Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

The Natural History of an Arctic Oil Field: Development and the Biota
The Natural History of an Arctic Oil Field: Development and the Biota
The Natural History of an Arctic Oil Field: Development and the Biota
Ebook856 pages8 hours

The Natural History of an Arctic Oil Field: Development and the Biota

Rating: 0 out of 5 stars

()

Read preview

About this ebook

In spite of the harsh conditions that characterize the Arctic, it is a surprisingly fragile ecosystem. The exploration for oil in the Arctic over the past 30 years has had profound effects on the plants and animals that inhabit this frozen clime. The Natural History of an Arctic Oil Field synthesizes decades of research on these myriad impacts. Specialists with years of field experience have contributed to this volume to create the first widely available synopsis of the ecology and wildlife biology of animals and plants living in close association with an actively producing oil field.
  • First widely available synthesis of arctic oil field ecology and wildlife biology
  • Concise yet readable treatment of a diverse polar ecosystem
  • Useful for land managers, policy makers as well as ecologists, and population biologists
  • Chapters authored by recognized authorities and contributions are peer-reviewed for accuracy and scientific rigor
  • Illustrations attractively designed to enhance comprehension
LanguageEnglish
Release dateJun 9, 2000
ISBN9780080512419
The Natural History of an Arctic Oil Field: Development and the Biota

Related to The Natural History of an Arctic Oil Field

Related ebooks

Biology For You

View More

Related articles

Related categories

Reviews for The Natural History of an Arctic Oil Field

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    The Natural History of an Arctic Oil Field - Joe C. Truett

    Truett

    Preface

    Oil development commenced 30 years ago on the coastal plain of arctic Alaska. Since that time development has expanded, and oil fields now encompass a substantial portion of the central coastal plain and coastal marine environment between the Colville and Sagavanirktok rivers. Environmental research, stimulated largely by the desire to minimize impacts on fish and wildlife populations and habitats, escalated with development.

    Over the years several entities have sponsored portions of this research. The federal government contributed funding through agencies such as the National Oceanic and Atmospheric Administration, the Bureau of Land Management, and the Minerals Management Service. The State of Alaska conducted research through existing agencies such as the Alaska Department of Fish and Game. The oil companies have contributed most of the research funding; BP Exploration (Alaska) Inc. (BPX), its predecessors Standard Alaska Production Company and SOHIO, and ARCO Alaska have been the major oil-field operators and the major industry sponsors of research. EXXON Company sponsored some studies in the early 1980s. Researchers have produced literally hundreds of reports.

    In 1996 BPX’s Environmental and Regulatory Affairs Department determined to sponsor a technical book as a means for more widely disseminating results of this research. They commissioned contractors to edit the volume and agreed to provide honoraria to authors. They, in collaboration with the senior editor, developed a list of fish and wildlife species, groups, and habitats they wished to see addressed, and they selected senior authors for these topics.

    BPX set the book’s focus and authorship and ultimately reviewed all manuscripts in near-final form but otherwise did not monitor the book’s preparation. Manuscripts were written and reviewed by outside experts independent of BPX’s purview. The senior editor worked with the authors to promote uniformity in style and format among manuscripts, objectivity and brevity in presentation, logic in interpretation, and acceptability in grammar and syntax. He obtained one to three peer reviews of each manuscript, reviewers being selected on the basis of their knowledge of the discipline and the editor’s judgment of their ability to provide a thorough and objective review. Three chapters—Introduction, North Slope Oil Field Development, and Synthesis—are not technical presentations of research and were not peer reviewed.

    The senior editor required the author or authors of each technical manuscript to address reviewers’ comments by revising as appropriate. He then edited each revised manuscript for responsiveness to review and for final details of grammar, readability, and format before sending it to the publisher. Each senior author reviewed page proofs of his manuscript, which he received directly from the publisher. The junior editor helped in the final stages of editing and in final preparations for publication.

    We thank the following for their technical reviews of chapters: R. M. Anthony, W. Ballard, S. Boyd, P. Clarkson, W. B. Collins, E. Cooch, P. Craig, M. Cronin, J. Davis, L. Dickson, R. Drewein, S. Fancy, H. M. Feder, E. H. Follmann, W. B. Griffiths, J. E. Hemming, L. Jacobs, M. S. Lindberg, N. Lunn, P. Martin, K. Moiteret, S. Murphy, D. Norton, A. Ott, R. E. Schweinsburg, R. Suydam, E. J. Taylor, and K. Whitten.

    This book could not have taken shape without the help of numerous organizations and individuals. BPX provided funding support for the writing, preparation of final illustrations, and editing. Chris Herlugson and Ray Jakubczak of BPX’s Environmental and Regulatory Affairs Department deserve much credit for supporting this project. In many cases authors supplemented their BPX honoraria with support from their employers or with time or money of their own. James Lukin collated and prepared final versions of artwork. Anne Brown coordinated activities between BPX and the editors and authors. Michelle Gilders read the entire manuscript and offered many useful suggestions. Judy Truett supported the editorial process in many ways.

    BPX sponsorship of this book expresses its support for the editorial and peer-review process that promotes objectivity in science. This sponsorship does not imply agreement with views expressed by the authors.

