+ All Categories
Home > Documents > Arctic Biodiversity Assessment - Amazon S3€¦ · 628 Arctic Biodiversity Assessment Species...

Arctic Biodiversity Assessment - Amazon S3€¦ · 628 Arctic Biodiversity Assessment Species...

Date post: 20-Jun-2020
Category:
Upload: others
View: 4 times
Download: 0 times
Share this document with a friend
24
628 Arctic Biodiversity Assessment Species protection has focused on preventing overharvest, which has historically been the largest threat to Arctic biodiversity. Seabirds are an example of biodiversity that is susceptible to such overharvest, and this has caused population declines in some parts of their range. In these areas, careful regulation of harvest is necessary as part of a conservation and restoration strategy. Kippaku, NW Greenland. Photo: Knud Falk.
Transcript
Page 1: Arctic Biodiversity Assessment - Amazon S3€¦ · 628 Arctic Biodiversity Assessment Species protection has focused on preventing overharvest, which has historically been the largest

628 Arctic Biodiversity Assessment

Species protection has focused on preventing overharvest, which has historically been the largest threat to Arctic biodiversity. Seabirds are an example of biodiversity that is susceptible to such overharvest, and this has caused population declines in some parts of their range. In these areas, careful regulation of harvest is necessary as part of a conservation and restoration strategy. Kippaku, NW Greenland. Photo: Knud Falk.

Page 2: Arctic Biodiversity Assessment - Amazon S3€¦ · 628 Arctic Biodiversity Assessment Species protection has focused on preventing overharvest, which has historically been the largest

629

Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 630

19.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 630

19.2. Disturbance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63019.2.1. Human footprint in the Arctic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63019.2.2. Direct impacts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63119.2.3. Indirect and cumulative impacts . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63219.2.4. Other indirect or cumulative impacts . . . . . . . . . . . . . . . . . . . . . . . 63319.2.5. Herbivore responses to disturbance . . . . . . . . . . . . . . . . . . . . . . . . 63419.2.6. Future prospects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 634

19.3. Feedbacks to ecosystems and climate . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63519.3.1. Greenhouse gases: carbon dioxide and methane . . . . . . . . . . . . 63519.3.2. Permafrost degradation and changes to soil conditions . . . . . 63619.3.3. Air pollutants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63619.3.4. Snow and ice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63619.3.5. Vegetation shifts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63619.3.6. Changes in lake area . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63719.3.7. Future prospects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 637

19.4. Conservation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63719.4.1. Habitat protection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 637

19.4.1.1. Aichi target 11 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63919.4.1.2. Sacred sites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64019.4.1.3. Potential for habitat shifts . . . . . . . . . . . . . . . . . . . . . . . . . . 64019.4.1.4. Future prospects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 641

19.4.2. Species protection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64119.4.2.1. Challenges . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64219.4.2.2. Management and regulation . . . . . . . . . . . . . . . . . . . . . . . 64319.4.2.3. Future prospects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 643

19.4.3. Conservation through community involvement . . . . . . . . . . . . . 64419.4.3.1. Monitoring approaches . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64419.4.3.2. Validity of CBM data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64419.4.3.3. Challenges . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64519.4.3.4. Contributions to biodiversity monitoring . . . . . . . . . . . 64619.4.3.5. Future prospects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 647

19.5. Discussion and conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 647

Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 648

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 648

Chapter 19

Disturbance, Feedbacks and ConservationLead AuthorHenry P. Huntington

Contributing AuthorsTom Arnbom (species protection), Finn Danielsen (conservation through community involvement), Martin Enghoff (conservation through community involvement), Eugénie Euskirchen (feedbacks to ecosystems and climate), Bruce C. Forbes (disturbance), Tiina Kurvits (habitat protection), Nette Levermann (conservation through community involve-ment), Peter Løvstrøm (conservation through community involvement), Kaisu Mustonen (traditional knowledge), Tero Mustonen (traditional knowledge), Martin Schiøtz (conservation through community involvement), Martin Sommerkorn (feedbacks to ecosystems and climate), Michael Svoboda (conservation through community involvement), Elmer Topp-Jørgensen (conservation through community involvement), Geoff York (species protection)

Contents

» We were told not to hunt animals for the sense of killing. Because you are not able to use that animal for eternity. I believe we were also taught that there is a certain pur-pose here in this particular time for us to utilize these marine mammals. That was what I heard the elderly people say from the older generation, like Pelaasi and others, used to say. They were saying: the ‘plan’ has been already made. The ‘master plan’ is that our purpose is to hunt marine mammals, but that we should not take that for granted. This is why conservation is so important in our culture.

George Noongwook in Oozeva et al. 2004.

Page 3: Arctic Biodiversity Assessment - Amazon S3€¦ · 628 Arctic Biodiversity Assessment Species protection has focused on preventing overharvest, which has historically been the largest

630 Arctic Biodiversity Assessment

SUMMARY

Humans disturb the environment in various ways, notably from industrial development and other activi-ties in formerly pristine areas. Components of the earth system affect one another in a web of feedbacks, includ-ing between ecosystems and climate. Conservation is the human attempt to avoid or minimize negative impacts of human activity on species and habitats. This chapter examines all three topics.

Disturbance here refers to the disruption of normal eco-logical functions or distributions at the landscape level. While many types of human activity can affect local en-vironments, industrial development is most likely to af-fect larger areas, followed by spatially extensive practices such as reindeer herding that can lead to heavy grazing and trampling. Around the Arctic, human activity is increasing, with more roads and other infrastructure, leading to a greater overall impact, especially in areas with oil and gas or other valuable commodities.

Feedbacks occur in many forms at many scales. Here, we look at the primarily positive feedbacks from Arctic warming to global climate, which are likely to lead to still greater warming. For example, the loss of ice and snow leaves a darker surface, so that more sunlight is ab-sorbed, leading to greater warming and so on. Changes in the Arctic’s role in the carbon cycle, through release of carbon dioxide and methane and possible increased uptake of carbon dioxide through increased vegetation growth, will affect global climate. Forcing through positive feedbacks is likely to outweigh the impacts of negative feedbacks within the Arctic.

Three measures of conservation are addressed next. Habitat protection is usually measured in terms of protected areas, which are generally strong on land in the Arctic but nearly absent in the marine environment. Species protection includes those species listed in various categories at risk of extinction, and unfortunately these lists appear to be growing in the Arctic as elsewhere in the world. Effective conservation also requires the participation of the people who are likely to either create threats or be affected by management measures. A grow-ing number of programs seek to include Arctic residents in gathering, analyzing and making use of observational data, which often cannot be obtained in other ways.

