press,

Prince RC, Clark JR

2004

Bioremediation of marine oil spills (2)

In: Petroleum biotechnology: Prince, R.C., Bare, R.E., Garrett, R.M., Grossman, M.J., Haith, C.E., Keim, L.G., Lee, K., Holtom, G.J., La

Prince RC, Elmendorf DL, Lute JR, Hsu CS, Haith CE, Senius JD, Dechert GJ, Douglas GS, Butler EL

1994

17a(H),21b(H)-hopane as a conserved internal marker for estimating the biodegradation of crude oil

Environ. Sci. Technol. 28: 142–145

Prince RC, Garrett RM, Bare RE, Grossman MJ, Townsend T, Suflita JM, Lee K, Owens EH, Sergy GA, Braddock JF, Lindstr

2003b

The roles of photooxidation and biodegradation in long-term weathering of crude and heavy fuel oils

Spill Sci. Technol. Bull. 8: 145-156

Prince RC, RE Bare, RM Garrett, MJ Grossman, et al.

2003a

Bioremediation of stranded oil on an Arctic shoreline

Spill Sci & Technol Bull 8(3):308-312

Prince, R.C., McFarlin, K.M., Butler, J.D., Febbo, E.J., Wang, F.C.Y. and Nedwed, T.J.

2012

The primary biodegradation of dispersed crude oil in the sea

Chemosphere 90: 521-526

Pritchard PH, Mueller JG, Rogers JC, Kremer FV, Glaser JA

1992

Oil spill bioremediation: experiences, lessons amd results from the Exxon Valdez oil spill in Alaska

Biodegradation 3: 315-335

Rabus R, Wilkes H, Schramm A, Harms G, Behrends A, Amann R, Widdel F

1999

Anaerobic utilization of alkylbenzenes and n-alkanes from crude oil in an enrichment culture of denitrifying bacteria affiliating with the beta-subclass of Proteobacteria

Environ Microbiol 1: 145-157

Raloff J

2010

Bug traps in Gulf to use BP oil as bait

Society for Science and the Public; Vol. August 25 ed.^eds.), p.^pp

Ravenschlag K, Sahm K, Ama R

2001

Quantitative molecular analyses of the microbial communities in marine Arctic sediments (Svalbard)

Appl. Environ. Microbiol. 67: 387-395

Reddy CM, Arey JS, Seewald JS, et al.

2012

Composition and fate of gas and oil released to the water column during the Deepwater Horizon oil spill

Proc Natl Acad Sci U S A 109: 20229-20234

Reed M, Daling PS, Brakstad OG, Singsaas I, Faksness L-G, Hetland B, Ekrol N

2000

OSCAR 2000: A multi-component 3-dimensional oil spill contingency and response model

In: Proceedings to the 2000 Arctic and Marine Oilspill Program Technical Seminar, Vancouver, Canada, June 14-16, 2000. Environment

Rike, A.G., Haugen, K.B., Børresen, M., Engene, B., and Kolstad, P

2003

In situ biodegradation of petroleum hydrocarbons in frozen Arctic soils

Cold Regions Sci.Technol. 37: 97-120

Robador, A., Bruchert, V. and Jorgensen, B.B

2009

The impact of temperature change on the activity and community composition of sulfate-reducing bacteria in arctic versus temperate marine sediments

Environ. Microbiol. 11: 1692-1703

Rojo F

2009

Degradation of alkanes by bacteria

Environ Microbiol 11: 2477-2490

Röling WFM, de Brito IRC, Swannell RPJ, Head IM

2004

Response of archaeal communities in beach sediments to spilled oil and bioremediation

Appl. Environ. Microbiol. 70: 2614-2620

Rosnes JT, Torsvik T, Lien T

1991

Spore-Forming Thermophilic Sulfate-Reducing Bacteria Isolated from North Sea Oil Field Waters

Applied and Environmental Microbiology 57: 2302-2307

Rowland S, Donkin P, Smith E, Wraige E

2001

Aromatic Hydrocarbon "Humps" in the Marine Environment: Unrecognized Toxins?

Environmental Science & Technology 35: 2640 -2644

Ruberto L, Vazquez SC, Mac Cormack WP

2003

Effectiveness of the natural bacterial flora, biostimulation and bioaugmentation on the bioremediation of a hydrocarbon contaminated Antarctic soil

International Biodeterioration and Biodegradation 52:115-125

Russel, N

2008

Membrane components and cold sensing

In: Margesin, R., Schinner, F., Marx, J.-C. and Gerday, C. (Eds.). Psychrophiles. From biodiversity to biotechnology. Springer Verlag,

Rysgaard S, Glud RN

2004

Anaerobic N2 production in Arctic sea ice

Limnol. Oceanogr., 49: 86–94

Schippers A, Neretin LN, Kallmeyer J, Ferdelman TG, Cragg BA, Parkes RJ & Jorgensen BB

2005

Prokaryotic cells of the deep sub-seafloor biosphere identified as living bacteria

Nature 433: 861-864

Schluep M, DM Imboden, R Galil, J Zeyer

2001

Mechanisms affecting the dissolution of nonaquaueous phase liquids into the aqueous phase in slow stirring batch systems.

Environ Toxicol Chem 20:459-466

Scott AC, MD MacKinnon, PM Fedorak

2005

Naphthenic acids in Athabasca oil sands tailings waters are less biodegradable than commercial naphthenic acids

Environ Sci Technol 39:8388-8394

Sei K, Sugimoto Y, Mori K, Maki H, Kohno T

2003

Monitoring of alkane-degrading bacteria in a sea-water microcosm during crude oil degradation by polymerase chain reaction based on alkane-catabolic genes

Environ Microbiol 5: 517-522

Simonato F, Campanaro S, Lauro FM, et al.

2006

Piezophilic adaptation: a genomic point of view

J Biotechnol 126: 11-25

Siron R, Pelletier É, Brochu C

1995

Environmental factors influencing the biodegradation of petroleum hydrocarbons in cold seawater

Archives of Environmental Contamination and Toxicology 28: 406-416

Smith E, Wraige E, Donkin P, Rowland S

2001

Hydrocarbon humps in the marine environment: Synthesis, toxicity, and aqueous solubility of monoaromatic compounds

Environmental Toxicology and Chemistry 20: 2428-2432

Smith VA, Graham DW, Cleland DD

1998

Application of resource-ratio theory to hydrocarbon biodegradation

Environ Sci Technol 32:3386-3395

Sogin ML, Morrison HG, Huber JA, et al.

2006

Microbial diversity in the deep sea and the underexplored "rare biosphere"

Proc Natl Acad Sci U S A 103: 12115-12120

Stevens FL

2004

Thin layer chromatography – flame ionization detection analysis of in-situ petroleum biodegradation

Master of Science Thesis at the Texas A&M University

Sul WJ, Oliver TA, Ducklow HW, Amaral-Zettler LA & Sogin ML

2013

Marine bacteria exhibit a bipolar distribution

Proc Natl Acad Sci U S A

Sveum P, Ladousse A

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 DC, pp 439-445

Swannell RPJ, Lee K, McDonagh M

1996

Field evaluation of marine spills bioremediation

Microbiological Reviews 60:242-365

Syldatk C, F Wagner

1987

Production of biosurfactants

In: Kosaric, N., Cairns, W.L., Gray, N.C. (Eds.), Biosurfactants and Biotechnology. In: Surfactant Science Series, vol. 25. Marcel Dekk

Tang, S., Qin, D., Ren, J.,Kang, J. and Li, Z

2007

Structure, salinity and isotopic composition of multi-year landfast sea ice in Nella Fjord, Antarctica"

Cold Regions Science and Technology 49. 170–177

Tapilatu Y, Acquaviva M, Guigue C, Miralles G, Bertrand JC, Cuny P

2010

Isolation of alkane-degrading bacteria from deep-sea Mediterranean sediments

Lett Appl Microbiol 50: 234-236

Teramoto M, Ohuchi M, Hatmanti A, Darmayati Y, Widyastuti Y, Harayama S & Fukunaga Y

2011

Oleibacter marinus gen. nov., sp. nov., a bacterium that degrades petroleum aliphatic hydrocarbons in a tropical marine environment

Int J Syst Evol Microbiol 61: 375-380

Teramoto M, Suzuki M, Okazaki F, Hatmanti A, Harayama S

2009

Oceanobacter-related bacteria are important for the degradation of petroleum aliphatic hydrocarbons in the tropical marine environment

Microbiology 155: 3362-3370

Terrado, R., Lovejoy, C., Massana, R., Vincent, W.F.

2008

Microbial food web responses to light and nutrients beneath the coastal !rctic Ocean sea ice during the winter–spring transition

J. Mar. Syst. 74: 964-977

Teske A, Durbin A, Ziervogel K, Cox C, Arnosti C

2011

Microbial community composition and function in permanently cold seawater from an Arctic fjord of Svalbard

Appl. Environ. Microbiol. 77: 2008-2018

Throne-Holst M, Markussen S, Winnberg A, Ellingsen TE, Kotlar HK, Zotchev SB

2006

Utilization of n-alkanes by a newly isolated strain of Acinetobacter venetianus: the role of two AlkB-type alkane hydroxylases

Appl Microbiol Biotechnol 72: 353-360

Throne-Holst M, Wentzel A, Ellingsen TE, Kotlar HK & Zotchev SB

2007

Identification of novel genes involved in long-chain n-alkane degradation by Acinetobacter sp. strain DSM 17874

Appl Environ Microbiol 73: 3327-3332

Tran TC, Logan G, Grosjean E, Ryan D, Marriott P

2010

Use of comprehensive two-dimensional gas chromatography/time-of-flight mass spectrometry for the characterization of biodegradation and unresolved complex mixtures in petroleum

Geochim. Cosmochim. Acta 74: 6468-6484

Valentine DL, Kessler JD, Redmond MC, et al.

2010

Propane respiration jump-starts microbial response to a deep oil spill

Science 330: 208-211

Valentine DL, Mezic I, Macesic S, et al.

2012

Dynamic autoinoculation and the microbial ecology of a deep water hydrocarbon irruption

Proc Natl Acad Sci U S A 109: 20286-20291

van Beilen JB, Marin MM, Smits TH, Rothlisberger M, Franchini AG, Witholt B & Rojo F

2004

Characterization of two alkane hydroxylase genes from the marine hydrocarbonoclastic bacterium Alcanivorax borkumensis

Environ Microbiol 6: 264-273

Varadaraj R, Robbins ML, Bock J, Pace S, MacDonald D

1995

Dispersion and biodegradation of oils spills on water

In: Proceedings to the 1995 International Oil Spill Conference, American Petroleum Institute, Washington DC, pp 101-106

Venosa AD, Holder EL

2007

Biodegradability of dispersed crude oil at two different temperatures

Mar. Poll. Bull. 54: 545-553

Venter JC, Remington K, Heidelberg JF, et al.

2004

Environmental genome shotgun sequencing of the Sargasso Sea

Science 304: 66-74

Vilchez-Vargas R, Junca H & Pieper DH

2010

Metabolic networks, microbial ecology and 'omics' technologies: towards understanding in situ biodegradation processes

Environ Microbiol 12: 3089-3104

Wang Z, Fingas M, Blenkinsopp S, Sergy G, Landriault M, Sigouin L, Lambert P

1998

Study of 25-year old Nipisi oil spill: Persistence of oil residues and comparisons between surface and subsurface sediments

Environ. Sci. Technol. 32:2222-2232

Wasmund K, Burns KA, Kurtboke DI, Bourne DG

2009

Novel alkane hydroxylase gene (alkB) diversity in sediments associated with hydrocarbon seeps in the Timor Sea, Australia

Appl Environ Microbiol 75: 7391-7398

Wentzel A, Ellingsen TE, Kotlar HK, Zotchev SB & Throne-Holst M

2007

Bacterial metabolism of long-chain n-alkanes

Appl Microbiol Biotechnol 76: 1209-1221

Whyte LG, Bourbonniere L & Greer CW

1997

Biodegradation of petroleum hydrocarbons by psychrotrophic Pseudomonas strains possessing both alkane (alk) and naphthalene (nah) catabolic pathways

Appl Environ Microbiol 63: 3719-3723

Whyte LG, C Bourbonniere, C Bellrose, WC Greer

1999

Bioremediation assessment of hydrocarbon contaminated soils from the high Arctic

Bioremediation Journal 3:69-79

Widdows J, Donlkin P, Brinsley MD, Evans SV, Salkeld PN, Franklin A, Law RJ, Waldock MJ

1995

Scope for growth and contaminant levels in North-Sea mussels Mytilus edulis

Marine Ecology Progress Series 127: 131-148

Woese C, Kandler O, Wheelis M

1990

Towards a natural system of organisms: proposal for the domains Archaea, Bacteria, and Eucarya

PNAS 87: 4576–4579

Yergeau E, Hogues H, Whyte LG & Greer CW

2010

The functional potential of high Arctic permafrost revealed by metagenomic sequencing, qPCR and microarray analyses

ISME J 4: 1206-1214

Yergeau E, Sanschagrin S, Beaumier D, Greer CW

2012

Metagenomic analysis of the bioremediation of diesel-contaminated Canadian high Arctic soils

PLoS One 7: e30058

Zinger L, Amaral-Zettler LA, Fuhrman JA, et al.