    Joe C. Truett and Stephen R. Johnson

    PART I

    Introduction to Arctic Ecosystems

    CHAPTER 1

    Introduction

    Joe C. Truett,     Truett Research, Glenwood, New Mexico

    Introduction

    Physical Environment

    Wildlife, Fish, and Habitats

    Anthropogenic Changes

    The Investigative Focus

    References

    INTRODUCTION

    In June 1968, Atlantic Richfield Company announced the discovery of a major oil accumulation near Prudhoe Bay, Alaska. Close upon the heels of this news, the U.S. government passed the National Environmental Policy Act (NEPA) of 1969, requiring full disclosure of the environmental costs of major development ventures. This coincidence set the stage for an experiment that in time not only would measure responses of biota to arctic oil development, but in so doing would test the ability of intensive scientific investigation to respond to the high hopes of NEPA This book presents key findings of that experiment.

    Arctic scientists had long anticipated the potential for the scale of development set in motion by the Prudhoe Bay discovery, and two decades earlier they had begun to lay the groundwork for measuring impacts of human-induced disturbance. In 1947, the Naval Arctic Research Laboratory (NARL) took shape at Point Barrow, the farthest north landscape in Alaska, and scientists based at NARL initiated studies on the effects of disturbance on soils and nutrient cycling (Shaver, 1996). In 1958 the U.S. Atomic Energy Commission authorized environmental studies near Cape Thompson on the northwestern coast of Alaska in anticipation of an experimental harbor excavation by nuclear blast (which never materialized) (Wilimovsky and Wolfe, 1966). In the early 1970s the National Science Foundation initiated two arctic Alaska programs: (1) the International Biological Program (IBP) and its Coastal Tundra Biome Studies at Barrow (Brown et al., 1980) and (2) the Research on Arctic Tundra (RATE) program inland from Barrow at Atkasook (Batzli and Brown 1976).

    Preparations for the development of Prudhoe Bay oil elevated substantially the level of inquiry into the potential impacts of arctic disturbance. In 1969, the year following the oil strike, applications filed by industry to develop a trans-Alaska pipeline system to export the oil southward spawned research along the proposed pipeline corridor (Alexander and Van Cleve, 1983). In 1975, the U.S. government initiated the oil-related Alaska marine studies program known as the Outer Continental Shelf Environmental Assessment Program (OCSEAP), billed by its director as the largest environmental program in the history of our nation and probably of the world (Engelmann, 1976). Between 1976 and 1979, the U.S. Geological Survey conducted environmental studies on the vast National Petroleum Reserve in Alaska (NPR-A) west of the Prudhoe Bay region in anticipation of potential oil development there (USGS, 1979). In 1983–1984 the U.S. Department of Energy initiated new studies to help assess impacts of arctic energy development; the program, called Response, Resistance, Resilience, and Recovery of Arctic Ecosystems to Disturbance (R4D), centered on a site near Toolik Lake in the Brooks Range foothills 250 km south of Prudhoe Bay (Reynolds and Tenhunen, 1996). During 1980–1985, the U.S. Fish and Wildlife Service conducted a baseline study to determine the size and diversity of fish and wildlife populations in a coastal plain portion (the 1002 area) of the Arctic National Wildlife Refuge judged vulnerable to petroleum development (Garner and Reynolds, 1986). In 1987 scientists began to investigate the potential impacts of petroleum development in the 1002 area on key fish and wildlife species and their habitats (McCabe, 1994).

    The greatest intensity of investigation took place in the oil fields themselves. Studies here, funded largely by industry and monitored by agencies, distinguished themselves by one major difference from the other research programs—they overlapped temporally and spatially with development. Many researchers built on this advantage to measure responses of biota to development. The budgets for oil field environmental studies, which focused strongly on wildlife and fish, have been conservatively estimated to have totaled more than $4 million per year since the early 1980s (Maki, 1992).

    This book assembles findings of studies sited in and near arctic Alaska oil fields. The focus is on fish, wildlife, habitats, and communities that are important from public and agency perspectives; information from outside the oil-field region is incorporated as necessary. The following paragraphs provide an introduction to the physical and biological characteristics of the oil field region, the general nature of disturbance, and the investigative approach of the studies.

    PHYSICAL ENVIRONMENT

    Producing oil fields in arctic Alaska lie between the Colville and Sagavanirktok rivers on the arctic coastal plain and extend into nearshore waters of the Beaufort Sea (see chapter by Gilders and Cronin, this volume). Land and water intermix to create a complex wetland in much of the onshore environment. Climate exerts a major influence on the land and water, not only severely constraining the landscape’s ability to support wildlife and fish but also imposing landscape changes that often mimic the changes wrought by development.

    Climatological extremes prevail. The cold, dark winter months, with January temperatures commonly averaging between −20 and −90°C, contrast sharply with the continuous daylight of summer, when minimum and maximum July temperatures reach 1 and 8°C, respectively. The scanty precipitation averages 13–18 cm annually with most coming as snow (Selkregg, 1975, p. 18). Winds blow mainly from the east and secondarily from the west (Dygas, 1975); they are persistent and strong (Selkregg, 1975, p. 19).