19.1. INTRODUCTIONHumans interact with Arctic ecosystems in many ways. This chapter examines three types of interactions: disturbance, feedbacks and conservation. Disturbance is the effect that human activity has on the natural environ-ment. Taken broadly, nearly everything humans do cre-ates some form of disturbance, since the natural world is altered by our presence and our activities. Some of this interaction is the normal result of people living as part of the ecosystem, and thus does not constitute a threat.

For the purposes of this assessment, disturbance refers more narrowly to the disruption of normal ecological functions or distributions at the landscape level, posing a threat to biodiversity. While many types of human activ-ity can affect local environments, industrial development is most likely to affect larger areas, followed by spatially extensive practices such as reindeer Rangifer tarandus herding that can lead to heavy grazing and trampling of vegetation.

Feedbacks are in one sense a part of the natural world, constraining the natural cycles of weather, climate and biology. Feedbacks large and small are thus pre-sent throughout the world. While negative feedbacks tend to push a system back to its original state, posi-tive feedbacks lead to ever greater or faster change. Melting snow, ice and permafrost in the Arctic are one such positive feedback, and are described herein. This feedback is already important at the global scale, and likely to become even more significant in the near future as sea ice retreats in summer, snow cover becomes less extensive in space and time, and permafrost degrades and thaws, all of which will lead to greater warming and thus further change.

On a more optimistic note, conservation efforts are hu-mankind’s attempt to reduce its negative impacts on the environment. Habitat protection recognizes that biodi-versity requires intact ecosystems for natural processes to continue. Species protection focuses directly on popu-lations that are at risk, aiming at its simplest to avoid extinction from human causes. Environmental monitor-ing is essential to determine what is at risk and whether conservation efforts are succeeding. In the Arctic, sparse populations and remote areas create a special need for the involvement of local residents in community-based initiatives. These three conservation measures are ad-dressed here.

This chapter is neither exhaustive nor definitive on the topics of disturbance, feedbacks and conservation. It aims instead at describing key aspects of human-ecosys-tem interactions, focusing on matters of special signifi-cance in the Arctic, with the expectation that these areas of focus will be relevant markers for future consideration of trends in Arctic ecosystems and their relationship to humans and to the world as a whole.

19.2. DISTURBANCE

19.2.1. Human footprint in the Arctic Widespread concern regarding the prospect of large-scale resource development in the Arctic was first expressed 40 years ago (e.g. Bliss 1970a, 1970b, Bellamy et al. 1971). Human activities and associated disturbance re-gimes are much more extensive within the tundra biome than in the past. Sizable portions of the region are faced with widespread threats ranging from onshore and off-shore oil and gas development to long-distance transport

Page 4: Arctic Biodiversity Assessment - Amazon S3€¦ · 628 Arctic Biodiversity Assessment Species protection has focused on preventing overharvest, which has historically been the largest

Chapter 19 • Disturbance, Feedbacks and Conservation 631

of contaminants, from mining to tourism. Direct effects on certain Arctic ecosystems with a significant human presence, such as the hydrocarbon fields of northern Rus-sia, are likely to be even more imperative than climate change in the next few decades. These effects include direct and indirect impacts associated with, for example, resource exploitation and altered grazing regimes due to changing patterns of reindeer husbandry. Evidence shows that even small scale, low intensity disturbances can ac-cumulate spatially from local to regional scales.

Besides habitat disturbance, human activities may cause disturbance in the form of displacement (scaring) of wildlife from preferred habitat. This is not dealt with here, but see Meltofte et al. (Chapter 1), Reid et al. (Chapter 3) and Ganter & Gaston (Chapter 4).

19.2.2. Direct impactsIn general, the direct mechanical disturbance of Arctic terrain, including vegetation, soil and the underlying permafrost layer, can lead to erosion where slope and/or ice-rich permafrost are present (Forbes et al. 2001). Unchecked, severe erosion can progress to eventually de-grade entire landscapes. Among the aforementioned three components, vegetation has special importance, not only as the basic link to the upper trophic levels of an ecosys-tem, but also in terms of its controls over permafrost and ground-ice maintenance in tundra substrates. In addition, the regeneration of an ecosystem after disturbance is de-pendent upon revegetation, which is the essential first step of ecosystem recovery. Vegetation cover, therefore, is one of the best criteria to assess overall ecosystem status in the wake of previous environmental degradation. Resto-ration efforts are generally lacking in tundra ecosystems because of the constraints imposed by climate, although assisted revegetation efforts can succeed under certain circumstances when viable seeds or vegetative cuttings are properly cultivated and subsidies of nutrients and water are sufficient. Such efforts can be expensive, however, and most disturbed terrain is left to revegetate naturally, ex-cept where control of aeolian erosion is essential (Forbes & McKendrick 2002). Regeneration is slower in the high Arctic compared with the low Arctic and proceeds more

quickly on moist-wet terrain, unless there has been sub-sidence from thawing permafrost. Some of the scars from oil exploration on Alaska’s North Slope in the 1950s are still visible today (Forbes et al. 2001).

Anthropogenic impacts are complex in that various hu-man activities can influence ecosystems simultaneously and cumulatively, and can have both immediate cata-strophic and long-term effects. In practice it can be dif-ficult to distinguish between direct and indirect impacts, and scientists may use different methods for classifying disturbances (Crawford 1997, Gilders & Cronin 2000, Nellemann et al. 2001, National Research Council 2003). For example, Russian scientists distinguish three main classes of disturbed areas: ochagovyi (local), lineinyi (lin-ear) and fonovyi (spatial) (Khitun & Rebristaya 2002). The most striking example of the first type includes sites sur-rounding petroleum bore-holes (drill sites). Recent data from Arctic Russia indicate that each drilling denudes vegetation over an area of about 120-200 m in diameter, with moderate impacts beyond that distance (Forbes et al. 2009). Transport corridors appearing in connection with road and pipeline construction constitute linear disturbances. Large territories affected by air pollution are examples of spatial disturbances. Especially in the older gas and oil fields, the amount of terrain disturbed on Russian territories exceeds by an order of magnitude that from North American (Tab. 19.1). The track record in Russia has improved in recent years, particularly in the case of post-Soviet joint ventures, such as the Ardalin Oil Field first developed in 1993-94 by ConocoPhillips’ Polar Lights Co. in the Timan-Pechora basin (Rasmussen & Koroleva 2003, Stammler & Forbes 2006).