2011

Global patterns of bacterial beta-diversity in seafloor and seawater ecosystems

PLoS One 6: e24570

Hylland K, Tollefsen KE, Ruus A, Jonsson G, Sundt RC, Sanni S, Utvik TIR, Johnsen S, Nilssen I, Pinturier L, Balk L, Barsien

2008

Water column monitoring near oil installations in the North Sea 2001-2004

Mar Pollution Bull 56(3):414-429.

Melbye AG, Brakstad OG, Hokstad JN, Gregersen IK, Hansen BH, Booth AM, Rowland SJ, Tollefsen KE

2009

Chemical and toxicological characterisation of an unresolved complex mixture (UCM)-rich biodegraded crude oil

Environ. Toxicol. Chem. 28: 1815-1824

6.0 ECOTOXICOLOGY OF OIL AND TREATED OIL IN THE ARCTIC

Executive Summary

Photo 6-1: Boreogadus saida (Lionel Camus)
Photo 6-1: Boreogadus saida (Lionel Camus)

This section presents a review of the current knowledge on the effects of oil and treated oil on aquatic Arctic organisms, focusing on available ecotoxicology data from peer-reviewed literature (includes 85 published studies and 14 papers in review). These studies were conducted with a variety of media, test organisms, and experimental designs: 

  • Crude oil was the main test media 
  • Test organisms represented 44 species from 11 taxonomic groups. Where possible, the test species represent vital linkages in the Arctic trophic web (valuable ecosystem components or VECs) 
  • The various experimental designs included spiked and continuous exposures of fresh and/or weathered oil in the presence or absence of dispersant; different preparations of exposure media [water soluble fraction (WSF), water accommodated fraction (WAF and chemically enhanced WAF) and a newly introduced method of oil-in-water dispersions (OWD)]; and reporting of nominal or quantitatively measured exposure concentrations. Results were typically reported in terms of total petroleum hydrocarbons (TPH) or as toxicant-specific compounds (such as naphthalene and its congeners). 

The majority of the toxicity experiments were conducted in laboratories, very few ecotoxicity studies have been conducted with Arctic field experiments. These disparate experimental designs and particularly methods used in quantifying data can lead to variable interpretations of relative toxicity and species sensitivity estimations brought forward to ecological risk assessments and net environmental benefit analyses. Because of the relative difficulty in conducting Arctic toxicology studies at extremely low temperatures with authentic Arctic species, there are relatively few comprehensive investigations. However, relatively recent attention has focused on the issue of relative sensitivity of Arctic species to temperate species and several assessments have similarly concluded that arctic and temperate species show little difference in relative sensitivity when toxicity studies were conducted with similar methodologies. 

6.1 Introduction

An extended ice-free season has expanded opportunities for oil and gas exploration in the Arctic and has increased the need to better understand the environmental consequences of oil spill response alternatives under Arctic conditions. Oil spill counter measures include physical removal, the application of chemical or physical dispersants, and in-situ burning, as well as emerging technologies such as chemical herders and oil-mineral aggregates (OMAs).  Oil spill responders need to quickly evaluate the best treatment options for the event at hand that will cause the least amount of environmental perturbations.  This requires a multi-tiered assessment of interactive attributes and processes, such as:

  • What are the chemical components of the oil spilled? How would the toxic fraction be characterized?
  • What volume of oil was released and what is the probable fate of the oil if untreated? – If treated with chemical dispersants or oil/mineral aggregates?
  • What VECs might be impacted by short- and long-term releases?

Each of these different alternatives may alter the fate, bioavailability and potential environmental impacts of a release.  Numerical models are available to describe the fate and spreading of the oil (e.g.DREAM, OSCAR) or to predict the population levels effects of a release (e.g. SYMBIOSES).  However, in order to link the composition and fate of released petroleum to predictions of the potential population level impacts, toxicological data is required (see Figure 6-1).   The oil and gas industry is compiling and centralizing an adequate ecotoxicity database to be used as an integral part of forecasting probable effects caused by an oil spill in order to best mitigate the environmental damages, both short term and long term.  Knowledge about the relative toxicity consequences of each scenario is necessary to ensure that the most effective spill response option is chosen.  Often the most effective treatment may be a combination of spill response options. 

This section reviews the current knowledge on the effects of oil and treated oil on aquatic Arctic organisms, focusing on available ecotoxicology data available in the peer-reviewed literature. Approximately 85 peer-reviewed publications and 14 papers in preparation have been reviewed. These publications evaluate the sensitivities of 44 species from 11 different taxonomic groups. Data on 14 algae species and six bacteria species could be identified in the literature. Additionally, four papers address effects on communities.  Nine papers reported toxicity studies conducted at relatively high temperature (i.e. 10oC) with boreal species that occur in the Arctic for a portion of their life history. Eight papers have been identified reporting baseline levels of biomarkers and/or petroleum related contaminant levels. The majority of the studies have been conducted with crude oil or single polycyclic aromatic compounds. One paper described a field trial of releasing oil in the field in order to study the short and long term impact with or without any response. There is a large variety in exposure methods, oil type and treatments, complicating the interpretation and application of these data in risk assessment procedures. Different exposure systems and exposure quantifications may lead to different interpretations on the relative toxicity of oil and treated oil. 

Figure 6-1. Toxicity empirical data (outlined in red) are used in advanced predictive models <em>(e.g. PETROTOX, EPISUITE, QSAR, OSCAR, TLM)</em>
Figure 6-1. Toxicity empirical data (outlined in red) are used in advanced predictive models (e.g. PETROTOX, EPISUITE, QSAR, OSCAR, TLM)

There is a substantial body of data regarding the toxicity of oil to aquatic organisms.  A majority of the existing toxicity data is for acute lethality of oil-water mixtures to temperate species (DeHoop et al. 2011; Word and Gardiner In prep).  More recent research efforts have begun to focus on Arctic test organisms and have included sublethal endpoints to acute and chronic exposures, as well as acute lethality.  The following paper presents a summary of existing toxicological studies on the effects of oil and treated oil on pelagic, ice, and benthic community species.  This review includes toxicological studies on Arctic and boreal species that are available in the scientific literature, including peer-reviewed journal articles, published books, and meetings proceedings.  Scientific papers that have been accepted for publication were also included.  Papers that have been submitted but have not yet been accepted are referred to here, but were not included in the summary data tables.  Reports published in the peer-reviewed literature and in books offer the advantage of being openly available as well as representing a high level of professional review to ensure data quality.  References were further evaluated for the use of acceptable test methods and controls, including use of standardized methods for test media preparation and exposure and analytical chemistry, where present.   Our review focuses on species that lend themselves to controlled laboratory toxicity tests with oil and treated oil using standardized test methods.  For aquatic species this is typically limited to fish and invertebrate species.  This review does not include seabirds and mammals.  Laboratory exposures to oil and treated oil with upper trophic levels are problematic due to complex routes of exposure, long term response times, and ethical considerations.  In a review to oil toxicity on sea birds, Leighton (1993) notes that there are three different ways in which such oils may affect birds.  External contamination of plumage is the most common form of exposure. Feathers are coated in oil, they become matted, and lose their critical properties of water repellency, insulation, and flight. Death results from combinations of hypothermia, starvation, and drowning.  Avian embryos are highly sensitive to oil that contaminates the egg shell; amounts as little as 1–10 μL are lethal to embryos during the first half of incubation. Birds ingest oil when preening oiled plumage or ingesting oiled nutrients. At least three toxic effects of ingested oil are well documented: a nonspecific response as a stressor that is additive or synergistic with those of other stressors, impairments in reproduction ranging from lowered fertility to abandonment of reproductive effort, and severe oxidative damage to red blood cells. The effect of oil pollution on bird populations is very difficult to document and is likely to remain uncertain due to the many ecological factors that may occur in association with an oil spill event.

Similarly, the mammalian VECs (baleen and toothed whales, ice seals, polar bears, and walruses) cannot be studied using standard laboratory testing models.  Much of the data regarding oil and treated oil effects on marine mammals rely on field exposures during spill events in uncontrolled exposures (Geraci and St. Aubin 1989).  Furthermore, effects occur over a long period of time and rely more sub-lethal effects based on biomarker studies.  Important routes of exposure can include dermal exposure, inhalation at the sea surface, and ingestion.  As with seabirds, dermal contact is an important route of exposure for marine mammals (Engelhardt 1983; NRC 2003).  Baleen whales may also be impacted by oil fouling of the baleen (Braithwaite et al. 1983).  Feeding studies with captive seals (Engelhardt 1982) have been associated with changes in blood chemistry and immune system function.

There is a body of evidence that Arctic species are not more sensitive than temperate species to petroleum related compounds (single PAH, chemically dispersed, physically dispersed) when using acute toxicity tests. Moreover, there is a growing body of information that indicates that chemically dispersed oil may not be more toxic than physically dispersed oil. Some studies reveal that species assemblages and community functioning respond differently in the Arctic in response to oil exposure when compared to temperate communities. More effort could be devoted to study subtle long-term effects at the community and population level either by carrying out higher tier exposure studies or by extrapolating individual responses to population and community level effects using predictive models.

6.1.1 General Methods and Relevant Endpoints in Laboratory Testing

The methods for evaluating the toxicity of oil-water emulsions, physically and chemically dispersed oil and oil in sediment are varied and have evolved over the recent decades.  While there are accepted methodologies for test media preparation and exposure, there is still a wide range of methods applied.  Methods for preparing petroleum-water mixtures for toxicity testing were first developed by Anderson et al. (1974) and have been modified by several investigators to create more consistent petroleum-water and petroleum-dispersant mixtures.  Other guidance documents, such as those provided by UNESCO, the American Society for Testing and Materials (ASTM), and the US Environmental Protection Agency (USEPA) provide standardized test methods that may be adopted for conducting toxicity tests with fish and invertebrates.  The Chemical Response to Oil Spills Ecological Effects Research Forum (CROSERF) program developed and published methods for creating preparation of WAF as well as CEWAF and conducting toxicity tests (Aurand and Coelho 2005).  Since the procedures were put forth by the CROSERF program, they have been critically reviewed by Barron (2003), Clark et al. (2001), and Nordtug and Oistein (2007).  Barron (2003) suggested changes to the CROSERF methods that would make exposures more suitable for understanding effects in open, cold-water environments.  Other groups (SINTEF, Akvaplan-Niva, and NOAA) have developed methods for creating oil-water dispersions (OWD).  Many of the studies summarized here have included methods developed by CROSERF with the modifications to better emulate conditions in the Arctic.  A summary of several key considerations are presented in the following sections.

6.1.1.1 Test Exposure

The type of exposure that is used in toxicity evaluations may influence the effect-level estimates for individual species. Two different types of exposures, continuous and spiked exposures, are commonly used in laboratory evaluations of oil and treated oil. Continuous exposures are conducted with a constant dose throughout the test period. This is accomplished by either performing renewals with newly prepared test media or by performing static tests. Continuous exposures represent a maximum exposure to test media and represent field conditions where the petroleum concentrations are expected to be more or less constant over the period of several days to weeks. Spiked exposures are conducted with the doses that are highest at test initiation and then decrease throughout the test period. A number of researchers have noted that spiked exposures better mimic the petroleum exposures in the water column, particularly when chemical dispersants are used (Singer et al. 1991; Gardiner et al. 2013, in press; Fuller et al. 2004). When conducted side-by-side, continuous exposures typically show lower LC50 or EC50 estimates (greater toxicity) than spiked exposures (Singer et al. 1991). It is therefore important to consider the toxicity exposure method and the spill scenario when applying toxicity data to ecosystem-level evaluations of OSRs for different environmental compartments. 