    The temperature regime strongly affects habitat qualities of the landscape. Surface soils remain frozen and snow-covered for 8–9 months each year (Hobbie, 1984; Walker, 1985). Snowmelt and the thawing of surface soil and water commence in May a few to several tens of kilometers inland; these phenomena are delayed for 2 weeks or so at the coast because of the cooling effect of the nearby frozen ocean. In the oil-field area, soil thaw begins in June, reaches 50–100 cm deep by late summer, and reverses itself such that soils begin to refreeze during September (Hobbie, 1984; Walker, 1985); biological activity restricts itself to the thaw layer. Below this, permafrost (permanently frozen ground) extends to depths of 660 m or more (Walker, 1985). The shortness of the period of thaw in concert with the blockage of subsurface drainage by permafrost curtails water loss by evaporation and percolation so that, despite the meager precipitation, much of the tundra in summer resembles a marshy grassland (Hobbie, 1984).

    The action of ice strongly molds emergent landforms. Tundra landscapes show naturally patterned ground formed by the wedging effect of ice alternately freezing and thawing at the margins of soil polygons 5–10 m or more in diameter (Walker, 1985). Surface disturbances caused by the freeze-thaw cycle and other forces often reduce locally the albedo, or surface reflectance, resulting in thermokarst, which is surface subsidence caused by thawing of ice-rich soils. Occasional pingos (ice-cored hills) rise here and there as much as 15 m higher than the surrounding tundra (Walker, 1985; Walker and Walker, 1991).

    Polygons, thermokarst depressions, and other cryogenic (ice-generated) features provide microrelief (<1.0 m vertical range) that results in fine-scale horizontal variation in the moisture content of surface soils. This in turn causes corresponding horizontal heterogeneity in the composition of the vegetation.

    As with terrestrial habitats, aquatic environments yield to ice. Surface waters of ponds, lakes, and the coastal ocean remain frozen for about 9 months each year.

    Typically, ice begins to form on these waters in September or October, reaches a maximum thickness of 2 m or so by March or April, and completely thaws in June or July (Kovacs and Mellor, 1974; Hobbie, 1984). In the nearshore ocean during the open-water period, winds may occasionally bring in ice from the permanent ice pack offshore. The aquatic food base in general—plankton, benthos, and sessile plants—peaks markedly in productivity and availability to consumers during the open-water period (Newbury, 1983; Hobbie, 1984).

    Water, like ice, can cause major landscape changes. River discharge during spring breakup erodes and reworks tremendous amounts of peat, gravel, sand, and finer sediments in floodplains and deltas (Ritchie and Walker, 1974). Movement of water in the nearshore marine zone transports sediments and reworks coastal islands and beaches (Short et al., 1974); storm surges at the coast erode shorelines (Reimnitz and Maurer, 1979). Because the tidal range is small, winds constitute the primary driving force behind water movement in the marine system (Dygas, 1975). The major erosional phenomena caused by water, both in coastal and inland localities, tend to be episodic.

    Natural agents of landscape disturbance often generate changes ecologically analogous to anthropogenic disturbance, although the scale and conformation of change may be different (Walker and Walker, 1991). Caribou trails, troughs at polygon margins, and vehicle tracks on tundra resemble each other in their tendency to cause thermokarst. Both river flooding and the deposition of fill by industry result in gravel-surfaced features. The thermokarst effects of wave action and heat absorption in natural lakes resemble those where surface water is impounded by industry. Both wind-driven water and construction crews rework coastal landforms and nearshore marine substrates.

    WILDLIFE, FISH, AND HABITATS

    Public and political consensus hold that the primary measure of arctic Alaska’s environmental quality is its ability to support wildlife and fish populations (Garner and Reynolds, 1986, p. 1; Walker et al., 1987a; Maki, 1992; Jorgenson and Joyce, 1994). Thus, the relationships of animals to habitats, and the alteration of habitats by people, become of central concern to management. The research forming the basis for this book originated largely from the perceived need to protect animal populations and maintain habitat quality in the face of increased human activity. Although not intended as impact assessments, the contributions herein provide the kinds of information useful for managing species and habitats influenced by human activities.

    From an ecological viewpoint, mammal, bird, and fish species of arctic Alaska may be categorized broadly by their food-chain and habitat affiliations. Thus classified, major animal groups are (1) herbivorous mammals, (2) carnivorous and omnivorous mammals, (3) herbivorous waterfowl, (4) carnivorous waterfowl (including loons), (5) shorebirds, (6) freshwater fishes, and (7) anadromous fishes.

    Herbivorous mammals of arctic Alaska range in size from lemmings (e.g., Lemmus sibericus, Dicrostonyx spp.) and arctic ground squirrels (Spermophilus parryi) to caribou (Rangifer tarandus), moose (Alces alces), and muskoxen (Ovibos moschatus) (Brooks et al., 1971; Garner and Reynolds, 1986). The smaller species occupy small home ranges year-round; the larger ones range widely, typically occupying different areas and habitats seasonally. In this book we focus on caribou, the only large obligate herbivore common in the oil-field region.