» Freedom…this modern situation cannot be compared to the case when I was young. You cannot compare these times, and

it did not change naturally, instead it is like a revolution has taken place. Once the highway was made here [in 1957], it changed every aspect of our system here… I think of this so that I have already witnessed several revolutions so far, even though I am not older than I am.

(North Saami hunter and Elder Aslak Ola Aikio, Utsjoki, Finland, regarding the arrival of roads, in Mustonen 2012).

Northwest Siberia Area (km2) Area (km2) Alaska

Yamal-Gydan region 235,000 230,000 Arctic Coastal Plain

Yamal Peninsula 112,000 71,000 Canning to Colville River

Total area disturbed (Yamal-Gydan)a 6,000-7,000 785 Total area disturbed (Trans-Alaska Pipeline System)b

Bovanenkovo Gas Field (BGF)c 2,052 991 Prudhoe Bay Oil Field (PBOF)c

Bovanenkovo Gas Fieldd 200 16.9 Prudhoe Bay Oil Fieldd

Severely disturbed terrain (BGF) 277-287 8.8 Severely disturbed terrain (PBOF)

Indirect impact zone (BGF) 448 na Indirect impact zone (PBOF)

To be disturbed in the near futuree 500 na To be disturbed in the near future

a) Estimate from 1994. b) Includes territory south of the Brooks Range. c) Above-ground facilities. d) Below-ground lease area. e) Includes territory between BGF and Kharasavey.

Table 19.1. Comparison of land areas impacted by oil and gas exploitation in northern Alaska and NW Siberia, respectively (Walker 1996, Khitun 1997, Gilders & Cronin 2000, Khitun & Rebristaya 2002, VNIPIGazdovycha 2005, Forbes et al. 2009).

Page 5: Arctic Biodiversity Assessment - Amazon S3€¦ · 628 Arctic Biodiversity Assessment Species protection has focused on preventing overharvest, which has historically been the largest

632 Arctic Biodiversity Assessment

The point at which small disturbances create large impacts can be hard to identify, but for indigenous peoples it may begin with a sense of breaking the ‘whole’ that constitutes their environment and the role of humans therein.

One of the most widespread types of direct impact is damage to tundra from off-road vehicle traffic (Forbes et al. 2001, Kumpula et al. 2011). In vehicle tracks, plant and soil nutrient cycling regimes can become significantly different than in undisturbed areas, with increases and decreases variable among species, growth forms and soil types. Although the actual ruts may be small to begin with, the shift from scale-of-impact to scale-of-response can be several orders of magnitude, as in the case of drained wetlands. Even shallow ruts from as little as a single-pass vehicle track are capable of effectively diverting runoff from spring snowmelt away from wet and mesic sedge fens that depend upon this source of moisture. Such desiccation of wet tundra has resulted in the local extinction of aquatic sedges Carex spp., sphagnum mosses Sphagnum spp. and other hydric bryophytes, as well as an increase in surface albedo (Forbes 1997). Similarly, as little as a single passage of a vehicle in summer is sufficient to significantly reduce the abundance of soil arthropods (Kevan et al. 1995). In areas with substantial ground-ice, thermokarst activity can expand appreciably. In northern Alaska, some disturbances on silty sediments covered at least twice the original area of impact after 30 years, but most off-road traffic has been effectively banned since the mid-1970s (Forbes et al. 2001). A similar ban has been in place in the hydrocarbon fields of Arctic Russia since the late 1980s, but has proven far less effective (Khitun & Rebristaya 2002, Kumpula et al. 2011).

» We often learn of plans for industrial development only by accident; for example we have learned about the shelf

oil-field development project from the American side, and we began to write about it. The authorities often hold back ecological information from us that is important for the society. That’s why we often support something without being aware of the ecological consequences.

(Tatiana Yuryevna Achyrgina in Novikova 2008).

19.2.3. Indirect and cumulative impactsIn addition to direct disturbances of the ground surface, other, less visible impacts can accumulate over time. These may occur independently of each other, or may be exacerbated through synergy among various proxi-mal effects. These indirect or cumulative impacts are well-documented, especially in the hydrocarbon fields of North America and Russia. Whereas in the early years of development they were often unforeseen, scientists are now better able to predict them (e.g. Gilders & Cronin 2000, Forbes et al. 2001, Kumpula et al. 2011).

Given the large amount of hydrocarbon extraction activi-ties that have taken place over several decades on both the North Slope of Alaska and the Yamal-Nenets Autonomous Okrug (YNAO), it is worthwhile to compare the ex-

tent of impacts in these two regions (Tab. 19.1). In most cases, the extent of indirect impacts exceeds the physical footprint of an Arctic oil or gas field complex, although efforts at mitigation have continued to improve (Forbes et al. 2009). For example, construction of the entire 1,288 km Trans-Alaska Pipeline directly disturbed 785 km2 of land (Walker 1996). It is claimed that if Alaska’s Prudhoe Bay Oil Field were developed today using current tech-nology and consolidation of facilities, gravel would cover at least 80% less area, and the oil field’s direct footprint would be less than half its current size (Gilders & Cronin 2000, National Research Council 2003). In the case of the YNAO and the Nenets Autonomous Okrug (NAO), where migratory Nenets and their large reindeer herds move back and forth across actively exploited fields, it is important to note that it is not only a matter of how much territory is affected, but what kind of territory and which migrations routes are affected by the losses. Further-more, research based on extensive participant observation and interviews with nomads has revealed that each terri-tory has its own particular meaning and importance for users, so that one territory is not equivalent to another (Forbes et al. 2009, Kumpula et al. 2011).

By 1994, disturbed terrain comprised an estimated 6,000-7,000 km2 of the YNAO. Most of the damage to date has been in the southern portion of the region (Khitun & Rebristaya 2002). The impact in the northern tundra zone is still in the early stages, but includes in-tensive terrestrial and aquatic impacts in and around the Bovanenkovo Gas Field on the territory of the Yarsalins-ki sovkhoz. At that time, the gas field encompassed more than 200 km2, of which half was severely disturbed and which affected narrow migratory corridors. As of 2005, the visibly affected area around Bovanenkovo, includ-ing both direct and indirect impacts, had encompassed about 450 km2 (Forbes et al. 2009). This has increased in the last several years as the gas field expands to the northeast so the affected area was 836 km2 as of sum-mer 2011 (Kumpula et al. 2012). Oil development in the Arctic regions of the NAO is some years behind that of the YNAO but is catching up quickly.