6.1.1.2 Test Media Preparation

The test media can be prepared in a number of ways, with each method having advantages and disadvantages. In the following literature review, water-soluble fractions (WSF) are typically created using a generator column or gravel column. In such preparations, water is pushed through an oil-coated matrix, with the soluble fractions “dosing” the seawater. Recently, SINTEF developed a droplet generator that allows for the formation of physical dispersions with neutrally buoyant droplets of a diameter similar to chemically dispersed oil. This allows for a more direct comparison of physically and chemically dispersed oil. 

6.1.1.2.1 Water Soluble Fractions (WSF)

Water-soluble fractions (WSF) represent the soluble fractions of oil and are typically created using a generator column or gravel column. In such preparations, water is pushed through an oil-coated matrix, with the soluble fractions “dosing” the seawater. Oil is coated on to granular materials (e.g. gravel, or glass beads) and contained in a column having an entry port and exit port to accommodate the flow of filtered seawater through the column. Generally, fresh oil is added initially and naturally weathers over time. Adult female Calanus glacialis and Calanus finmarchicus were exposed to water soluble fraction (WSF) crude oil in a continuous flow experiment; hatching success of eggs and production of fecal pellets were assessed (Jensen and Carroll 2010). The WSF solutions were prepared by pumping filtered seawater through columns containing glass beads coated with crude oil (Duesterloh et al. 2002; Camus and Olsen 2008). Chemical analyses of WSF for both experiments included polynuclear aromatic hydrocarbons (PAH; based on USEPAs 16 priority compounds). Authors noted that for the C. glacialis experiment only 4 PAHs of the 16 analytes measured exceeded detection limits in the low and high dose treatments (naphthalene, fluorene, phenanthrene, and acenaphthylene). The low dose (3.6 μg/L PAH) treatment had a much lower naphthalene concentration relative to the high dose treatment. Total concentrations of 16 PAHs gradually decreased over time (naphthalene decreased more rapidly whereas heavier compounds were more stable). Results from C. glacialis tests showed no significant differences between highest treatment (10.4 μg/L PAH) and controls for cumulative egg or fecal pellet production; however, hatching success was significantly reduced. Exposure of C. finmarchicus to 7.0 μg/L PAH reduced feeding efficiency of adults. 

6.1.1.2.2 Water Accommodated Fractions (WAF, CEWAF)

Water-accommodated fractions (WAF) are a preparation method where oil is placed over a volume of seawater and mixed with a magnetic stirrer, generally for 18 to 24-hours. The amount of mixing energy may vary, but is typically adjusted such that visible oil droplets do not become entrained in the seawater. The mixing period is then followed by a “resting” period which allows larger oil droplets to rise to the surface. The test media are removed via a port at the bottom on the mixing container. WAF preparations are intended to include the dissolved fraction and the smaller, neutrally buoyant oil droplets. A dispersant can be added to the WAF preparation to create a dispersed WAF or Chemically-Enhanced WAF (CEWAF). Methods for preparing petroleum-water mixtures for laboratory toxicity testing were first developed by Anderson et al. (1974) and have since been modified by several investigators to create more consistent petroleum-water and petroleum-dispersant mixtures that are more representative of field conditions (Singer et al. 2000, Nordtug and Oistein 2007, Aurand and Coelho 2005; Barron 2003, Clark et al. 2001). Many of the studies conducted with Arctic species have been conducted with WAF or CEWAF preparations. 

6.1.1.2.3 Oil-in-Water Dispersions (Oil Droplets)

An additional method has been developed to create oil-in-water preparations that may better approximate the distribution of smaller, neutrally buoyant oil droplets in the physically dispersed oil. Nordtug et al. (2008) present a method for continuous and predictable in-line production of oil dispersions with defined size distribution of oil droplets based on theoretical models for droplet formation. Test solutions enter the experimental chamber via multiple nozzles which create turbulent dispersions of oil droplets and water soluble components. The mean droplet size and the droplet size distribution was determined by a combination of energy input (flow rate) and the number of nozzles used. The system enables simultaneous comparison between the effects of different concentrations of the dispersion and their corresponding water soluble fractions and the net effects of the oil droplet fractions. Flow-through into each individual exposure chamber was controlled by pumping exposure solution out near the surface, causing fresh solution to flow passively by entrainment into the chamber at a corresponding rate. Chemical measurements have been used to verify concentrations and quality of the dispersions and water soluble phases. The hydrocarbon profiles of the oil and the dispersions were very similar, whereas only a fraction of the components are in the dissolved phase. Tests with early life stage cod larvae (Gadus morhua) and copepods (C. finmarchicus) have been conducted with OWD preparations, allowing for evaluations of both the toxicity associated the dissolved fractions as well as adhesion of particulate oil (Olsvik et al. 2010 and Melbye et al. 2006). 

6.1.1.2.4 Oil Type/Weathering

The type of oil used in test media preparation may affect toxicity, as well as factors like fresh or weathered oil is used. After oil is released into the environment, its chemical and physical characteristics change through processes of: evaporation, dispersion, emulsification, dissolution, oxidation, and biodegradation. The weathering process is described in more detail in other sections of this report. For laboratory toxicity studies, typically weathering is simulated by evaporation causing a loss of the most volatile hydrocarbons and a reduction in the oil volume. The volatile hydrocarbons generally represent the more acutely toxic fractions of oil. It is important to note that the evaporative weathering only accounts for one portion of the weathering process. Many studies are conducted with fresh oil as it represents a worst-case scenario, particularly for acute toxicity estimates. There have been several toxicity studies with Arctic species that have used weathered oil in test media. 

Photo 6-2: Preparing for bioassay test under Arctic conditions (William Gardiner)
Photo 6-2: Preparing for bioassay test under Arctic conditions (William Gardiner)

6.1.1.2.5 Exposure Concentrations

The manner in which exposure concentrations are presented can also affect the usefulness and comparability of data. Measured effects concentrations may be reported as a percentage of the original test media (nominal concentrations) or as measured concentrations. Toxicity reported as nominal concentrations represent a percentage of a particular loading rate and are most closely tied to the laboratory method. Nominal concentrations are perhaps the most difficult data to relate to field data or to use in models, unless coupled with measured concentrations of hydrocarbons. Measured concentrations can include several analytical methods for expressing petroleum hydrocarbons. TPH are the most commonly reported measure for both laboratory and field studies. As such, they allow for comparisons between lab and field exposures. However, TPH includes many petroleum compounds that do not contribute to toxicity. This can be a significant source of variability and error when using toxicity data in models. Total petroleum aromatic hydrocarbons (PAHs) are believed to be more specific to the toxic fractions; however, this measure also includes non-toxic fractions. Finally, toxicity may be expressed in terms of a specific compound, such as naphthalene or 2-methylnaphthalene. While this removes the concern of including too many compounds, it also may not account for toxicity associated with secondary hydrocarbons. In many cases, it is important to understand how data are being applied and their limitations. 

Historically, the toxicity of oil, dispersed oil and single compounds was evaluated using continuous exposures, with test concentrations remaining more or less constant over the test period. In some cases, such as dispersed oil in open water, this may overestimate exposure. In such cases, spiked, or declining exposures have been used to better simulate field conditions (Photo 6-2). Each exposure method may be appropriate depending on the application of the data (e.g. spiked exposures for dispersed oil and continuous exposures for oil in ice). This topic is dealt with later in this chapter. 

6.1.1.2.6 Test Organisms

Where possible, test species that are included in toxicity studies are valuable ecosystem components (VECs) that represent the relevant environmental compartments potentially exposed to oil or treated oil. The VECs that were selected for the Arctic are discussed in Chapter 2. However, species that are VECs do not necessarily lend themselves to toxicity studies. Characteristics for suitable test organisms include the following: 

  • A sensitivity to oil and treated oil; 
  • Relative abundance and an availability for collection or culture; 
  • The ability to withstand laboratory handling; 
  • Meaningful and measurable endpoints over the time period of the study; 
  • Native to the site-specific conditions (e.g. cold water). 

There were toxicity data for a number of Arctic VECs including copepods, Arctic and Atlantic cod, sculpin, capelin, ice amphipods, benthic amphipods, crab, shrimp, and echinoderms. While there were toxicity data for a number of the VECs, in many cases the data are limited and endpoints. The suitability of data for conducting ecosystem-level evaluation of different OSR alternatives is considerably more limited, with copepods and Arctic cod having the most complete data sets. 

6.1.1.2.7 Test Endpoints and Exposures

Endpoints, or measures of effect, are dependent upon the test organism and the study goal. In general, toxicity is expressed as a dose-response to a certain concentration of test media. The most common endpoint is mortality, with the median Lethal Concentration (or LC50) being that concentration of test media that will cause 50 % mortality to the exposed organisms. Sub-lethal effects such as growth or reproduction are usually expressed as a median Effect Concentration (EC50) representing the concentration at which 50 % of the exposed organisms have a significant change in those end-points. These indicators are often extrapolated to an assessment endpoint depending on the purpose of the assessment. Effect concentrations can for example be used to assess environmental threshold values like a Predicted Environmental No-effects Concentration (PNEC) or to construct a species sensitivity distribution used to assess the fraction of species that is at risk. Other common measures of toxicity include No-observed Effect Concentrations (NOEC) and Lowest-Observed Effects Concentrations (LOEC). 

6.1.1.2.8 Data Extrapolation and Population Models

In order to evaluate population or ecosystem level effects, there are a variety of extrapolation models that utilize the toxicity endpoints (e.g. LC50) as input. Population models coupled with field-collected data on the distribution, abundance, natural mortality and fecundity are used to predict population level effects associated with modeled exposure scenarios. Simulated exposures using physical fates models such as the Integrated Oil Spill Impact Model System (SIMAP) allow for different spill and response scenarios to be evaluated. Gallaway et al. (2007; 2013, in prep) used toxicity data from laboratory exposures of Arctic cod to WAF and CEWAF, field collected data from the western Beaufort Sea, and the SIMAP model simulations to predict potential population effects from physically and chemically treated oil in open water. 

Another approach to population models is to incorporate many ecological components of an ecosystem to predict the potential cascade effects and repercussion in the trophic chain over time (Peterson et al. (2003). An example of such a model is the SYMBIOSES model that is under development for the Barents Sea (Carroll and Smit 2011; DeLaender et al. 2010). The main objective of such an impact analysis tool is to examine extent of effects on a VEC (e.g. larval Arctic cod) that result in significant effects on fish recruitment and stock size. While this is similar to the goal of the model used by Gallaway et al. (2007; in prep), this ecosystem approach considers indirect effects of a discharge that originates from ecological interactions. Data used to quantify effects of oil associated compounds on the Arctic trophic linkages include many food web links from the planktonic community to fish larvae and adult fish. 

The SYMBIOSES model integrates food web analyses of the structure and functions of the whole microbial and plankton community of the Barents Sea (bacteria, protozoa, phytoplankton, zooplankton, fish eggs and fish larvae) in terms of response to oil; for instance using an in situ mesocosm experiment. Today, such mesocosms exist and can be used in boreal or Arctic ecosystems to address plankton community responses to petroleum discharges (Hjorth et al. 2008). The mesocosm could be deployed in the Arctic waters during the ice free season especially at the peak of the phytoplankton bloom when the ice break-ups occur. While mesocosms represent an increase in the level of complexity of exposure design, they are subject to similar artifacts as laboratory studies. Interactions observed in the laboratory should be confirmed with field data when possible. 

Empirical toxicity data may also be used as part of predictive environmental fate and effect models that aim to integrate the chemical relationship of petrochemicals with prevailing biological processes (Figure 6-1). To form a holistic ecosystem perspective, complex relationships have been simplified according to basic assumptions, such as: 1) Hydrocarbon block and Quantitative Structure Activity Relationship (QSAR) models, i.e. hydrocarbons sharing similar structure and physiochemical properties will behave similarly in the environment; 2) Partitioning theory, i.e. chemicals belonging to similar structural groups will partition to organic carbon, water, and biological membranes in a similar manner and further, most hydrocarbons partition to lipids within the organisms (e.g. Target Lipid Model; McGrath et al. 2004)); 3) Critical body residues (CBR) coupled with median effect levels will provide a good estimation of impact or injury at the species level of investigation (threshold response levels have been summarized for many species groups); 4) Sensitivity of early life stages – incorporation of data representing early life stage endpoints result in the most conservative estimation of population disruption (e.g. OECD guidelines have been set for fish embryo toxicity; LC50 values greater than 100 mg/L are considered non-toxic). 