    Carnivorous and omnivorous mammals that characterize coastal arctic Alaska include polar bear (Ursus maritimus), the grizzly subspecies of brown bear (Ursus arctos horribilis), wolf (Canis lupus), and arctic fox (Alopex lagopus) (Brooks et al., 1971; Garner and Reynolds, 1986). Polar bears occupy the coastal zone primarily in winter but also occasionally in summer. Grizzly bears occur in the oil-field area year-round but hibernate in winter. Arctic foxes are the most abundant and ubiquitous of this group. Wolves have been scarce in the oil-field region during the period of development, perhaps because of rabies (Maki, 1992). This book addresses polar bears, grizzly bears, and arctic foxes.

    Herbivorous waterfowl, and indeed nearly all birds found in coastal arctic Alaska, use the region mainly for breeding and brood-rearing and are present only during summer (Derksen et al., 1981; Garner and Reynolds, 1986). Tundra swans (Cygnus columbianus), brant (Branta bernicla), snow geese (Chen caerulescens), Canada geese (Branta canadensis), and white-fronted geese (Anser albifrons) dominate this group of obligate herbivores. Swans, brant, and snow geese are the most conspicuous in the oil-field region and have received the largest amount of study; this book addresses all three of these.

    Large waterbirds relying mainly on animal foods include diving ducks and loons (Derksen et al., 1981; Garner and Reynolds, 1986). The most abundant diver in arctic coastal Alaska is oldsquaw (Clangula hyemalis); less abundant are common eider (Somateria mollissima), spectacled eider (S. fischeri), and king eider (S. spectabilis). Three loons breed regularly in or near the oil fields—Pacific loon (Gavia pacifica), red-throated loon (G. stellata), and yellow-billed loon (G. adamsii) [Derksen et al., 1981; Garner and Reynolds, 1986; Troy Ecological Research Associates (TERA), 1992]. Contributions in this volume address common eider and Pacific loon.

    More than a dozen shorebird species nest on the arctic coastal plain (Derksen et al., 1981). As a group they are the most abundant of the waterbirds. They consume mainly macroinvertebrates in tundra, littoral-zone, and coastal lagoon habitats (Connors and Risebrough, 1979; Johnsgard, 1981). In the oil-field region the most common of these include black-bellied plover (Pluvialis squatarola), American golden-plover (P. dominica), semipalmated sandpiper (Calidris pusilla), pectoral sandpiper (C. melanotos), dunlin (C. alpina), stilt sandpiper (C. himantopus), buff-breasted sandpiper (Tryngites subruficollis), red-necked phalarope (Phalaropus lobatus), and red phalarope (P. fulicaria) (TERA, 1990, 1992). These species arrive in the oil fields in early summer and nest in tundra habitats; adults and young typically depart the oil fields by late July or early August. This book’s contribution on shorebirds addresses these numerically dominant species.

    Among the fishes of arctic Alaska, a habitat gradation exists between species that occur exclusively in freshwater streams and lakes (freshwater species) and those that also range seasonally into the coastal waters of the Beaufort Sea (amphidromous and anadromous species). Arctic grayling (Thymallus arcticus) and ninespine stickleback (Pungitius pungitius) typify the region’s freshwater fauna. Broad whitefish (Coregonus nasus), least cisco (C. sardinella), and Dolly Varden (Salvelinus malma) are largely freshwater species with some amphidromous populations that range seasonally into brackish or marine waters. The anadromous humpback whitefish (Coregonus pidschian) and arctic cisco (C. autumnalis) spend more of their lives outside freshwater habitats; the latter makes up the bulk of the commercial fishery of arctic Alaska (Kogl and Schell, 1975; Morrow, 1980; Craig, 1989). This book includes a contribution on freshwater fishes and one on amphidromous and anadromous fishes.

    Terrestrial vegetation and aquatic invertebrates comprise food-chain entities particularly important to the wildlife and fish populations addressed herein. With the exception of vertebrates eaten by other vertebrates, they comprise nearly the entire food base. A thorough analysis of the trophic importance of invertebrates in the nearshore marine environment has been published (Craig et al., 1984). Similar analyses of terrestrial vegetation and freshwater invertebrates have not been published, and this book includes a chapter on each.

    Within the rather uniform benthic environments of arctic Alaska bays and lagoons there exists, within the oil-field region, an anomalous arctic ecosystem—a kelp community associated with clusters of boulders (Dunton et al., 1982). The importance of this boulder patch community in supporting regional fish and wildlife populations is perhaps small, but the community itself has attracted much public and scientific interest because of its uniqueness. This book includes a contribution on its ecology.

    ANTHROPOGENIC CHANGES

    For centuries, native peoples in pursuit of subsistence hunted and fished the coastal regions of arctic Alaska (Jamison, 1978). Whalers of European descent, arriving in the late 1800s (Milan, 1978), augmented the ability of people to affect wildlife and fish when they brought firearms and other accouterments of the industrial age. Shortly thereafter, nonnative Americans began to settle in the region, gradually exacerbating subsistence pressures on the wildlife and fishery resources (Brower, 1942). The beginning of oil development in the 1970s brought a final quantum jump in the power of people to influence not only the animals but also their habitats.

    Changes introduced by the oil industry to arctic coastal Alaska fit within five major categories of action that could influence wildlife and fish populations. Four alter habitats: (1) addition of gravel fill, (2) disruption of tundra surfaces, (3) creation of impoundments, and (4) introduction of elevated structures. The fifth is the increase in human numbers and activity. Gilders and Cronin (this volume) describe these changes in detail; a summary follows.