Several other indirect (and direct) impacts in both ter-restrial and aquatic habitats stem from roads and rail-ways. One example is the construction of a transport corridor on Yamal Peninsula between Bovanenkovo and the port of Kharasavey to the northwest (Fig. 19.1). Ac-cording to nomadic Nenets reindeer herders, the build-ing of about 130 bridges initially degraded key rivers and lakes so that the supply of fish for daily subsistence, especially critical for reindeer herders during summer migration, was less reliable for several years (Forbes et al. 2009). However, in 2011 the same herders reported that fish had begun to return in significant numbers (Kumpula et al. 2012). In the absence of strict regulatory oversight, poaching can be a chronic problem whenever access to formerly remote regions becomes possible for the general population. Thus, the ongoing influx of workers in both the NAO and YNAO is certain to in-crease fishing pressure and accelerate this process since

Page 6: Arctic Biodiversity Assessment - Amazon S3€¦ · 628 Arctic Biodiversity Assessment Species protection has focused on preventing overharvest, which has historically been the largest

Chapter 19 • Disturbance, Feedbacks and Conservation 633

enforcement of existing regulations remains lax relative to North America (Forbes et al. 2009). Throughout the Arctic, gravel roads and sand quarries are subject to wind erosion and can spread sand and dust up to one kilometer from the source (Forbes 1995, Myers-Smith et al. 2006). Road dust is alkaline and is capable of rapidly smothering bryophytes, lichens and mushrooms on the surface. Dust significantly increases the pH of soils and surface waters, and alters the nutrient contents of abun-dant vascular plants and mosses in as few as four years. During the same time period, blowing sand can bury all mosses and lichens, and many vascular plants, up to a distance of 250 m from the source (Forbes 1995).

19.2.4. Other indirect or cumulative impacts

Additional types of impacts result from oil leaks and spills. Massive oil spills, such as the Exxon Valdez or the Deepwater Horizon, have not yet occurred in truly Arctic waters. However, the multifaceted concerns sur-rounding offshore development continue to grow quickly as the economic viability of fields in the Beaufort, Bar-ents and Kara Seas increases annually, where difficulties in cleaning up after a spill would likely be exacerbated by cold temperatures and ice in part of the year (Mar-gesin & Schinner 1999). The scope of actual and poten-tial development, and the environmental risks involved,

Figure 19.1. Map of the Yamal Peninsula, Russia, showing industrial infrastructure and other features.

Page 7: Arctic Biodiversity Assessment - Amazon S3€¦ · 628 Arctic Biodiversity Assessment Species protection has focused on preventing overharvest, which has historically been the largest

634 Arctic Biodiversity Assessment

has been detailed in two landmark reports by AMAP (1998, 2007). Onshore, one of the largest spills in his-tory occurred near Usinsk, Komi Republic, in 1994, resulting in a release of oil into tributaries of the Pechora River estimated to be eight times greater than the Exxon Valdez spill (Crawford 1997). Experts argue that since cleanup of such accidents can be difficult or impossible, as well as astronomically expensive, the best solution is a sustained, all-out effort at prevention (Jernelöv 2010). However, the spotlight on British Petroleum’s (BP) prac-tices in recent years in Alaska and the Gulf of Mexico has detailed the extent to which certain companies clearly prefer to channel resources into production and profits rather than field safety and longevity (Graham & Reilly 2011, Goldenberg 2011, New York Times 2011a, 2011b). Other important issues are airborne pollution from the flaring of excess gas within active fields and, in the Murmansk and Norilsk regions, from smelters. Evi-dence from both North America and Eurasian high lati-tudes indicates that pollutants, including heavy metals, accumulate in terrestrial and aquatic systems downwind and can persist for decades (AMAP 1998). As in the case of drained wetlands cited earlier (Forbes et al. 2001), the shift from scale-of-impact to scale-of-response can be several orders of magnitude when pollutants spread out-ward from point sources and either settle on the surface or are entrained in atmospheric air currents for longer periods and carried out to sea, contributing to Arctic haze (AMAP 1998, 2007).

» The economic development of Chukotka has influenced the life of Eskimos because the environment has deteriorated and

land rehabilitation is not being realized. Atmospheric pollution leaves its mark. Living in the permafrost zone means that filtration occurs very slowly, all the pollutants remain on top of the ground, the reindeers then eat it, and finally people eat their meat. The same thing happens to the sea. Nobody cleans oil spills, especially in cold waters. The pollutants do not disappear without a trace. Once we brought a whale and the meat was polluted with spilled oil.

(Raisa Mikhaylovna Zotova in Novikova 2008).

19.2.5. Herbivore responses to disturbanceHerbivory is important as a force to contend with, both as a form of disturbance in itself and as a potential limiting factor during succession (Forbes & McKendrick 2002) and under a warming climate (Post & Pedersen 2008, Olofsson et al. 2009). This is particularly the case in the relatively lush lowland and coastal tundra ecosystems with high herbivore densities where extensive hydrocar-bon extraction is now active (e.g. Prudhoe Bay, Alaska; NAO, Russia) and expected to spread to (e.g. Naval Petroleum Reserve-Alaska and Arctic National Wild-life Refuge, Alaska; YNAO, Russia). Equally important is the displacement of populations of large herbivores (Cameron et al. 2005). The responses of caribou/rein-deer to disturbance can be complex and highly variable depending upon sex, age and season, among other factors (Cameron et al. 2005, Haskell et al. 2006). In the case of

new infrastructure, such as roads or power lines, animals (and particularly pregnant females) may initially avoid these. To date, the evidence that animals can adapt to the presence of infrastructure and associated disturbance within or across years is inconsistent (Haskell et al. 2006, Vistnes & Nellemann 2008).