Hydrocarbon models, such as PETROTOX, organize the numerous hydrocarbon compounds in solution into structurally based ‘blocks’. If the behavior and toxicity of each block is known, the model can represent the relative predicted toxic activity level of the combined compounds present in different environmental compartments and estimate the resulting toxicity. Such an approach relies on laboratory estimates of toxicity for a variety of hydrocarbon constituents. This model assumes additivity of the toxicity for each block which may not necessarily be true. At this point in time, the quantity and quality of data with specific hydrocarbons for Arctic species are limited, however, if temperate data are also used, this approach may be feasible. 

The CBR approach couples toxicity data with tissue residue levels to predict effects. Such an approach is based on the premise that the toxicity of oil and treated oil is related to those hydrocarbons taken into the tissues. While this may have merit for exposures that occur over a long enough period to allow for uptake, it does not account for acute effects from short term exposures or physical effects due to adhesion. This method also relies on tissue residue data which are limited for Arctic species. 

6.2 Knowledge Status

Table 6-1 presents a summary of Arctic-relevant studies included in this review (link).  These studies are summarized in the following sub-sections.

6.2.1 Species represented in the data set

Amphipods are the most studied in toxicity testing with 11 different species. This is not surprising since amphipods are a very diverse and abundant taxa with species represented in all aquatic habitats of the arctic and they are relatively easy to catch and maintain in a laboratory. Fish and bivalves share an equal number of 8 species per group.  The predominant number of studies are the Arctic or Polar cod (Boreogadus saida) and the North East Atlantic cod (Gadus morhua) followed by the decapods (4 species), gastropods (4), copepods (3), isopod (2), Mysidacea (2) and lastly pycnogonidae (1) Cnidaria (1) and echinoderms (1). Although only three copepods species have been used in laboratory testing, a considerable number of studies have been identified: 14 papers with toxicity tests performed at low temperature, 4 at warmer temperature, and 9 papers are in preparation.  A few studies (5) report community level effects, either as a mesocosm or field study with relatively few studies (9) that examined the impact on the early life stages (4 with copepods, 3 with fish species, and one with amphipod).

6.2.2 Arctic ecosystem compartments in the dataset

6.2.2.1 Pack ice

Five different animal species (1 fish, 1 copepod and 3 amphipod species) have been used in toxicity testing comprising a total number of 31 studies. Five papers have investigated the biological effects of petroleum related compounds on sea ice bacteria and algae.

6.2.2.2 Pelagic

Toxicity data evaluating the effects of oil and treated oil on pelagic communities were dominated by studies on three species of copepods (14 studies) and fish (20 studies).  Two studies dealt with the effects of oil on water-column phytoplankton.  None of the studies included euphausiids (krill).

6.2.2.3 Benthic

The benthic compartment has the most diverse group of species used in ecotoxicological research with 32 different animal species dominated by crustaceans (17 species) and molluscs (11 species). Threemacroalgae species have also been reported. Eight studies at the community level (whole community, meiofauna, microalgae and bacteria) could be identified.

6.2.3 Review by Taxa

6.2.3.1 Phytoplankton and seaweed

The growth of different Arctic marine phytoplankton and macroalgae exposed to oil in seawater and Corexit (formulation not provided in paper) dispersant were evaluated by Hsiao (1978).  Primary production rates varied with type and concentration of crude oil and species composition.  Diatom growth (Chaetoserus septentrionalis, Navicula bahusiensis and Nitzschia delicatissima) was inhibited at 10 days of exposure at 10 ppm; whereas, growth in green flagellate (Chlamydomonas pulsatilla) was stimulated by two different oil treatments.   Primary production of natural populations of phytoplankton from different locations in the Beaufort Sea ranged from stimulation to inhibition.  Primary production of macroalgae (Laminaria saccharina and Phyllophora truncate) collected from the Beaufort Sea was inhibited by the different oil treatments. 

The effects of oil and dispersed oil exposure on the macroalgae Stictyosiphon tortilis, Dictyosiphon foeniculaceus, and Pilayella littoralis were evaluated during a surface and subsea field release of weathered oil dispersed with Corexit® 9527 (Cross et al. 1987a).  There were no detectable adverse effects on biomass, diversity, or reproductive condition of the shallow subtidal macroalgae.  Similarly Cross (1987) found no decrease in cell densities, chlorophyll a concentrations or productivity in the ice algal communities dominated by the pennate diatoms Nitzschia grunowii and N. frigida; there was some evidence of slightly increased production in the untreated oil exposures.  Exposure concentrations in the sea ice were estimated to be 0.15 to 0.28 ppm in the oil-only exposures and 5.8 to 36.5 ppm in the dispersed oil exposures.  Cross (1987) notes that the first measures of production and biomass were made approximately two days after dispersed oil application and any immediate and transitory effects would not have been detected for any of the oil treatments.

In more controlled laboratory conditions, Van Baalen and O’Donnell (1984) grew cultures of the sympagic diatoms Nitzchia sp and Chaetoceros sp. collected from the ice edge in the Bering Sea the presence of two crude oils and two fuel oils.  Growth inhibition was observed for Nitschia sp. at concentrations >50 ppm for both the crude oil and fuel oil treatments.  Chaetoceros sp. growth was inhibited at concentrations between 50 and 500 ppm.  For both species, effects were greater in the 0°C treatments than in the 10 °C treatments.

Results of studies with Arctic phytoplankton and macroalgae are somewhat equivocal.  However, based on these studies there appears to be inhibition of sea ice algae and phytoplankton in the presence of oil and dispersed oil.  However, this effect does not appear to persist if petroleum concentrations decrease.  Effects may persist in sea ice if petroleum is incorporated into the sea ice.

6.2.3.2 Mysids

The mysid (Mysis oculata) is an epibenthic shallow-water invertebrate that is common throughout the coastal waters of the Arctic (Photo 6-3).  Mysis relicta is a boreal pelagic mysid found in colder waters of Baltic and Barents Sea.  Both species represent a component of fish diets, as well as some marine mammals and birds.  Mysids are considered sensitive indicators of aquatic toxicity and the temperate mysid,Americamysis bahia is one of the most common laboratory test species.

Photo 6-3: Mysid from Kongsfjorden (Ida Beathe Overjordet)
Photo 6-3: Mysid from Kongsfjorden (Ida Beathe Overjordet)

Photo 6-4: Calanus finmarchicus (Bjørn Hansen)
Photo 6-4: Calanus finmarchicus (Bjørn Hansen)

Photo 6-5: Calanus glacialis (Bjørn Hansen)
Photo 6-5: Calanus glacialis (Bjørn Hansen)

Acute lethal toxicity levels were derived for young-of-the-year mysids (M. oculata) exposed to oil in water dispersions (OWD) and WSF of Norman wells crude oil (Riebel and Percy 1990).  The 96-hour LC50s for continuous exposures to OWD were 4.51 to 7.57 ppm nominal.  In the water-soluble fractions, LC50 values were approximately 0.5 ppm nominal.  No apparent physical effects were observed in the OWD treatment.  The 96-h LC50 for the mysid M. relicta was 2.7 ppm, in continuous exposures to water-soluble fractions of Cook Inlet crude oil.

6.2.3.3 Copepods

As noted above, there have been a number of studies focusing on the sensitivity of Arctic copepods to oil and treated oil.  Copepods are an Arctic keystone species, transferring energy from lower to higher trophic levels of the Arctic and sub-Arctic food web (Photos 6-4, 6-5). 

Hansen et al. (2011) evaluated the survival and WAF-mediated induction of the gene encoding glutathione S-transferase (GST) for the Arctic copepod Calanus glacialis and the boreal copepod C. finmarchicus exposed to water-accommodated fractions of oil.  The 96-h LC50values differed between the two species with the Arctic copepod being less sensitive than the temperate-boreal species.  However, LC50s were similar when copepods were observed for a longer period of time.  The lipid-rich copepods survived longer than lipid-poor copepods at the same exposure concentration. In terms of GST expression, both species showed concentration-dependent and exposure time-dependent trends. The Arctic copepod appeared to respond slower and with a lower intensity indicating that a delayed response and that longer test duration is required to observe effects.

Photo 6-6. C. glacialis (Jack D Word)
Photo 6-6. C. glacialis (Jack D Word)

Gardiner et al. (2013) found similar results in tests with water-accommodated fractions and chemically dispersed WAF preparations, conducting 12-day spiked exposures with C. glacialis (Photo 6-6).  The mean LC50 for WAF with fresh Alaska North Slope crude was 4 ppm TPH, similar to the range of values (2.5 – 9.6 ppm) of Hansen et al. (2011) and Nordtug et al. (2013).   The mean LC50 for the dispersed WAF was 22 ppm and 62 ppm for early season and late seasonC. glacialis.  Similar to Hansen et al. (2011) the low lipid reserves likely affected the sensitivity of copepods in the early season.  Gardiner et al. (2013) found that when expressed as total PAH or parent naphthalene, the relative toxicity of the WAF and dispersed WAF was more similar, showing that TPH is a gross measure of exposure concentration and includes numerous constituents that do not contribute to toxicity.

Hansen et al. (2012) used a newly develop droplet generator to create physical dispersions of oil with droplets of a similar size distribution as those from chemically dispersed oil.  When the toxicity of the physical dispersions to C. finmarchicus were compared to those of the chemical dispersion, the 96-h LC50s differed by a factor of 1.6, whereas the LC10 (that concentrations the causes 10% mortality) and LC90 (that concentration that causes 90% mortality) differed by factors of 2.9 and 0.9, showing that at high effects levels, there was little difference between the physically and chemically dispersed oil and indicating that the full dose-response curve should be considered in risk assessments.

Jensen and Carroll (2010) evaluated the effects of continuous sublethal WSF exposures on reproduction and feeding for C. glacialis and C. finmarchicus.  A reduction in feeding was observed for C. finmarchicus at concentrations of 7 ppb total PAH and a decrease in egg hatching success was observed forC. glacialis.  

Photo 6-7. In situ burning test (Liv-Guri Faksness)
Photo 6-7. In situ burning test (Liv-Guri Faksness)

Faksness et al. (2012) conducted one of the few toxicity testing programs with the residuals of in-situ burning (Photo 6-7).  Seawater samples and oil collected immediately after in-situburning were evaluated for chemical composition and toxicity.  Acute toxicity tests with the marine copepod C. finmarchicus showed no increase in toxicity in the underlying water after in-situ burning.

A number of studies have evaluated the effects of pyrene on copepods including survival, egg hatching, development time and dynamic energy budget for C. glacialis and C. finmarchicus (Hjorth and Nielsen 2011; Grenvald et al. 2013; Klok et al. 2012, 2014).  Copepod survival and gene expression have also been evaluated for exposures with naphthalene.  Recent efforts are focusing on coupling copepod toxicity data with models for uptake and population level effects (Hansen et al., in preparation) and in many ways the data on copepods are perhaps the most complete for Arctic species.

An additional study conducted with C. finmarchicus by Melbye et al. (2001) focused on the uptake of hydrocarbons from dispersed oil droplets using a flow-through system with a drop generator.  This study was designed to specifically examine the uptake and bioaccumulation of the oil droplets.  The dispersion was created by using an ultrasonic probe and mixing 6.05 g of Totalfina oil in 500 mL of seawater for 10 minutes.  Water and organism tissues were collected and monitored over specific intervals throughout 14 days of testing. The copepods were field collected from the Trondheim fjord.  This species was chosen for uptake studies of dispersed oil because it actively filters water for particles in the 10 – 50µm size range.

As reported in Melbye et al. (2001), C. finmarchicus accumulated hydrocarbons in the concentration range of 200 to 1000 ppb, based on analysis of the stock solution.  The uptake of physically dispersed oil did increase from 7 to 14 days but results also showed that a significant portion of the oil was adsorbed to the outside surface of the copepods.