    Gravel fill is used to support facilities and vehicular traffic on tundra and in shallow water because it provides a dry, stable surface. Gravel airstrips, camp pads, oil drilling pads, and roadways have been used for more than 30 years (Brown and Grave, 1979; Everett et al., 1985), and gravel fill is now in standard use for roadways on tundra, causeways in shallow water, and facilities support pads (Hanley et al., 1981). Earlier exploratory-well pads on tundra were often 0.6 m or less thick (Hanley et al., 1981, p. 127); most pads and roads in oil fields now are 1.5 m or more thick (Walker et al., 1987a). Gravel causeways and artificial islands in nearshore waters rise several meters above sea level to withstand inundation from storm surge.

    Industry disrupts tundra surfaces by removing, compacting, dusting, oiling, partially covering, or otherwise altering the vegetative cover and the peaty soil. Historically, disruptions were caused by the building of roads of near-surface peat, by moving heavy equipment at sites of construction, or by driving vehicles across the tundra (MacKay, 1970; Brown and Grave, 1979; Everett et al., 1985; Walker et al., 1986); visible signatures of these activities persist. Recently such activities have lessened (Hanley et al., 1981), and most new disruptions are associated with the building and use of snow and ice roads in winter, the dusting of vegetation near heavily traveled roads, the occasional spilling of oil or other contaminants on the tundra surface, and the mining of gravel.

    Except where soil, organic matter, or debris are piled, disrupting the tundra surface almost invariably results in thermokarst (MacKay, 1970). Compaction or removal of the organic mat and its vegetation by such actions as vehicular traffic (Felix and Jorgenson, 1985) or building of peat roads (Walker et al., 1987b) reduces surface insulation (Brown and Grave, 1979, p. 6; Walker et al., 1987b, p. 17). Deposition of road dust on snow hastens snowmelt, and spillage of oil on tundra reduces albedo. All of these promote thermokarst by enhancing heat absorption during summer (McKendrick, 1987; Walker et al., 1987b, pp. 34, 35). Subsequent to thermokarst, water typically accumulates or ponds on the surface (MacKay, 1970; Adam and Hernandez, 1977). Ponding reduces surface thermal insulation and albedo even further, exacerbating subsidence (Walker et al., 1987b). Rates of subsidence eventually lessen, and in the cases examined, a new thermal equilibrium established itself within several years to several decades (MacKay, 1970; Lawson, 1986).

    Impoundments form behind blockages to surface runoff. Within the oil fields, impoundments occur where gravel roads and pads block drainage in previously drained thaw-lake basins or other low-lying areas (Walker et al., 1986, 1987b, p. 30). The total area flooded by impoundments usually peaks as snowmelt ends (i.e., mid-June); most of the flooded areas drain by midsummer after road culverts unclog and surface run-off subsides (Walker et al., 1987b, p. 30).

    Impounded areas typically retain rooted vegetation, but vascular plant cover can disappear when impoundments are deep and persistent (Walker et al., 1987b, p. 30). Though the vegetation in temporarily flooded areas and in shallow, permanently flooded areas typically is not killed, it often is changed. The most noticeable effect is an enhanced greening of the vegetation in summer (Walker et al., 1987b, pp. 22, 30).

    Elevated structures include buildings and horizontal pipes and their supports. All camp facilities, well houses, processing plants, and other buildings sit on gravel pads; most are elevated above the pad surface by stanchions 1–2 m tall that minimize heat conduction to the underlying gravel and permafrost. Stanchions also elevate unburied pipelines 0.5–2.5 m above the tundra surface. Pipelines occur singly or in groups of up to 8 adjacent pipes. Since 1980, all new pipelines have been elevated at least 1.5 m above ground to expedite passage by caribou (Curatolo and Murphy, 1986).

    Human activities peak in time during facilities construction periods and in space along road corridors and at concentrations of infrastructure. Aircraft overflights have escalated with the general increase in oil-related activity. During summer, research biologists and others walk about on the tundra or operate small boats in nearshore waters at some locations. A special type of activity that stimulates response from some animals is the disposal of refuse, which during the past 2 decades has been deposited temporarily in dumpsters near sites of operation and permanently in a landfill in the oil-field region.

    THE INVESTIGATIVE FOCUS

    The species and habitats addressed in this book are those that have attracted primary attention from environmental management agencies and public interest groups. In most cases, identical or similar species and habitats occur widely in arctic regions, reflecting the circumpolar similarities in the physical forces and biological adaptations that characterize arctic ecosystems. Further, most animals we address also have high public profiles elsewhere in the world. In recognition of these interregional similarities, and the likelihood of continuing human presence in the Arctic, the authors have attuned their contributions to potentially broad management applications in space and time.

    To this end, their contributions inquire into the processes and factors that influence populations and communities. Such a focus promotes broad applicability. It is hoped that biologists and administrators working far beyond the oil fields of arctic Alaska will find within these pages information they can use in their own attempts to manage species and ecosystems in the almost ubiquitous presence of human enterprise.

    REFERENCES

    Adam, K.M., Hernandez, H. Snow and ice roads: Ability to support traffic and effects on vegetation. Arctic. 1977;30:13–27.