Grazing by vertebrate herbivores can have profound effects on dynamic processes in Arctic ecosystems, particularly in successional communities. In addition to favoring graminoids and weedy mosses at the expense of lichens and certain selected dwarf shrubs, grazing is an ecologically important limiting factor in the regenera-tion of many vascular plant species (Forbes & Kumpula 2009). Numerous Arctic researchers have noted that her-bivores ranging in size from lemmings to caribou/rein-deer are attracted to the plants growing on experimental fertilization plots and that they can affect the structure, cover/abundance and successional trajectory of the af-fected communities. Caribou, for example, may use sites with high forage nitrogen concentrations more intensely as a strategy of maximizing nutrient intake, leading to a positive feedback loop over the long-term (Forbes & McKendrick 2002). The same pattern has been observed in the boreal zone, where selective grazing of vegeta-tion plots has been reported for periods of up to five years after a one-time addition of NPK fertilizer (John & Turkington 1997). This has serious implications for areas where assisted revegetation is attempted because, although plants are selected primarily to prevent erosion, many of the chosen species also provide important food for wildlife. Since it is standard practice to apply organic matter and/or chemical fertilizers to subsidize the initial stages of growth, managers trying to restore the origi-nal vegetation cover need to be wary about the access of herbivores to sites either naturally recovering from disturbance or actively revegetated. On the other hand, if the aim is simply to encourage the growth of forage for herbivores then regardless of the species composition, free access to fertilized sites is encouraged because of the positive feedbacks likely to encourage biomass accumula-tion (Forbes & McKendrick 2002).

With regard to climate change, the prevailing assump-tion until recently has been that the dominance of woody shrubs will increase under ongoing and future warming. At least in systems with low erect shrubs (e.g. dwarf birch Betula nana, willow Salix spp.), large herbivores like reindeer and muskoxen may be able to check increases in shrub biomass/height (Post & Pedersen 2008, Olofs-son et al. 2009, Ims & Ehrich, Chapter 12). However, in cases where high erect shrubs are already above the browse line (c. 1.5-1.8 m), herbivory does not limit shrub growth (Forbes et al. 2010, Macias-Fauria et al. 2012).

19.2.6. Future prospectsDuring the next few decades, hotspots to observe for extensive terrestrial ecosystem disturbances include the major oil and gas bearing regions, such as the Alaskan

Page 8: Arctic Biodiversity Assessment - Amazon S3€¦ · 628 Arctic Biodiversity Assessment Species protection has focused on preventing overharvest, which has historically been the largest

Chapter 19 • Disturbance, Feedbacks and Conservation 635

North Slope, the Mackenzie River delta in Canada and the Timan-Pechora and W Siberian basins of NW Rus-sia. Cumulative impacts from piecemeal tourism and residential development are of concern to rural commu-nities in northernmost Fennoscandia, where new cabins, power lines and expanding ski areas and road networks threaten to further fragment territories used for forage and migration by reindeer and other forms of wildlife (Vistnes & Nellemann 2008). Another emerging topic being watched closely in Fennoscandia is that of the rap-idly-expanding mining industry. Similarly, the develop-ment of wind power in northern and high elevation areas in Fennoscandia is being closely watched. The empirical literature is still quite limited, but a recent national re-view in Sweden identified groups of terrestrial mammals potentially affected by existing and planned develop-ments (Helldin et al. 2012). Interestingly, the authors conclude that one of the presumed knock-on effects of wind power is very similar to the aforementioned piecemeal developments. Namely, that the expanding road system to access turbines will enhance access for recreation, hunting and leisure traffic, likely resulting in impacts on populations of wild and domestic reindeer, moose and large carnivores (Helldin et al. 2012). For the time being, however, the data remain inconclusive.

19.3. FEEDBACKS TO ECOSYSTEMS AND CLIMATE

Changes in climate affect the structure and function of ecosystems. The biosphere and the atmosphere are a fully coupled system, therefore changes in the structure and function of terrestrial ecosystems may, in turn, feedback to the climate. In order to project Arctic (and global) climate variability into the future with certainty, these feedback loops must be understood. In this sec-tion, we focus on: (1) the influence of climate on Arctic ecosystems and (2) the regional and global feedbacks to climate by these ecosystems. We examine a number of these climatically sensitive processes and feedbacks, including carbon and methane cycling, permafrost dynamics, soil conditions, air pollutants, snow and ice cover dynamics, vegetation shifts, fire regimes and lake area. Cryosphere phenomena such as snow cover extend far beyond the Arctic, but we have not attempted to separate the Arctic component of feedbacks such as the snow-albedo effect. It is simpler, and more consist-ent with current research and modeling, to use the full winter snow cover for the northern hemisphere. Similar considerations have been used for other feedbacks.

19.3.1. Greenhouse gases: carbon dioxide and methane

The Arctic plays an important role in the global carbon budget, making significant contributions to the global fluxes of carbon dioxide and methane between ecosys-tems and the atmosphere. Increasing concentrations of atmospheric greenhouse gases are key driving factors

in warming trends in the Arctic. Both carbon dioxide (CO2) and methane (CH4) are increasing in the atmo-sphere, and are estimated to have caused a c. 1.66 W/m2 and c. 0.48 W/m2 increase in radiative forcing globally since 1750, respectively (Forster et al. 2007). Methane is present in the atmosphere in much smaller concen-trations compared with carbon dioxide, but is relatively more potent with a high potential for global warming. Over a 100-year time scale, methane is 25 times more effective per molecule than CO2 at absorbing long-wave radiation, despite its shorter lifetime in the atmosphere. Since terrestrial ecosystems fix CO2 through photosyn-thesis and release it through respiration, any change that impacts these processes will feedback to climate.

These fluxes of carbon are particularly important since, in their efforts to regulate carbon emissions, govern-ments rely on estimates of carbon losses and gains re-lated to climate change. Notably, the ‘social cost of car-bon’, the estimated price of damages caused by each ton of CO2 released into the atmosphere, varies by country. Consequently, any mechanism that results in increased CO2 sequestration in Arctic ecosystems, such as in-creased vegetation growth, would have a positive impact on climate mitigation. Likewise, any mechanism that causes decreases in CO2 sequestration, including the po-tential loss of carbon from Arctic ecosystems caused by increased development and human disturbance, would have a negative impact on climate mitigation. While the monetary value of carbon sequestration potential in the tropics is now part of the UN-led climate negotiations (through the instrument Reducing Emissions from De-forestation and Forest Degradation (REDD)), this issue is not recognized in the case of Arctic ecosystems.