6.2.3.4 Amphipods

Photo 6-8 Amphipod (Jack D Word)
Photo 6-8 Amphipod (Jack D Word)

Amphipods are an essential component of the ice, benthic, and pelagic communities, forming a link between lower trophic levels (e.g. algae and copepods) and upper trophic levels (Photo 6-8).  Amphipods have also been used as sensitive indicators of toxicity in both sediments and waters, particularly hyperiid and gammarid amphipods.  Existing data include benthic, epibenthic, and sea ice amphipods.  Studies with hyperiid amphipods, an important prey resource, were not identified. The sea ice amphipod Gammarus wilkitzkii has been tested with water-soluble fractions of crude oil.  Hatlen et al. (2009) evaluated mortality, respiration, molting and oxidative stress of water-soluble fractions of crude oil on the Arctic ice amphipod, G. wilkitzkii in continuous exposures using rock-column oil-generator column.  No mortality or molting effects were observed at test concentrations ranging from 120 to 764 ppb total PAH; however, respiration effects and oxidative stress was observed indicating sub-lethal effects.  Camus and Olsen (2008) exposed eggs of the ice amphipod G. wilkitzkii to water-soluble fractions of crude oil for 30 days using continuous-flow oil generator columns.  A significant increase in embryo aberrations was observed in the highest test concentration (55 ppm total PAH in overlying water on Day 0).  Other sublethal endpoints, such as energy allocation and oxidative stress, have been evaluated for G. wilkitzkii in similar exposures.  Olsen et al. (2008) evaluated the energy budget of ovigerous females of G. wilkitzkii in continuous exposures to water soluble fractions of crude oil.  Sublethal exposures ranged from 5 to 55 ppm total PAH, decreasing over the 30-day exposure period.  Protein levels were significantly higher in the 10 ppm total PAH dose, potentially indicating a stress response.  Other measures of energy allocation were unaffected by the WSF preparations.

Two other sympagic (ice-associated) amphipods (Gammaracanthus loricatus and Onismus nanseni) were tested in a similar water-soluble fraction generator column (Carls and Korn 1985) with Cook Inlet crude oil.  The 8-day LC50 concentrations were 3.7 ppm and >1.7 ppm total PAH, respectively.  As has been shown with other Arctic species, there was a delayed response in the tests with ice amphipods, with 8-day or longer exposures better representing toxicity.

Benthic amphipods are commonly used in toxicity testing in many regulatory programs and can be sensitive indicators of impacts to nearshore or shelf benthic communities.  There have been numerous toxicity studies evaluating the effects of oil, dispersed oil, and individual PAHs on amphipods.   Benthic amphipods G. zaddachi, O. affinis, Anonyx nugax had LC50s ranging from 2.3 to 8 ppm total PAH in water dispersions. 

Tests with single PAH compounds have included total naphthalene and 2-methylnapthalene.  Carls and Korn (1985) report 96-h and 8-d median lethal concentrations for O. nanseni, A. nugax, and G. loricatusranging from 2.2 to 3.5 ppm.  Olsen et al. (2011) report 2-methylnaphalene 96-h LC50 values of 1.34 and 1.92 ppm for the amphipods Gammarus sp. and A. nugax.  In each case the relative sensitivity to single PAH compounds was remarkably similar to other species within the phylum (Pandalus borealis, Sclerocrangon boreas, Mysis sp.).

6.2.3.5 Benthic organisms

Photo 6-9: Decapod crab larvae (Ida Beathe Overjordet)
Photo 6-9: Decapod crab larvae (Ida Beathe Overjordet)

Photo 6-10: Larval shrimp (Bjørn Hansen)
Photo 6-10: Larval shrimp (Bjørn Hansen)

Echinoderms, molluscs, and epibenthic decapods (Photo 6-9) play an important role in supporting upper trophic levels (e.g. the clams Serripes sp. and Mya sp.), support fisheries (Pandalus borealis), and are known to be sensitive indicators of toxicity (Strongylocentrotus droebachiensis).  Olsen et al. (2011) report data for a number of benthic organisms tested with 2-methylnaphthalene including Chlamys islandica (scallop) adult S. droebachiensis (sea urchin), Acmea tessulata (gastropod snail), Littorina littorea (gastropod snail), Margarites helisina (gastropod snail), Pandalus borealis (shrimp; Photo 6-10), and Nympon gracile (sea spider).  Among the most sensitive were S. droebachiensis and A. tessulata whereas the snail M. helisina and the sea spider were the least sensitive.  The amphipods Gammarus sp. and A. nugax and the fish B. saida were also included in this evaluation.  The amphipods were similarly sensitive to other benthic invertebrates, while the B. saida was among the most sensitive to 2-methylnaphthalene.  This study represents one of the most comprehensive datasets for benthic and epibenthic species and is used to compare arctic and temperate species using species sensitivity distributions.

The effects of oil in sediment and dispersed oil have also been evaluated with the clams Mya truncata (Camus et al. 2003; Mageau et al. 1987) and Serripes groenlandica (Mageau et al. 1987), the sea urchin S. droebachiensis(Mageau et al. 1987) and the crab Hya araneus (Camus et al.  2002) and Chionocetes bairdi (Perkins et al. 2003).

6.2.3.6 Fish

Photo 6-11: Arctic cod (Brian Hester)
Photo 6-11: Arctic cod (Brian Hester)

Photo 6-12: Arctic cod in acclimation tank (Lionel Camus)
Photo 6-12: Arctic cod in acclimation tank (Lionel Camus)

Relatively few Arctic fish species have been tested with oil and treated oil.  Understandably, the majority of available data are for the Arctic cod, Boreogadus saida.  Arctic cod are ubiquitous throughout the Arctic, representing a critical link between the zooplankton community and higher trophic levels (e.g. seals, beluga whales).  B. saida is a truly pan-Arctic, occupying nearshore, pelagic, and sea ice habitats, residing both at depth and near the surface waters, depending upon age and season (Photos 6-11, 6-12).  The sensitivity of Arctic cod has been evaluated for both oil and dispersed oil.  Gardiner et al. (2013) conducted spiked exposures one-year-old B. saida with WAF and dispersed WAF to emulate open-water pelagic conditions.  Median-lethal concentrations were reported for TPH, total PAHs, and parent naphthalene.  Data were then used to compare the relative sensitivity of Arctic and temperate species using species sensitivity distributions.   Carls and Korn (1985) also report median lethal concentrations to 8-day continuous exposures of water soluble fractions (total PAH) and naphthalene. 

Chronic exposures have been conducted both with water-soluble fractions and oil-dosed food.  Nahrgang et al. (2009) conducted a series of studies with B. saida, exposing the fish to water-soluble fractions of oil, injecting fish with benzo (a) pyrene, and feeding prey items exposed to crude oil.  Following chronic (4 weeks) exposure to water soluble fractions, biomarkers showed evidence of PAH uptake and genotoxicty.  Christiansen (2000) and Christiansen and George (1995) evaluated the effects of oil-contaminated prey on cod appetite, growth and osmoregulation.  Food contaminated with 200 or 400 ppm crude oil did not affect appetite but resulted in decreased growth; foods contaminated with concentrations of 500 ppm were rejected.

Capelin (Mallotus villosus) is a keystone fish species, particularly in the Barents Sea.  As with B. saida, M. villosus represent the trophic link between zooplankton and larger fish, birds, and marine mammals.  Despite their importance, there were few studies evaluating the toxicity of oil or dispersed oil with capelin.  Khan and Payne (2005) exposed capelin to oil and dispersed oil (Hibernia and Corexit 9527) in continuous flow-through tests of 96-h.  Similarly, Frantzen et al. (2012) evaluated the toxicity of crude oil to M. villosus embryos.  Concentrations of 46 ppb total PAHs increased embryo mortality rates and decreased hatching success. 

Photo 6-13: Myoxocephalus sp. (William Gardiner)
Photo 6-13: Myoxocephalus sp. (William Gardiner)

Sculpin (Myoxocephalus sp.) are among the most common demersal fish species in the Arctic.  Toxicity tests have been conducted with oil and dispersed oil (Photo 6-13).  Gardiner et al. (2013) conducted spiked exposures to WAF and dispersed WAF with larvalMyoxocephalus sp.  Median lethal concentrations were generally lower than those for the Arctic cod; however, the sensitivity was similar when older cohorts of larval sculpin were compared.  Carls and Korn (1985) exposed the sculpin, Oncocottus hexacornis to water-soluble fractions of Cook Inlet crude, with median lethal concentrations >1.7 ppm total PAH.

Table 6-1.  Toxicological Investigations

TaxaHabitatChemicalExpEndpointAuthorsYear

Cnidaria (1 species)

Halitholus cirratus

P

Dispersed oil

A

Locomotion

Percy, Mullin

1977

Echinodermata (1 species)

Strongylocentrotus droebachiensis*

B, NS

Dispersed oil

Field

Physiological, behavioral

Mageau et al.

1987

2-MN

A

Survival

Olsen et al.

2011

Dispersed oil

Field

Bioaccumulation

Humphrey et al.

1987

Condensate

A,C

Survival, respiration, reproduction

Kosheleva et al.

1997

Pycnogonidae (1 species)

Nymphon gracile

B, NS

2-MN

A

Survival

Olsen et al.

2011

Amphipoda (11 species)

Anonyx nugax*

B, DW

WSF, Naphthalene

A

Survival

Carls, Korn

1985

2-MN

A

Survival

Olsen et al.

2011

Oil in sediment

 

Body burden

Neff, Durell

2012

Corophium clarencense

B, NS

Oil in sediment

A

Acute, avoidance

Percy

1977

Gammaracanthus loricatus

SI

WSF, Naphthalene

A

Survival

Carls, Korn

1985

Gammarus oceanicus*

B, NS

WAF

C

Respiration, osmoregulation

Aunaas et al.

1991

food tainted

A

Avoidance

Percy

1976

2-MN

A

Survival

Olsen et al.

2011

Gammarus setosus*

B, NS

WAF

C

Cellular energy allocation

Olsen et al.

2007

Pyrene

C

Bioaccumulation, biotransformation, excretion

Carrasco Navarro

2013

Gammarus wilkitzkii*

SI

WSF

C

Cellular energy allocation

Olsen et al.

2008

WSF

C

Embryo development

Camus and Olsen

2008

WSF

C

Oxidative stress, molting

Hatlen et al.

2009

None

C

Baseline oxidative stress

Krapp

2009

Gammarus zaddachi*

B, NS

Oil in sediment

A

Survival

Atlas et al. 1977

1977

Onisimus affinis*

B, NS

Oil in sediment

A

Avoidance

Percy

1977

Dispersed oil

A

Respiration

Percy

1977

Tainted food

A

Avoidance

Percy

1976

Oil in sediment

A

Survival

Atlas et al. 1977

1977

Oil Dispersion

A

Locomotion

Percy, Mullin

1977

WSF

C

Survival, movement, foraging

Budosh

1981

Onisimus litoralis*

B, NS

WAF

C

Cellular energy allocation

Olsen et al.

2007

None

 

Baseline oxidative stress

Nygård et al.

2010

Onisimus nanseni*

SI

WSF, Naphthalene

A

Survival

Carls, Korn

1985

Orchomenella pinguis

NS

Oil in Sediment

Field

Population

Bach et al.

2009

Oil in Sediment

Field

Population

Bach, Dahllöf

2012

Isopoda (2 species)

Mesidotea sibirica

NS

Oil in Sediment

A

Avoidance

Percy

1977

Mesidotea entomon

NS

Oil in Sediment

C

Avoidance

Percy

1977

Tainted Food

C

Avoidance

Percy

1976

Mysidacea (2 species)

Mysis oculata

B, NS

WSF, oil dispersion

A

Survival

Riebel, Percy

1990

Mysis relicta

B, NS

WSF

A

Survival

Carls, Korn

1985

Copepoda (3 species)

>Calanus glacialis*

P, SI, DW

WAF

A

Survival, gene induction

Hansen et al.

2011

WAF, CEWAF, Naphthalene, Dispersants

A

Survival

Gardiner et al.

2013

Dispersants

A

Survival

Hansen et al.

in prep

WSF

A

Survival

Jager, Hansen

in prep

Pyrene

A

Feeding, reproduction

Jensen et al.

2008

Pyrene

C

Egg, hatching, fecal pellet

Hjorth, Nielsen

2011

WSF

C

Feeding, reproduction

Jensen, Carroll

2010

Pyrene

C

Survival, hatching, development

Grenvald et al.

2013

Pyrene, WSF

C

Dynamic energy budget

Klok et al.

2012

Calanus finmarchicus*

P, DW

WAF

A

Survival, gene induction

Hansen et al.

2011

OWD, Dispersed Oil

A

Survival, feeding, adhesion

Hansen et al.

2012

ISB residue

A

Survival

Faksness

2012

Dispersed oil

A

Survival, reproduction, hatching success

Olsen et al.

2013a,b

WSF, dispersed oil

A

Survival, biomarkers, adhesion

Hansen et al.