    Alexander, V., Van Cleve, K. The Alaska Pipeline: A success story. Annu. Rev. Ecol. Syst.. 1983;14:443–463.

    Batzli, G.O., Brown, J. RATE—the influence of grazing on arctic tundra ecosystems. Arctic Bull.. 1976;2(9):153–160.

    Brooks, J.W., Bartonek, J.C., Klein, D.R., Spencer, D.L., Thayer, A.S. Environmental influences of oil and gas development in the Arctic Slope and Beaufort Sea. U.S. Fish Wildl. Serv., Resour. Publ.. 1971;96:1–24.

    Brower, D.C. Fifty Years Below Zero: A Lifetime of Adventure in the Far North. New York: Dodd, Mead; 1942.

    Brown, J., Grave, N.A. Physical and thermal disturbance and protection of permafrost. U.S. Army Cold Reg. Res. Eng. Lab., Rep.. 1979;SR79-5:1–43.

    Brown J., Miller P.C., Tieszen L.L., Bunnell F.L., eds. An Arctic Ecosystem: The Coastal Tundra at Barrow, Alaska. Stroudsburg, PA: Dowden, Hutchinson & Ross, 1980.

    Connors, P.G., Risebrough, R.W. Shorebird dependence on arctic littoral habitats. NOAA/OCSEAP. Environ. Assess. Alaskan Cont. Shelf. 1979;1:271–329.

    Craig, P.C. An introduction to anadromous fishes in the Alaskan arctic. Biol. Pap. Univ. Alaska, Spec. Rep.. 1989;24:27–54.

    Craig, P.C., Griffiths, W.B., Johnson, S.R., Schell, D.M. Trophic dynamics in an arctic lagoon. In: Barnes P.W., Schell D.M., Reimnitz E., eds. The Alaskan Beaufort Sea: Ecosystems and Environments. New York: Academic Press; 1984:342–380.

    Curatolo, J.A., Murphy, S.M. The effects of pipelines, roads, and traffic on the movement of caribou, Rangifer tarandus. Can. Field-Nat.. 1986;100(2):218–224.

    Derksen, D.V., Rothe, T.C., Eldridge, W.D. Use of wetland habitats by birds in the National Petroleum Reserve—Alaska. U.S. Fish Wildl. Serv. Resour. Publ.. 1981;141:1–27.

    Dunton, K.M., Reimnitz, E., Schonberg, S. An arctic kelp community in the Alaskan Beaufort Sea. Arctic. 1982;35:465–484.

    Dygas, J.A., A study of wind, waves, and currents in Simpson Lagoon. Environmental Studies of an Arctic Estuarine System—Final Report. Rep. EPA-660/3-75-026. Washington, DC: U.S. Environ. Prot. Agency; 1975:15–44.

    Engelmann, R.J. Environmental assessment research in Alaska—an overview. Sci. Alaska. 1976;27:83–108.

    Everett, K.R., Murray, B.M., Murray, D.F., Johnson, A.W., Linkeur, A.E., Webber, P.J. Reconnaissance observations of long-term natural vegetation recovery in the Cape Thompson region, Alaska, and additions to the checklist of flora. U.S. Army Cold Reg. Res. Eng. Lab., Rep. 85-11. 1985:1–76.

    Felix, N.A., Jorgenson, M.T. Effects of winter seismic exploration on the coastal plain of the Arctic National Wildlife Refuge, Alaska, 1984. U.S. Fish Wildl. Serv., ANWR Prog. Rep.. FY 85-1, 1985.

    Garner G.W., Reynolds P.E., eds. Arctic National Wildlife Refuge Coastal Plain Resource Assessment. Final Report Baseline Study of the Fish, Wildlife, and their Habitats. Anchorage, AK: U.S. Fish and Wildl. Serv.; 1986.

    Hanley, P.J., Hemming, J.E., Morsell, J.W., Morehouse, A., Leask, L.E., Harrison, G.S. Natural resource protection and petroleum development in Alaska. U.S. Fish Wildl. Serv., Rep. FWS/OBS-80/22. 1981.

    Hobbie, J.E. The ecology of tundra ponds of the arctic coastal plain: A community profile. U.S. Fish Wildl. Serv., Rep. FWS/OBS-83/25. 1984:1–52.

    Jamison, P.L. Enthnohistory of research populations. In: Jamison P.L., Zegura S.L., Milan F.A., eds. Eskimos of Northwestern Alaska: A Biological Perspective. Stroudsburg, PA: Dowden, Hutchinson & Ross; 1978:31–39.

    Johnsgard, P.A. The Plovers, Sandpipers, and Snipes of the World. Lincoln: University of Nebraska Press; 1981.

    Jorgenson, M.T., Joyce, M.R. Six strategies for rehabilitating land disturbed by oil development in arctic Alaska. Arctic. 1994;47:374–390.

    Kogl, D., Schell, D., Colville delta fisheries research. Environmental Studies of an Arctic Estuarine System—Final Report. Rep. EPA-6603-75-026. Washington, DC: U.S. Environ. Prot. Agency; 1975:483–504.