While Arctic terrestrial ecosystems are currently esti-mated to be a sink of atmospheric CO2, the strength of this sink in a warmer Arctic is forecast to deteriorate and may switch to acting as a source in the future (Canadell & Raupach 2009, McGuire et al. 2009). This is due to a variety of factors, including increases in the decompo-sition of soil organic matter under a warming climate, permafrost degradation and acclimation of the plants to increased atmospheric CO2. Consequently, this means that while Arctic terrestrial ecosystems currently exert a decelerating (i.e. negative) feedback to raising atmo-spheric greenhouse gas concentrations, in the future they are projected to wield a less decelerating, or even accelerating (i.e. positive) feedback to climate warming, although this depends on the effect of increased veg-etation growth, as mentioned below. As the amounts of carbon stored in Arctic soils that are vulnerable to the effects of warming are vast, the potential of Arctic terrestrial ecosystems to accelerate climate warming is significant. In terms of methane, the Arctic is currently a source of atmospheric methane, due in large part to methane emissions from lakes and wetlands. It is thought that methane emissions will continue or increase in the warmer Arctic of the future, providing a positive feed-back to climate warming (McGuire et al. 2009).

Page 9: Arctic Biodiversity Assessment - Amazon S3€¦ · 628 Arctic Biodiversity Assessment Species protection has focused on preventing overharvest, which has historically been the largest

636 Arctic Biodiversity Assessment

19.3.2. Permafrost degradation and changes to soil conditions

Predicting the response of permafrost thaw to climate warming is complicated by the wide variety of factors that influence soil temperature, including air tempera-ture, snow depth, topographic effects on insolation, soil texture, organic layer depth, surface water and runoff, groundwater movement and soil moisture. Studies have documented increases in permafrost degradation across the Arctic (Jorgenson et al. 2001, Zhang et al. 2005), and with ongoing warming these trends are expected to continue (Lawrence et al. 2008). Permafrost is a strong heat sink that reduces surface temperature and heat flux to the atmosphere, and consequently, the thawing of permafrost releases heat, causing a positive climate feedback. Moreover, recent estimates suggest that the top three meters of permafrost soils contain more than twice the amount of carbon as the atmosphere (Tarnocai et al. 2009). This carbon has been accumulating over long periods of time as a result of cold and waterlogged permafrost soil conditions. Degradation of permafrost – from deepening of the annually developing thawed layer to its complete disappearance near the ground surface – fundamentally enhances the conditions for soil microorganisms to decompose old soil carbon. As a result, greenhouse gas releases from thawing permafrost may act as a more sustained and much larger positive climate feedback than previously thought (Schuur et al. 2008). An important landscape aspect controlling both magnitude and direction of climate feedbacks following permafrost thaw is the resulting soil wetness near the surface. An increase in wetness will promote anaerobic conditions and increased methane emissions while slow-ing overall decomposition and CO2 release. In contrast, dryer conditions can promote greater decomposition and release of CO2 and decrease methane emissions. On an ecosystem scale, climate feedbacks associated with carbon and methane releases related to permafrost degradation are likely substantial during this century – in the range of those projected to be released by global deforestation scenarios (Zhuang et al. 2006).

19.3.3. Air pollutantsWhile greenhouse gases are the dominant driving factor in warming trends and climate feedbacks in the Arctic and globally, Arctic air pollutants, including aerosols, are also important (Key & Stohl 2007). The aerosols are usually introduced to the Arctic from Eurasia in the form of sulfates and black carbon (soot). Boreal forest fires and tundra fires (Hu et al. 2010) act as significant aerosol pollution sources in the circumpolar Arctic and may become more prevalent in the future. The pollu-tion influences the Arctic climate through changes in surface radiative forcing, i.e. heat being absorbed at, or near, the Earth’s surface. Some of these particles of pollution absorb sunlight, acting as a positive feedback to warming, while others reflect sunlight, acting as a negative feedback to warming. For example, deposition of black carbon on snow reduces surface albedo, and

acts as a positive feedback to warming, while sulfates scatter incoming solar radiation, resulting in a cooling effect. Globally, the radiative forcing of aerosols is a negative feedback to warming (Myhre 2009), but the net radiative forcing of these pollutants in the Arctic is still uncertain and a topic of study. However, recent work suggests that decreasing concentrations of sulfate aero-sols and increasing concentrations of black carbon have substantially contributed to rapid Arctic warming during the past three decades (Shindell & Faluvegi 2009).

19.3.4. Snow and iceSnow and ice albedo feedback loops in the Arctic are strong: as snow or ice melts, a dark surface is exposed, less solar energy is reflected back to space, and more energy is absorbed and transferred to the atmosphere, causing a positive feedback loop that reinforces warming. Across the Arctic, and between 1970 and 2000, a de-crease in duration of approximately 2.5 days per year of the snow season translate to a 2.5 W/m2 decade warm-ing during this same period (Euskirchen et al. 2007). Changes in ice cover also represent a strong positive feedback to warming. The extent of sea ice has declined since the beginning of the record in 1953, with the low-est value recorded in 2012 (Stroeve et al. 2007, Perovich et al. 2012; see Fig. 1.5 in Meltofte et al., Chapter 1), and a strong thinning of multiyear ice and an increase in the area of melt ponds. All of these factors exacerbate the ice-albedo positive feedback loop to warming (Light et al. 2008, Pedersen et al. 2009). The additional or amplified warming caused by the loss of sea ice is not constrained to the Arctic Ocean, but also influences adjacent land ar-eas, especially during autumn and winter, and may lead to hastened degradation of certain types of permafrost (Lawrence et al. 2008; see also Section 19.3.2). There is also a negative feedback following the loss of sea ice due to an increase in evapo-transpiration, causing an increase in summer clouds, which then increases net radiation, and decreases heating to the atmosphere (Chapin et al. 2005). However, this negative feedback is expected to be relatively weak, and will likely not counteract the strong ice-albedo feedback loop (Chapin et al. 2005).

19.3.5. Vegetation shiftsStudies have documented recent changes in the vegeta-tion in the Arctic. This has included treeline advance-ment in some areas, retreat in other areas, and an en-croachment of tall, woody shrubs in the tundra. Treeline advancement and tall, woody shrub encroachment are likely due to a longer growing season with increases in temperature and moisture (Sturm et al. 2001, Lloyd et al. 2003). This replacement of tundra with boreal forest and increases of tall, woody shrubs will result in greater carbon uptake into the vegetation, acting as a negative feedback to climate warming. On an ecosystem scale, it is currently under study whether the vegetation shift will lead to releases of soil carbon that would affect magnitude or direction of this negative feedback. How-ever, any event that causes an advance of treeline and

Page 10: Arctic Biodiversity Assessment - Amazon S3€¦ · 628 Arctic Biodiversity Assessment Species protection has focused on preventing overharvest, which has historically been the largest

Chapter 19 • Disturbance, Feedbacks and Conservation 637

shrubs will reduce albedo, causing a positive feedback to warming (Sturm et al. 2005, Euskirchen et al. 2009). In fact, research has shown that the net uptake or release of carbon associated with changes in treeline is likely a much smaller feedback to climate than the feedback due to changes in surface energy balance (Betts 2000).