2009

Naphthalene

A

Gene expression (GST, p450, HSP)

   

Dispersants

A

Survival

Hansen et al.

in prep

WSF

A

Survival, biomarkers

Jager, Hansen

in prep

WSF

A

Survival, development

Hansen et al.

in prep

OWD, Dispersed Oil

A

Reproduction, hatching success

Hansen et al.

in prep

OWD, Dispersed Oil

A

Body burden, fouling, adhesion, filtration rates

Nordtug et al.

in prep

Monoaromatics

A

Survival

Nordtug et al.

in prep

WSF

A

Survival

Hansen et al.

in prep

WSF, dispersed oil

A

Survival, body burden

Hansen et al.

in prep

Phenanthrene, BaP

C

Bioaccumulation

Jensen et al.

2012

Pyrene

C

Egg, hatching, fecal pellet

Hjorth, Nielsen

2011

Pyrene

C

Survival, hatching, development

Grenvald et al.

2013

Microsetella norvegica

P

Pyrene

A

Survival, feeding, RNA/DNA

Hjorth, Dahllöf

2008

Decapoda (4 species)

Myas araneus

B

Oil in sediment

C

Heart rate, respiration, oxidative stress

Camus et al.

2002

Pandalus borealis*

B, DW

2- MN

A

Survival

Olsen et al.

2011

WAF

C

Survival, Lysosomal stability, DNA

Bechmann et al.

2010

Sclerocrangon boreas

B, NS

2-MN

A

Survival

Olsen et al.

2011

Chionoecetes bairdi*

B

WAF

A

Survival

Perkins et al.

2003

Bivalvia (7 species)

Astarte borealis

 

Dispersed oil

Field

Bioaccumulation

Humphrey et al.

1987

Chlamys islandica

B, NS

BaP

A

Injected dose; membrane stability, oxidative stress

Camus et al.

2002

2-MN

A

Survival

Olsen et al.

2011

Dispersed oil

C

Immune function

Hannam et al.

2010

Dispersed oil

C

Immunotoxicity, oxidative stress

Hannam et al.

2009

Dispersed oil

C

Enzymes, lysosomes, DNA

Baussant et al.

2009

Dispersed oil

 

Survival, growth, biomarker

Frantzen et al.

2013

None

A

Biomarker baseline

Nahrgang et al.

2012

Liocyma fluctuosa

B, NS

WAF

C

Cellular energy allocation

Olsen et al.

2007

Macoma calcarea*

B, NS

Dispersed oil

Field

Bioaccumulation

Humphrey et al.

1987

Dispersed oil

Field

Histology, biochemistry

Neff et al.

1987

Mya truncata

B, NS

Oil in sediment

C

Survival, feeding RNA/DNA

Hjorth, Dahloff

2008

Dispersed oil

Field

Respiration, membrane stability, oxidative stress

Camus et al.

2003

Dispersed oil

Field

Behavior

Mageau et al.

1987

Mytilus edulis*

B, NS

Phenanthrene

C

Lysosomal stability

Camus et al.

2000

Dispersant

C

Freezing tolerance, osmoregulation

Aasset, Zachariassen

1983

Dispersed oil

C

Enzymes, lysosomes, DNA

Baussant et al.

2009

Dispersed oil

Field

Histology, biochemistry

Neff et al.

1987

Serripes groenlandicus*

B

Dispersed oil

Field

Behavior

Mageau et al.

1987

Dispersed oil

Field

Bioaccumulation

Humphrey et al.

1987

Gastropoda (4 species)

Acmaea tessulata

B, NS

2-MN

A

Survival

Olsen et al.

2011

Littorina littorea

B, NS

2-MN

A

Survival

Olsen et al.

2011

Littorina obstusata

B, NS

Condensate

A, C

Reproduction

Kosheleva et al.

1997

Margarites helicina

B, NS

2-MN

A

Survival

Olsen et al.

2011

Fish (8 species)

Anarhichas minor

B

b-naphthoflavone

C

EROD (gills)

Jönsson

2003

Dispersed oil

C

Growth, biomarker

 

 

Boreogadus saida*

P, SI, DW

2-MN

A

Survival

Olsen et al.

2011

WSF, Naphthalene

A

Survival

Carls, Korn

1985

WAF, CEWAF, Naphthalene

A

Survival

Gardiner et al.

2013

Injected BaP

A

DNA adducts

Aas et al.

2004

Injected BaP

A

Gene expression, PAH metabolites, enzyme activity

Nahrgang et al.

2009

WSF

C

Gene expression, metabolites, DNA damage, reproduction

Nahrgang et al.

2010

Tainted food

C

Gene expression, PAH metabolites, enzyme activity

Nahrgang et al.

2010a

Tainted food, WSF

C

Biomarkers

Nahrgang et al.

2010b

WSF

C

Whole fish metabolism

Christiansen

2009

Tainted food

C

Growth, food selection, appetite

Christiansen, George

1995

Tainted food

C

Osmoregulation

Christiansen

2000

Tainted food

C

Biomarker, CYP1A

Gorge et al.

1995

Injected BaP

C

Excretion, radiography

Christiansen et al.

1996

Injected BaP

C

Excretion, radiography

Ingebritsen et al.

2000

Dispersed oil, nonylphenol

C

Biomarker, baseline

Jönsson et al.

2010

None

 

Biomarker, baseline

Nahrgang et al.

2010c

Clupea herengus

P, NS

Dispersed oil

C

Survival, development

Ingarsdóttir et al.

2012

Cyclopterus lumpus

B, NS

Dispersed oil

A,C

Survival, biomarker

Frantzen et al.

in prep

Gadus morhua*

P

Dispersed oil

C

Biomarkers

Skadsheim et al.

2009

WSF, dispersed oil

C

Transcriptional effects

Olsvik et al.

2010

WSF, dispersed oil

C

Transcriptional effects

Olsvik et al.

2011a,b

WSF, dispersed oil

C

Transcriptional effects

Olsvik et al.

2012

WSF, dispersed oil

C

Survival, food assimilation

Nordtug et al.

2011

Condensate

A,C

Hematology, respiration

Kosheleva et al.

1997

None

 

Biomarker baseline

Nahrgang et al.

2013

Mallotus villosus*

P

WSF

C

Survival, development

Frantzen et al.

2012

Myoxocephalus sp.*

B, NS

WAF, CEWAF, Naphthalene

A

Survival

Gardiner et al.

2013

Oncocottus hexacornis

B, NS

WSF

A

Survival

Carls, Korn

1985

Algae (12 species)* 

Chaetoceros septentrionalis

P

WSF, dispersed oil

A

Growth

Hsiao

1978

Navicula bahusiensis

P

WSF, dispersed oil

A

Growth

Hsiao

1978

Chlamydomonas pulsatilla

P

WSF, dispersed oil

A

Growth

Hsiao

1978

 Nitzchia delicatissima

P

WSF, dispersed oil

A

Growth

Hsiao

1978

Halosphaera viridis

P

Gas condensate

A,C

Photosynthesis, abundance

Kosheleva et al.

1997

Phizosolenia sp.

P

Gas condensate

A,C

Photosynthesis, abundance

Kosheleva et al.

1997

Nizschia sp.

SI

WSF

A

Photosynthesis

Van Baalen, O’Donnell

1984

Chaeotoceros sp.

SI

WSF

A

Photosynthesis

Van Baalen, O’Donnell

1984

Ice algae *

SI

Dispersed Oil

Field

Density, Chl a, production

Cross

1987

Stictyosiphon sp.

B

Dispersed oil

Field

Growth

Hsiao

1978

Pilayella littoralis

B

Dispersed oil

Field

Growth

Hsiao

1978

Dictyosiphon foeniculaceus

B

Dispersed oil

Field

Growth

Hsiao

1978

Laminaria saccharina

B

Oil, dispersant

Field

Growth

Hsiao et al.

1978

Phyllophora truncata

B

Oil, dispersant

Field

Growth

Hsiao et al.

1978

Phytoplankton community

B

Pyrene, UV

C

Nutrient uptake

Petersen, Dahllöf

2007

Communities

Sediment community

B, NS

Oil in water

C

Respiration

Olsen et al.

2007

Sediment community

B, NS

Oil in water

C

Diversity

Gulliksen, Tassen

1982

Hard bottom macrofauna

B, NS

Dispersed oil

Field

Diversity, density, community structure, avoidance

Cross et al.

1987

Under-ice sediment

B, NS

Dispersed oil

Field

Diversity, abundance, density

Cross, Marten

1987

Benthic algae and bacteria

B, NS

Pyrene, UV

C

Respiration, diversity

Petersen et al.

2008

Pyrene, UV

C

Diversity, abundance

Brakstad et al.

2008

   

Pelagic microbial community

P

WAF, CEWAF

C

Respiration, diversity

McFarland et al.

in prep

Benthic bacterial community

B, NS

Oil, PAHs

A

Nitrogen fixation

Knowles, Wishart

1970

Oil

C

Nitrogen fixation, denitrification, CO2, methane

Griffiths et al.

1982

   

Ice bacterial community

SI

Oil

C

Diversity, abundance

Brakstad et al.

2008

Oil

C

Diversity

Gerdes et al.

2005

   

Compilations

Multiple species

Various

WAF, naphthalene, 2-MN

A

Survival

DeHoop et al.

2012

Multiple species

Various

WAF, naphthalene, CEWAF

A

Survival

Word and Gardiner

2013

 

ABBREVIATIONS
HabitatChemicalsExposure

P:

Pelagic

2-MN

2-methyl-naphthalene

A:

Acute

B:

Benthic and Epibenthic

BaP

Benzo(a)pyrene

C:

Chronic

SI:

Sea ice

 

NS:

Nearshore

 

DW:

Deepwater

*

Valuable Ecosystem Component

6.3 Discussion

6.3.1 Petroleum related components

6.3.1.1 Crude oil

Approximately 56 studies have been conducted with crude oil.  There have been a number of different oil types included in the different testing programs, as well as different types of preparations with untreated (soluble fraction or physically dispersed) or treated oil (chemical dispersant or ISB). Peer-reviewed studies with photochemically enhanced toxicity (photooxidation) with Arctic species are not available. However, photoenhanced toxicity of physically and chemically dispersed Alaska North Slope crude oil has been evaluated with Pacific herring eggs and larvae (Clupea pallasi; Barron et al. 2003). 

Exposure systems vary between studies although there is an increasing trend to harmonize the approaches between laboratories.  Studies have been conducted as continuous or declining exposures, with WAF, dispersed WAF, or physical dispersions.  It is noteworthy that relatively few types of chemical dispersants have been used in toxicity studies, primarily Corexit®9500 and Dasic Slickgone. The vast majority of the exposures were waterborne and five studies were diet borne exposures.

Relatively few studies have included weathered oil.  There are some studies (e.g. Hansen et al. (in prep) have tested weathered and unweathered oil with the copepod C. finmarchicus. Given that OSR response times may be delayed in the Arctic due to the potential distances and inclement weather, this should be a future focus of study.

6.3.1.2 Single PAH

Single PAH toxicity data are an important component in developing predictive models and for comparing effects between different OSR alternatives.  Most studies performed with single compounds included 2-methyl naphthalene, naphthalene, and pyrene.  Fewer studies have evaluated benzo(a)pyrene and phenanthrene and primarily when evaluating uptake and tissue residues.  The naphthalene compounds are often included in testing programs since they are predominant in the crude oil and are known to elicit most of the acute toxicity.  Pyrene and benzo(a)pyrene have been used as model compounds to study a specific biological mechanism (uptake rate, metabolism, enzyme induction).

6.3.2 Chemically dispersed oil versus physically dispersed oil

An increasing number of research programs are being conducted with Arctic species exposed to chemically dispersed oil and physically dispersed oil. Such data allow for a direct comparison of two OSR alternatives and provides responders with information to determine whether chemical dispersants should be used on an oil spill in the Arctic.  For Arctic species this research has focused on calanoid copepods and fish larvae (e.g. Hansen et al. 2012; Gardiner et al. 2013).  An additional paper in preparation addressed the same issues with shallow water benthic species (Frantzen et al. 2013, in prep.). While only two chemical dispersants were used (Corexit®9500 and Dasic Slickgone) and three different exposure systems between laboratories were employed, the investigators concluded that the same concentration of chemically dispersed oil is not more toxic than physically dispersed oil.  These conclusions suggest that chemical dispersants do not increase the toxicity of oil.