    Kovacs, A., Mellor, M. Sea ice morphology and ice as a geologic agent in the southern Beaufort Sea. In: Reed J., Sater J., eds. The Coast and Shelf of the Beaufort Sea. Arlington, VA: Arctic Institute of North America; 1974:113–161.

    Lawson, D.E. Response of permafrost terrain to disturbance: A synthesis of observations from northern Alaska, U.S.A. Arct. Alpine Res.. 1986;18:1–17.

    MacKay, J.R. Disturbances to the tundra and forest tundra environment of the western Arctic. Can. Geotech. J.. 1970;7:420–432.

    Maki, A.W. Of measured risks: The environmental impacts of the Prudhoe Bay, Alaska, oil field. Environ. Toxicol. Chem.. 1992;11:1691–1707.

    McCabe, T.R. Assessing values of Arctic wildlife and habitat subject to potential petroleum development. Landscape Urban Plann.. 1994;28:33–45.

    McKendrick, J.D. Plant succession on disturbed sites, North Slope, Alaska. Arct. Alpine Res.. 1987;19:554–565.

    Milan, F.A. Demography and population parameters of the present inhabitants of Northwest Alaska. In: Jamison P.L., Zegura S.L., Milan F.A., eds. Eskimos of Northwestern Alaska: A Biological Perspective. Stroudsburg, PA: Dowden, Hutchinson & Ross; 1978:222–232.

    Morrow, J.E. The Freshwater Fishes of Alaska. Anchorage: Alaska Northwest Publishing Company; 1980.

    Newbury, T.K. Under landfast ice. Arctic. 1983;36:328–340.

    Reimnitz, E., Maurer, D.K. Effects of storm surges on the Beaufort Sea coast, northern Alaska. Arctic. 1979;32:329–344.

    Reynolds, J.F., Tenhunen, J.D. Ecosystem response, resistance, resilience, and recovery in arctic landscapes: Introduction. In: Reynolds J.F., Tenhunen J.D., eds. Landscape Function and Disturbance in Arctic Tundra. New York: Springer-Verlag; 1996:3–18.

    Ritchie, W., Walker, H.J. Riverbank forms of the Colville River. In: Reed J., Sater J., eds. The Coast and Shelf of the Beaufort Sea. Arlington, VA: Arctic Institute of North America; 1974:545–562.

    Selkregg, L.L. Alaska Regional Profiles. Fairbanks: University of Alaska Arctic Environ. Inf. and Data Cent.; 1975;Vol. 11.

    Shaver, G.R. Integrated ecosystem research in northern Alaska, 1947–1994. In: Reynolds J.F., Tenhunen J.D., eds. Landscape Function and Disturbance in Arctic Tundra. New York: Springer-Verlag; 1996:19–34.

    Short, A.D., Coleman, J.M., Wright, L.D. Beach dynamics and nearshore morphology of the Beaufort Sea coast, Alaska. In: Reed J., Sater J., eds. The Coast and Shelf of the Beaufort Sea. Arlington, VA: Arctic Institute of North America; 1974:477–488.

    Unpublished report sponsored by BP Exploration (Alaska) Inc., Anchorage Troy Ecological Research Associates (TERA), The fate of birds displaced by the Prudhoe Bay Oil Field: The distribution of nesting birds before and after P-pad construction. 1990.

    Unpublished report sponsored by BP Exploration (Alaska) Inc., Anchorage Troy Ecological Research Associates (TERA), Bird use of the Prudhoe Bay Oil Field. 1992.

    U.S. Geological Survey (USGS). An environmental evaluation of potential petroleum development on the National Petroleum Reserve in Alaska. Prepared by USGS under Section 105(b) of the Naval Petroleum Reserves Production Act of 1976. 1979.

    Walker, D.A. Vegetation and environmental gradients of the Prudhoe Bay region, Alaska. U.S. Army Cold Reg. Res. Eng. Lab., Rep. 85-14. 1985.

    Walker, D.A., Walker, M.D. History and pattern of disturbance in Alaskan arctic terrestrial ecosystems: A hierarchical approach to analyzing landscape change. J. Appl. Ecol.. 1991;28:244–276.

    Walker, D.A., Binnian, E.F., Lederer, N.D., Nordstrand, E.A., Walker, M.D., Webber, P.J. Cumulative Landscape Impacts in the Prudhoe Bay Oil Field, 1949–1983. Anchorage, AK: U.S. Fish and Wildl. Serv., Habitat Resour. Sect.; 1986.

    Walker, D.A., Webber, P.J., Binnian, E.F., Everett, K.R., Lederer, H.D., Nordstrand, E.A., Walker, M.D. Cumulative impacts of oil fields on northern Alaska landscapes. Science. 1987;238:757–761.

    Walker, D.A., Cate, D., Brown, J., Racine, C. Disturbance and recovery of arctic Alaskan tundra terrain. U.S. Army Cold Reg. Res. Eng. Lab., Rep.. 1987;87-11:63.

    Wilimovsky N.J., Wolfe J.N., eds. Environment of the Cape Thompson Region, Alaska. Oak Ridge, TN: U.S. At. Energy Comm., Div. Tech. Inf., 1966.