Treeline retreat is likely in some areas due to drought stress under high temperatures, which then interacts with slow recruitment and reduced seed sources to decrease the success of tree regeneration following disturbance (McGuire et al. 2010). As a result, increased proportions of forests may regenerate as open forests or shrubland. In addition, permafrost degradation may also cause a decline of forest extent as forests may be replaced by bogs. Open forest, shrubland, or bogs would store less vegetation car-bon but more soil carbon than a forest with the resulting net carbon feedback depending on the relative magnitude of these effects. However, the land surface of these less vegetated ecosystems would have a generally high albedo and act as a negative feedback to warming.

19.3.6. Changes in lake areaMethane emissions from Arctic lakes are substantial, and increases in their emissions act as a positive feedback to global climate warming (Walter et al. 2007). Changes in lake area in the Arctic have been documented due to permafrost thaw. Consequently, in the future, the amount of methane emissions may be highly dependent on changes in lake area. In southern areas of warm per-mafrost, studies have generally documented a decrease in lake area due to lake drainage following permafrost thaw (Smith et al. 2005, Riordan et al. 2006), whereas lake area tends to increase with permafrost thaw in northern ice-rich zones of cold permafrost (Smith et al. 2005; see also Wrona & Reist, Chapter 13). In addi-tion to changing methane emissions, these increases or decreases in lake area would also impact albedo, with increases in lake area resulting in an increase in albedo and a negative climate feedback. Decreases in lake area will likely not have the opposite positive feedback effect, as the new surface can also have an increased albedo, dependent on the type of vegetation colonizing (Rouse et al. 2005). Overall, while the magnitude of the climate feedback due to changes in lake area has not been quanti-fied, it would depend on the relative changes of methane emissions versus albedo.

19.3.7. Future prospectsIn the coming decades, we will continue to observe changes in the sink strength of the Arctic in terms of car-bon and methane, the duration of the snow and ice cover, the integrity of the permafrost, and vegetation shifts, all of which will generally promote positive feedbacks to cli-mate. In fact, the number and strength of positive feed-backs to climate will likely continue to be larger than the number and strength of negative feedbacks (McGuire et al. 2006, Euskirchen et al. 2010), and there is indi-cation that some positive feedbacks, such as albedo loss

and permafrost thaw, accelerate each other (Canadell & Raupach 2009). Currently, the primary positive climate feedbacks are likely related to changes in surface albedo due to changes in ice and snow cover. While negative feedbacks to climate have been quantified, including increased carbon uptake by vegetation due to a longer growing season, these negative feedbacks are increasingly understood not be large enough – or last long enough – to counterbalance the larger and more sustained positive feedbacks. While models consistently simulate these feedbacks into the future, continuing to measure and monitor key indicators on integrated landscape and regional scales is extremely important. Key indicators to monitor include all those influencing energy fluxes and carbon cycling, for example, permafrost integrity, snow and ice cover duration, extent and thickness, landscape wetness and greenness, vegetation composition, as well as fire regimes and their related successional dynamics. The impacts of Arctic climate feedbacks will extend well beyond the Arctic, necessitating an integrated under-standing of the Arctic ecosystem processes and their representation in global climate models.

19.4. CONSERVATION

19.4.1. Habitat protection Protected areas have long been the foundation of biodi-versity conservation programs. Although many of the first protected areas were established primarily for the purposes of recreation, they have evolved since that time to become important tools for habitat protection and species conservation. This is as true for the Arctic as it is elsewhere.

The first protected areas in the North were established in Sweden and Alaska at the beginning of the 20th centu-ry. It was not until the 1980s that there was a significant increase in number of areas under protection. Recent decades have seen an exponential growth in the number of protected areas in the circumpolar north. As of 2010, there are 1,127 protected areas in the region, covering approximately 3.5 million km2 or 11% of the CAFF co-operation area (Tab. 19.2). These areas vary considerably in terms of size, type and nature of protection.

The International Union for the Conservation of Na-ture and Natural Resources (IUCN) has a classification system for protected areas as a means to help collate data from protected areas around the world. Since its inception, this international system has developed “to help provide a framework in which various protection strategies can be combined together, along with support-ive management systems outside protected areas, into a coherent approach to conserving nature” (Dudley 2008). The categories, therefore, can help countries and regions assess their progress to meeting defined conservation goals. In the Arctic, the majority of protected areas fall into category II, Ecosystem Conservation and Protection (Tab. 19.2, Fig. 19.2).

Page 11: Arctic Biodiversity Assessment - Amazon S3€¦ · 628 Arctic Biodiversity Assessment Species protection has focused on preventing overharvest, which has historically been the largest

638 Arctic Biodiversity Assessment

Cate­gory

Title Managed for No. of pro­tected areas

Total area (ha)

% of Arctic covered

Ia Strict protection: strict nature reserve Science 350 273,000 0.8

Ib Strict protection: wilderness area Wilderness protection 111 795,000 2.5

II Ecosystem conservation and protection (e.g. national park)

Ecosystem protection or recreation 102 1,530,000 4.7

III Conservation of natural features (e.g. natural monument)

Conservation of specific natural features

103 52,600 0.2

IV Conservation through active management (e.g. habitat/species management area)

Conservation through management intervention

125 154,000 0.5

V Landscape/seascape conservation and protection (e.g. protected landscape/ seascape)

Landscape/seascape conservation and recreation

60 64,600 0.2

VI Sustainable use of natural resources (e.g. managed resource protected area)

Sustainable use of natural ecosystems 120 648,000 2.0

No category assigned 156 30,800 0.1

Total 1,127 3,550,000 11.0

Table 19.2. Total Arctic territories in IUCN protection categories (CAFF 2010). N.B.: The definition of Arctic for this table is the CAFF bound-ary, not the ABA boundary.

Figure 19.2. Protected areas within the political cooperation area of CAFF.