6.3.3 Are Arctic species more sensitive than temperate species?

Arctic species require a longer period of time to exhibit effects associated with petroleum exposures.  A number of researchers have noted that Arctic species of fish (B. saida) and copepods (C. glacialis and C. hyperboreus) require up to 12-days to show effects associated with acute exposures in either spiked or typical 96-hour continuous exposures (Chapman and Riddle 2005; Olsen et al. 2011, Gardiner et al. 2013; Hansen et al. 2013).  Many factors may affect the increased response time of Arctic species to oil and treated oil.  Arctic species have a number of morphological and physiological adaptations to survive at cold temperatures (e.g. lipid stores, decreased metabolic rates for some larger body size compared to temperate counterparts, and slower digestion) that may affect toxic responses (DeHoop et al. 2011; Olsen et al. 2011).

Currently there are limited data regarding the toxicity of physically and chemically dispersed oil to Arctic species compared to the substantial body of data for sub-arctic, temperate, and tropical species.  Table 6-2 summarizes data from a subset of the available data reported for median lethal concentrations of oil and dispersed oil mixtures in spiked exposures, as well as 2-methyl-naphthalene and total naphthalene in continuous exposures.  For each of the different test treatments, there is a high degree of overlap among the different test species and phyla.  This is consistent with the comparisons among larger data sets that have included both Arctic and temperate species (Figures 6-2 and 6-3).

While regionally specific toxicity data are ideal, the practical challenges of testing Arctic species in standard laboratory tests provide incentive to determine if previous research on a broader array of species from different regions can be extrapolated to Arctic species. Some studies have attempted to answer the question whether the Arctic species are more or less sensitive to oil components than the temperate species at the individual level: using bivalves (Baussant et al. 2009), fish (Skadsheim et al. 2009, Christiansen et al. 1996; Ingebritsen et al. 2000), decapods (Perkins et al. 2005), copepods (Hansen et al. 2011; Jensen et al. 2008; Jensen and Carroll 2010; Hjorth and Nielsen 2011; Grenvald et al. 2013; Word and Gardiner, in preparation). Interestingly, Hansen et al. (2011) found that lipid rich individuals C. glacialis survived longer than the lipid-poor C. finmarchicus individuals exposed to crude oil highlighting that lipids which are crucial to most arctic species can influence the results of toxicity tests. Furthermore, Christiansen et al. (1996) and Ingebritsen et al. (2000) revealed that the cold water adapted kidney of the polar cod (the kidney is aglomerular to retain the antifreeze biomolecule in the blood) may prevent excretion of contaminants via the urine. Although the specific physiological cold water adaptation may render the B. saida more susceptible to contaminants, there is no evidence that this fish is more vulnerable than other fish.

Olsen et al. (2011) ran acute toxicity tests with Arctic and temperate species with a single PAH (2-methyl naphthalene) and developed species sensitivity distributions (SSD) which could be compared statistically. This study indicated that median estimates for the hazardous concentrations affecting 5 and 50 percent of the species (HC5 and HC50) based on both the NOEC and LC50 estimates were less than a factor 2 higher for temperate species than for Arctic species and were not statistically different. The authors concluded that there was no regional differences in tolerance to 2-methyl naphthalene either at the species level (LC50 and NOEC) or at the aggregated species level (HC5 and HC50).  Further they conclude that the values of survival metrics established for temperate regions are transferrable to the Arctic.  These findings are supported by Word and Gardiner (in prep) who compare the relative sensitivity of Arctic and non-Arctic species using measured and literature data. They come to a similar conclusion for parent naphthalene, WAF, and CEWAF in spiked exposures (see Figures 6-2 and 6-3).

Table 6-2.  Summary of Median Lethal Concentrations for Four Major Phyla

 ArthropodaEchinodermataMolluscaFish

WAF LC50 (mg/L TPH) in Spiked Exposures1

Mean LC50

SD

Range (n)

4.6

2.9

0.6 - 9.6 (12)

ND

ND

ND

ND

ND

ND

2.4

1.2

0.7 - 4.0 (6)

Dispersed WAF LC50 (mg/L TPH) in Spiked Exposures2

Mean LC50

SD

Range (n)

32

27

11 – 62 (3)

ND

ND

ND

ND

ND

ND

42

19

28 – 55 (2)

2-Methynaphthalene LC50 (mg/L) in Continuous Exposures3

Mean LC50

SD

Range (n)

2.4

1.7

1.3 - 5.4 (6)

0.7

-

- (1)

2.3

2.0

0.3 - 5.0 (4)

0.8

-

- (2)

Mean LC50

SD

Range (n)

2.4

1.3

1.2 - 5.3 (8)

ND

ND

ND

ND

ND

ND

1.4

0.3

1.0 - 1.7 (9)

ND: No data

1: WAF data include C. finmarchicus, C. glacialis, C. bairdi, M. oculata, E suckeyi, B. saida, Myoxocephalus sp, S. malma.

2: Dispersed WAF data include C. glacialis, C. bairdi, B. saida, Myoxocephalus sp.

3: 2-Methylnaphthalene data include C. finmarchicus, C. glacialis, B. saida, Gammarus sp., P borealis, S. boreas, A. nugax, N. gracile, C. islandica, A. tessulata, L. littorea, M. helicina, S droebachiensis.

4: Total naphthalene data include A nugax, B. nanseni, B. saida, Gammarus sp., O. hexacornis, C. finmarchicus.

Figure 6-2. Species-sensitivity distribution curves comparing the relative sensitivity for Arctic (solid line) and temperate (dashed line) species to for 2-methylnaphthalene based on (A) LC50s, and (B) No effect concentrations.
Figure 6-2. Species-sensitivity distribution curves comparing the relative sensitivity for Arctic (solid line) and temperate (dashed line) species to for 2-methylnaphthalene based on (A) LC50s, and (B) No effect concentrations.

[Thin dashed lines represent 95% confidence intervals (Olsen et al. 2011)].

Figure 6-3. The relative sensitivity of arctic and temperate species to naphthalene, WAF, and CEWAF exposures (Word and Gardiner, in prep)
Figure 6-3. The relative sensitivity of arctic and temperate species to naphthalene, WAF, and CEWAF exposures (Word and Gardiner, in prep)

A third study by DeHoop et al. (2011) also used the SSD approach to evaluate the relative sensitivity of naphthalene, 2-methylnaphthalene, and crude oil across data sets from a variety of literature sources.  The results support the general conclusion of Olsen et al. (2011) that values of survival metrics for temperate regions are transferrable to the Arctic for the chemical 2-methyl naphthalene, naphthalene, and physically and chemically dispersed crude oil as long as extrapolation techniques are properly applied and uncertainties are taken into consideration.

In assignment of risk for regulatory purposes, it is assumed that if species at the individual level are protected then higher levels of biological organizations (e.g. populations, communities, and ecosystems) will also be protected.  This implies that if there are similarities in the sensitivity of temperate and Arctic species, then they will be similarly sensitive at the population or community level.  However, a mesocosm study at the community level (Olsen 2007a) demonstrated that Arctic benthic communities may respond to oil-related compounds differently than temperate communities. Indeed, by measuring respiration, community structure and sediment characteristics as response variables of sediment community exposed to crude oil no effects could be observed with the temperate benthic communities as opposed to the Arctic sediment community. The authors highlighted that the effects were mainly due to differences in community structure and functioning between the two ecosystems.  Although the potential influences of a non-pristine temperate community could not be ruled out.  This study highlights the need to properly extrapolate from individual test models to population level for a proper environmental consequences analysis (ECA).

There has been a considerable effort in the past five to ten years to better understand the sensitivity of Arctic species to oil and treated oil, with nearly half of the identified studies published in the last five years.  Most of the recent studies provide relatively good information for risk assessment and biomonitoring, especially for the pelagic ecosystem and for assessing the toxicity of chemically or physically dispersed oil.

The copepod dataset is among the most complete for Arctic species, particularly for acute lethality.  Current data exist for oil-water dispersions, dispersed oil, ISB residuals, and single compounds.  There are similar data evaluating oil and chemically dispersed oil for pelagic fish, particularly Arctic cod (B. saida) and Atlantic cod (G. morhua).  For acute-lethal endpoints, data sets could be expanded with qualified data from boreal or temperate data sets.  There is also an emerging dataset regarding sublethal endpoints for copepods and cod; however, this dataset is still limited.  Other pelagic VECs that have little or no data (e.g. capelin, hyperiid amphipods, C. hyperboreus, and Arctogadus glacialis) as well as other types of petroleum hydrocarbons (e.g. ISB residuals) should be a focus of future efforts.   

Ice communities are a unique component of the Arctic and their ecology has been a focus over the past decade.  Recent toxicity studies have dramatically increased our understanding of effects of oil and dispersed oil on ice amphipods, an important component of the ice community.  These acute toxicity data are suitable for use in evaluating the effects of oil in ice.  However, additional work is necessary to better understand effects and exposures of oil and treated oil in and near sea ice and sublethal effects. 

There is a substantial amount of data on the effects of oil and treated oil on temperate benthic communities; however few data exist for Arctic species.  Based on the study by Olsen et al. (2011) some data on temperate species may be applicable for Arctic species however, the data supporting this conclusion are limited.  Further evaluations with other single PAHs, oil and treated oil for targeted species designed to strengthen this conclusion will dramatically improve the available data for benthic intertidal and subtidal communities.

6.4 Future Research Considerations

The review of the ecotoxicology of oil and treated oil in the Arctic described by the authors in this section led to suggestions of further research which can reduce remaining uncertainties.  The more generic suggestions compiled from this review are summarized below while recommendations that are important for improving Arctic NEBA are listed separately.

  1. Bunker oil and other fuel oils:  As the ship traffic increases in the Arctic, fuel oils are the most likely petroleum compound to be released.  The consequences of accidental discharges of bunker oil need to be better understood, particularly considering that currently limited technical solutions exist to recover bunker oil mixed with sea ice (communication by the coastal administration of Norway).
  2. Chemical dispersants: a limited number of chemical dispersants have been tested.  Indeed only Corexit® 9500 (NALCO 2013) and Dasic Slickgone (DASIC 2013) could be found in the recent literature while many other dispersants exist on the market and could potentially be used during an oil spill response. In order to select the best response option, it is important to understand the performance characteristics of chemical dispersants in low temperature applications.
  3. Ecotoxicological studies:
    1. To complement the existing lethal toxicity dataset, we recommend conducting toxicity studies to better characterize the effects of dispersed oil, including both acute and chronic exposures.  Where appropriate, longer-term studies should be conducted with treated and untreated oil focusing on ecosystem/species recovery.  Such long-term toxicity data can be used in combination with experiences from cold-water spills (e.g. Exxon Valdez) to identify critical factor for effects/recovery (See Ecosystem Recovery Section).
    2. Single PAHs:  Toxicity data with single compounds provide an important tool for comparing the toxicity of different OSRs, as well as providing model input.
    3. Herding chemicals:  Chemical herding agents are a promising enhancement for spill response amendment under certain conditions.  Examine the effects of herder chemicals that concentrate surface oil and are relatively unknown to date.
    4. Photooxidation: sea ice species living at the surface of the sea can be exposed to oil whose toxicity has been enhanced following exposure to UV, this could be addressed as part of a larger effort to resolve this issue for any marine spill (this issue is not unique to the arctic, and we do not have sufficient resolution for arctic or temperate exposures).
    5. Examine modes of action:  Knowledge of the quantitative effect of oil droplets and soluble components of oil on biota is also of great importance.  Examinations of toxicity modes of action need to address chemical uptake, body burdens and the narcosis related responses but also need to address the potential for the direct effects of fouling and implications of impacts of soluble or volatile compounds on epithelial tissues, especially those related to respiration.  
  4. Community level assessments:  As the need for an ecosystem management approach has been growing over the last few years, there is a need to move from single species toxicity tests towards community studies to produce data with the highest ecological possible relevance with a specific focus on the structural and functional changes. One methodological approach as discussed earlier would be the use of in situ and other mesocosms and also to work with early life stage impacts that can be fed into population and possibly ecosystem models.

Photo 6-14: Pelagic Hyperiid amphipod Themisto sp (Bjørn Hansen).
Photo 6-14: Pelagic Hyperiid amphipod Themisto sp (Bjørn Hansen).

6.4.1 Priority Recommendations to Enhance NEBA Applications in the Arctic

The recommendations presented below indicate where increased knowledge of ecotoxicology would result in reducing existing uncertainties in NEBA assessments.  No prioritization has been made to the list; for some of the recommendations, surrogate data may be already available.