    CHAPTER 2

    North Slope Oil Field Development

    Michelle A. Gilders,     Canmore, ALberta, Canada, and LGL Limited, Sidney, British Columbia, Canada

    Matthew A. Cronin,     LGL Alaska Research Associates, Inc., Anchorage, Alaska

    Introduction

    The History of North Slope Oil Exploration and Development

    The Nature of Oil Development

    Identifying and Monitoring Environmental Impacts

    Mitigating Environmental Impacts

    Conclusions

    References

    INTRODUCTION

    The Arctic Coastal Plain of northern Alaska covers an area of approximately 230,000 km² north of the Brooks Range, with 71,000 km² located between the Colville and Canning rivers (Senner, 1989). Within this region are several producing oil fields [BP Exploration (Alaska) Inc. and ARCO Alaska, Inc., 1997; BP Exploration (Alaska) Inc. (BPXA), 1998]: Prudhoe Bay, Kuparuk, Milne Point, Pt. McIntyre, Endicott, Lisburne, Niakuk, and Badami (Fig. 1, Table I). Several other fields (Alpine, Northstar, Liberty, and Pt. Thomson) are currently in the planning phase of development (Fig. 1, Table I) (Ambrosius, 2000). All lands currently used for oil and gas development and production on the North Slope are leased from the state of Alaska, although in the future federal lands will be involved, as in the proposed Liberty development. More than 2000 wells have been drilled north of the Brooks Range, a region commonly referred to as the North Slope. Since production began at Prudhoe Bay in 1977, more than 12 billion barrels of oil have been transported to the marine terminal in Valdez via the 1,288-km Trans-Alaska Pipeline (BPXA, 1998).

    TABLE I

    Oil Fields on Alaska’s Arctic Coastal Plain

    *Oil field refers to both units and participating areas (PA’s). There are 6 additional PA’s on the North Slope whose oil is processed by existing facilities (i.e. no additional surface impact).

    **Unit areas cannot be totaled because overlap exists among the units and participating areas.

    ***Alaska Department of Revenue, Oil and Gas Audit Division, January 25, 2000.

    FIGURE 1 The North Slope oil fields of arctic Alaska have a complex infrastructure of gravel roads, pads, and pipelines, located between the Colville and Sagavanirktok rivers.

    Total disturbed surface area (gravel placement, mine sites, and TAPS north of the Brooks Range) due to oil and gas activity is approximately 8793 ha (0.04% of the Arctic Coastal Plain, or 0.1% of the coastal plain between the Colville and Canning rivers) (Ambrosius, 2000).

    The North Slope portion of the Arctic Coastal Plain is generally flat with a few small streams and rivers and thousands of ponds and small lakes (Walker 1985, Walker and Walker, 1991) (Fig. 2; see color plate). The area is underlain by permafrost that may extend from within a meter of the surface to a depth of 660 m (Hobbie, 1984) (Fig. 3; see color plate). Permafrost sustains the ponds and lakes by preventing percolation of surface waters. The result is a summer wetland landscape with ponds and lakes dominating the surface, even though annual precipitation is less than 13–18 cm (Selkregg, 1975). In winter, all but the deepest water bodies freeze completely to the bottom.

    FIGURE 2 The North Slope summer landscape is covered with numerous small ponds, lakes, and wetland complexes. (Photo by Danny Lehman.)

    FIGURE 3 A cross-section of the tundra active layer reveals the underlying permafrost.

    During the brief 2- to 3-month summer, when temperatures can reach 27°C (although the July mean temperature is only 10°C) and wetlands dominate the landscape, wildlife is abundant (Walker, 1985; Walker and Walker, 1991; Brooks, et al., 1971; Garner and Reynolds, 1986). Some 15 species of terrestrial mammals, 6 species of marine mammals, and 240 species of birds (approximately 180 species breed on the North Slope, with absolute numbers approaching 10 million individual birds; Johnson and Herter, 1989; Derksen et al., 1981) are found in or near the oil fields. The vast majority of wildlife species are present only during the summer, arriving in late May or early June and leaving by late August or September.

    Only five species of birds are regular winter residents on the North Slope (Johnson and Herter, 1989): common raven (Corvus corax), snowy owl (Nyctea scandiaca), glaucous gull (Larus hyperboreus), ptarmigan (Lagopus spp.), and gyrfalcon (Falco rusticolus). Many of the terrestrial mammals either hibernate or migrate to avoid the harsh northern winter (56 days of total darkness and January temperatures of −30 to −40°C). Other mammals become nomadic (arctic foxes [Alopex lagopus]) or remain active beneath the snowpack (collared lemming [Dicrostonyx groenlandicus] and brown lemming [Lemmus sibiricus]). The marine mammals also are largely migratory, the majority moving into the Bering and Chukchi seas in September and October as the landfast ice connects with the floating pack ice of the Beaufort Sea.

    THE HISTORY OF NORTH SLOPE OIL EXPLORATION AND DEVELOPMENT

    The history of oil and gas exploration and development on Alaska’s North Slope is a relatively short one. The first geological surveys took place in the late 1950s and early 1960s. At that time, much of the interest was focused well to the west of the existing oil fields—within the boundaries of what became the National Petroleum Reserve in Alaska (NPRA)—and in the foothills of the Brooks Range. None of the wells drilled proved

    Enjoying the preview?
    Page 1 of 1