  1. Toxicity studies of weathered and unweathered OSR residuals.  Given the remoteness of the Arctic and the inclement weather, it is possible that responders will be treating weathered oil.  The toxicity of oil weathered under Arctic winter conditions is less well understood.  Priorities for toxicity data production should be defined based on the need for input into existing impact assessment tools which will be used by the stakeholders.
    1. ECs coupled with VEC test species:  In non-pelagic environments (surface microlayer, intertidal, shallow subtidal, deep water species, annual and multi-year ice); Pelagic: weathered oil
      • Create exposures to constant droplet size and concentration of OSR residuals; determine biological responses (survival, growth, lipid content, possible fecundity measure); determine impacts due to dissolved fraction and risk to pelagic organisms
      • Conduct in-situ assessment of rate of benthic recovery from controlled oil exposure
    2. Endpoints:  fouling, mortality, uptake and narcosis, histopathology associated with measured chemical concentration in exposure media.
    3. Exposures:  spiked, low constant dose, multiple spiked to simulate re-oiling (e.g. Droplet generator studies).  Although LC50 are used for regulatory risk assessment, the ecotoxicological information that they provide is often not environmentally relevant since the amount of oil tested far exceeds potential environmental concentrations.
    4. Evaluate SSD curves   Evaluate SSD curves for weathered and fresh oils for various modes of action.
    5. Evaluate consequences of OSR strategies.  To date there are very few toxicity studies that have been conducted with certain types of OSRs, most notably ISB residues and OMAs.  A full evaluation of OSR alternatives will require additional toxicity data for targeted VECs, defined by the anticipated fate of the treated oil.Additionally, very few chemical dispersants have been fully evaluated for acute and chronic effects.
  2. Augmentation of relative sensitivity dataset. The relative sensitivity assessment of arctic versus temperate species conducted by several authors indicates that the toxicity database for Arctic studies could be augmented with data from temperate species.  Further data are required for sublethal and chronic effects of oil and treated oil before the data can be extrapolated from temperate studies.  Data on benthic species and subtidal communities are sparse.  Further information is also needed for other Arctic VECs such as nearshore fish (e.g. capelin), sea ice, microlayer, and higher trophic groups (such as birds and mammals).
  3. Improve model relationships
    1. Impacts due to volatile fractions of oil.  Volatilization of hydrocarbons exposes marine mammals and seabirds to respiratory impacts (greatest with natural attenuation, herder application, mechanical recovery, ISB); exposure needs to model concentrations near air/water interface.
    2. Encompass all modes of toxic action

6.5 Further Information

Authors Dr. Lionel Camus [Akvaplan Niva], Dr. Jack Q. Word and William W Gardiner [ENVIRON], Dr. Bjorn H Hansen [SINTEF], Steinar Sanni [IRIS], Dr. Oleg Titov [PINRO], Dr. James Clark [HDR/EM&A] 

References

Aarset AV, Zachariassen KE

1983

Synergistic effects of an oil dispersant and low temperature on the freezing tolerance and solute concentrations of the blue mussel (Mytilus edulis L.)

Polar Research 1 (3):223-230

Aarset AV, Zachariassen KE

1983

Synergistic effects of an oil dispersant and low temperature on the freezing tolerance and solute concentrations of the blue mussel (Mytilus edulis L.)

Polar Research 1 (3):223-230

Aas E, Liewenborg B, Grøsvik BE, Camus L, Jonsson G, Børseth JF, Balk L

2004

DNA adduct levels in fish from pristine areas are not detectable or low, when analysed by the nuclease P1 version of the 32P-postlabelling technique

Biomarkers 8(6):445-460

Anderson JW, JM Neff, BA Cox, HE Tatem, GM Hightower

1974

Characteristics of dispersions and water soluble extracts of crude and refined oils and their toxicity to estuarine crustaceans and fish.

Mar Biol 27:75-88

Aunaas T, A Olsen, and KE Zachariassen

1991

The effects of oil and oil dispersants on the amphipod Gammarus oceanicus from Arctic waters

Polar Research  10(2):619-630

Bach L,  Forbes VE, Dahllöf  I

2009

The amphipod Orchomenella pinguis – A potential bioindicator for contamination in the Arctic

Mar Poll Bull  58(11):1664-1670

Bach L, I Dahllof

2012

Local contamination in relation to population genetic diversity and resilience of an arctic marine amphipod.

Aquat Toxicol 114-115: 58-66

Barron MG and L Ka’Kaihue

2003

Critical evaluation of CROSERF test methods for oil dispersant toxicity testing under subarctic conditions (Journal)

Mar Poll Bull   46:1191-1199

Barron MG, MG Carls, JW Short, SD Rice

2003

Photoenhanced toxicity of aqueous phase and chemically dispersed weathered Alaska North Slope crude oil to Pacific herring eggs and larvae.

Environ Toxicol Chem 22(3):650-60

Barry TW

1970

Likely effects of oil in the Canadian Arctic

Mar Poll Bull1 (5):73-74

Baussant T, Bechmann RK,  Taban IC, Larsen BK, Tandberg AH,  Bjørnstad A, Torgrimsen S, Nævdal A, Øysæd KB, Jonsson G,  Sanni S

2009

Enzymatic and cellular responses in relation to body burden of PAHs in bivalve molluscs: A case study with chronic levels of North Sea and Barents Sea dispersed oil

Mar Poll Bull, 58(12):1796-1807

Bechmann KR, BK Larsen, IC Taban, LI Hellgren, P Moller, S Sanni

2010

Chronic exposure of adults and embryos of Pandalus borealis to oil causes PAH accumulation, initiation of biomarker responses and an increase in larval mortality.

Mar Poll Bull 60(11):2087-2098

Busdosh M, Atlas RM

1977

Toxicity of oil slicks to Arctic amphipods

Arctic 30:85-92

Bushdosh, 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

Mar Environ Res 5 (3):167-180

Camus L Nahrgang J, Olsen GH

2010

Risk assessment and biomonitoring tools in the Arctic

Proceedings, 33rd AMOP Technical Seminar on Environmental Contamination and Response, Halifax, Canada. Volume 2, 1253-1266

Camus L, BE Grøsvik, JF Børseth, MB Jones, MH Depledge

2000

Stability of lysosomal and cell membranes in haemocytes of the common mussel (Mytilus edulis): effect of low temperatures.

Mar Environ Res 50:325-329 Environ Res 50:325-329 Environ Res 50:325-329 Environ Res 50:325-329

Camus L, Børseth JF, Jones MB, Regoli F, Depledge M

2002

Heart rate and total oxyradical scavenging capacity of the Arctic spider crab, Hyas araneus, following exposure to crude oil via sediment and injection

Aquat Toxicol 61:1-13

Camus L, GH Olsen

2008

Embryo aberrations in sea ice amphipod Gammarus wilkitzkii exposed to water soluble fraction of oil.

Mar Environ Res   66:221-222

Camus L, GH Olsen

2008

Embryo aberrations in the Arctic sea ice amphipod Gammarus wilkitzkii exposed to the soluble fraction of oil.

Mar Environ Res 66:221-222

Camus L, Gulliksen B

2005

Antioxidant defense properties of Arctic amphipods: comparison between deep-, sublittoral and surface-water species

Mar Biol 146:355-362

Camus L, Gulliksen B, Depledge MH, Jones MB

2005

Polar bivalves are characterized by high antioxidant defences

Polar Research 24(1-2 ):111-118

Camus L, Jones MB, Børseth JF, Grøsvik BE, Regoli F, Depledge MH

2002

Total oxyradical scavenging capacity and cell membrane stability of haemocytes of the Arctic scallop, Chlamys islandicus, following benzo(a)pyrene exposure

Mar Environ Res  54:425-430

Camus, L., Birkely S. R., Jones M. B., Børseth JF, Grøsvik BE, Gulliksen B, Lønne OJ, Regoli F, Depledge MH

2003

Biomarker responses and PAH uptake in Mya truncata following exposure to oil-contaminated sediment in an Arctic fjord (Svalbard)

Sci  Total Environment  308 (1-3):221-234

Carls MG, Korn S

1985

Sensitivity of Arctic marine amphipods and fish to petroleum hydrocarbons

Can Tech Rep Fish Aquat Sci 1368: 11-26

Carrasco Navarro V

2013

PAHs: Comparative biotransformation and trophic transfer of their metabolites in the aquatic environment.

Dissertation to University of Eastern Finland No. 106.  112 pp.

Carroll J, Smit M

2011

An integrated modeling framework for decision support in ecosystem management: case study Lofoten/Barents Sea

Society of Petroleum Engineer 140431

Chapman P, McPherson C

1993

Comparative zinc and lead toxicity tests with Arctic marine invertebrates and implications for toxicant discharges

Polar Record  29(168):45–54

Christiansen JS

2000

Sex differences in ionoregulatory responses to dietraty oil exposure in polar cod

Jnl Fish Biol 57:167-170

Christiansen JS, Dalmo RA, Ingebrigsten K

1996

Xenobiotic excretion in fish with aglomerular kidneys

Mar Ecol Prog Ser 136:303-304

Christiansen JS, George SG.

1995

Contaminatiuon of food by crude oil affects food selection and growth performance, but not appetite, in an Arctic fish, the polare cod (Boreogadus saida)

Polar Biology15:277-281

Christiansen JS, Karamushko LI, Nahrgang J

2010

Sub-lethal levels of waterborne petroleum may depress routine metabolism in polar cod Boreogadus saida (Lepechin, 1774)

Polar Biology 33:1049–1055

Christiansen JS, LI Karamushko, J Nahrgang

2009

Sub-lethal effects of waterborne petroleum may depress routine metabolism in polar cod Boreogadus saida

Polar Biol 33(8):1049-1055

Clark JR, GE Bragin, EJ Febbo,  DJ Letinski

2001

Toxicity of physically and chemically dispersed oils under continuous and environmentally realistic exposure conditions: Applicability to dispersant use decisions in spill response planning.

Proceedings, 2001 International Oil Spill Conference; American Petroleum Institute, Washington DC.  pp 1249-1255

DASIC

2013

Slickgone

http://www.dasicinter.com/oil-dispersant/products/range/oil-dispersant-product-range/show/slickgone-ew

De Laender F, Olsen GH,Frost T, Grøsvik BE, Hansen BH, Jan Hendriks A, Janssen CR, Klok C, Nordtug T, Smit M, X Carroll J, Camus L

2010

Ecotoxicological mechanisms and models in an impact analysis tool for oil spills

Jnl Toxicol  Environ Health Part A. 74(7):605-619

Duesterloh W, JW Short, MG Barron

2002

Photoenhanced toxicity of weathered Alaska North Slope crude oil to the calanoid copepods Calanus marshallae and Metridia okhotensis.

Environ Sci Tech 36(18):3953-3959

Engelsen O, Hop H,Hegseth EN, Hansen E, Falk-Petersen S

2003

Deriving phytoplankton biomass in the Marginal Ice Zone from satellite observable parameters

Intnl Jnl Remote Sensing 25:1453-1457

Faksness LG, BH Hansen, D Altin, PJ Brandvik

2012

Chemical composition and acute toxicity in the water after in situ burning--a laboratory experiment.

Mar Pollut Bull 64(1):49-55

Falk-Petersen S, Hop H, Budgell WP, Hegseth EN, Korsnes R, Løyning TB, Orbæk JB, Kawamura T, Shirasawa K

2000

Physical and ecological processes in the marginal ice zone of the northern Barents Sea during the summer melt period

Jnl Mar Syst 27(1-3):131-159

Falk-Petersen S, Sargent JR, Hopkins CCE

1990

Trophic relationships in the pelagic arctic food web

In: Barnes, M. and Gibson, R.N. (eds.), Trophic relationships in the marine environment. Scotl. Univ Press, Aberdeen, pp 315-333.

Frantzen M, IB Falk-Pettersen, J Nahrgang, T Smith, GH Olsen, TA Hagstad, L Camus

2012

Toxicity of crude oil and pyrene to the embryos of beach spawning capelin (Mallotus villosus).

Aquat Toxicol 108:42-52

Frantzen M, J Nahrgang, P Geraudie, W Locke, W Wmbrose, F Regoli, L Camus

2013 in prep

Biological effects of mechanically and chemically dispersed oil on the Icelandic scallop (Chlamys islandica)

In preparation

Fuller C, J Bonner, C Page, A Ernest, T McDonald, S McDonald

2004

Comparative toxicity of oil, dispersant, and oil plus dispersant to several marine species.

Environ Toxicol Chem 23(12):2941-2949

Gallaway BJ, WJ Gazey, JG