Page 12: Arctic Biodiversity Assessment - Amazon S3€¦ · 628 Arctic Biodiversity Assessment Species protection has focused on preventing overharvest, which has historically been the largest

Chapter 19 • Disturbance, Feedbacks and Conservation 639

While protected areas are a powerful conservation tool, they can also – depending on the exact rules in place – constrain some traditional activities, which can erode support for such conservation measures (see Box 19.1 for examples of indigenous views on conservation). Such a situation can be seen, for example, in the Malla nature reserve in Finland, where the removal of human activi-ties has led to conflicts over the impacts of reindeer grazing, which can benefit some species as well as have negative impacts on others (Jokinen 2005). As with other forms of conservation measures, the establishment and management of protected areas needs to address traditional practices and potential conflicts to achieve the overall goals of habitat protection and biodiversity conservation.

19.4.1.1. Aichi target 11

Target 11 of the Convention on Biological Diversity’s (CBD) Aichi Targets states that: “By 2020, at least 17 per cent of terrestrial and inland water, and 10 per cent of coastal and marine areas, especially areas of particular importance for biodiversity and ecosystem services, are conserved through effectively and equitably managed, ecologically representative and well-connected systems of protected areas and other effective area-based conser-vation measures, and integrated into the wider land-scapes and seascapes.” While terrestrial areas are relatively well represented in Arctic protected areas, the marine environment con-tinues to be under-represented (CAFF 2010). Although

The methods of habitat protection are not always viewed favorably by Arctic residents. In some cases, this is due to an interest in resource development. In others, it is a result of real or perceived clashes between protection regulations and traditional practices. The Aleut region of the Kommandorskye Islands in the Bering Sea is one such example (Meschtyb 2008).

Dorfei Semionovich Berezin was born on Bering Island. In general, he says, in the past fishing was good in the rivers that the local population has always used for their subsistence. Today, park regulations only allow three fish species to be taken over the whole summer period. This means that local people have started to poach. Berezin notes that the underlying cause for the illegal fishing, an almost universal phenomenon on the island including the fishing inspection personnel, is the deteriorating standard of life of islanders.

Zinaida Ivanovna Kvasiuk lives in the village of Nikolskoe on the Bering Island. She believes that poorly thought out economic and administrative policies have upset the ecological balance on Bering Island and are damag-ing the traditional Aleut way of life. Zinaida says that the local community used to have a structured economic life based on fishing, hunting of marine mammals and a little bit of farming, so despite the harsh island conditions they were self-sufficient people. She says the park does not do enough for environmental protection but hinders tradi-tional Aleut activities. Zinaida is adamant that the Aleut people cannot live without fish, seal and sea lion meat.

Gennadii Mikhailovich Yakolev was born on Mednyi Island in 1935, and now resides in the village of Nikolskoe. Tra-ditional livelihoods have been the basis of his way of life: “The fat of a seal for Aleut people is like butter for others.”

Box 19.1. Aleut perspectives on national parks on the Kommandorskye Islands, Russia

Tero and Kaisu Mustonen

For him, the traditional use of nature is not only a means of providing food, but is also a specific cultural legacy: “I try to take my grandsons with me so that they can be-come accustomed to real Aleut food.” He feels that the bu-reaucracy, which for him represents the majority culture at its most absurd, has done far too much harm with its directives and policies. The overall impact of the national park in the center of Bering Island has placed traditional Aleut activities under a vast array of regulations.

Nikolai Nikolaevich Tiuterev has similar views. He de-scribes how in the past the local community had hunted for seals in the summer and winter but now it is only per-mitted in the autumn and by specially accredited hunters. Tiuterev recalls earlier years when locals were permitted to hunt sea lions and seals but now this has been banned. He finds the official explanations difficult to understand:

“Our ancestors hunted these animals and their numbers never decreased. Yet, today, the authorities are afraid that we will exterminate them. In the past, when somebody needed a couple of seals for food, they would hunt for them. Why would we want to waste seals by over-hunt-ing? It provides meat for the whole village.

“New regulations state that it is necessary to go to Lake Sarannoe, but this is a considerable distance from the community, and there is no transportation for local fishermen. Many people do not always receive any fish, especially pensioners.”

All these regulations interfere with the practices of the traditional economy and Aleut livelihoods. Moreover, Tiuterev believes that they have a negative impact on the self-worth of the Aleut people as an indigenous culture.

Page 13: Arctic Biodiversity Assessment - Amazon S3€¦ · 628 Arctic Biodiversity Assessment Species protection has focused on preventing overharvest, which has historically been the largest

640 Arctic Biodiversity Assessment

over 40% of the protected areas inside the CAFF cooperation area have a coastal component, the extent to which the neighboring marine environment is incor-porated is undetermined for most (Barry & McLennan 2010). The Arctic marine environment is one of the least understood regions of Earth, especially in the high Arctic. Historically, the harsh conditions of this area have made it difficult to study. With some projections indicating that the Arctic Ocean could be ice-free in late summer by mid-century or even before, the develop-ment pressure will increase. There is a critical need for identification of ecologically important and vulnerable marine areas in the Arctic and recommendations for their management. A recent effort in this direction has been the identification of ecologically and culturally im-portant marine areas, following one of the recommenda-tions of the Arctic Marine Shipping Assessment (AMSA 2009). This work is being carried out under three Arctic Council working groups: the Arctic Monitoring and As-sessment Program (AMAP), CAFF and the Sustainable Development Working Group (SDWG).

19.4.1.2. Sacred sites

The IUCN broadly defines sacred sites as follows: “Sacred site – an area of special spiritual significance to peoples and communities; Sacred natural site – areas of land or water having special spiritual significance to peoples and communities” (Wild & McLeod 2008). They have been further described as “reflect[ing] the diversity of spiritual and cultural values that indigenous peoples attribute to their territories, landscapes, biota, and particular sites” (CAFF 2004).

In the Arctic, in particular, most sacred sites are natural areas, often high in biodiversity values, and so may contribute to biodiversity conservation with a strong link to customary livelihoods (CAFF 2004). A study by CAFF (2004) on the conservation values of sacred sites of indigenous peoples in the Arctic noted that “Not only are most sacred sites located on or in the vicinity to mi-gration routes, fishing sites, or pastures; the active use, maintenance and protection of these sites depend largely on healthy livelihood systems.”

Despite their conservation values, the role of sacred sites in habitat protection and biodiversity conservation has not received the same degree of attention as other types of protected areas. Ideally, sacred sites should be sup-ported by national and regional protected area systems, but this is often not the case (Wild & McLeod 2008). While tangible sacred sites (e.g. human-built structures) tend to be afforded a reasonable degree of protection, intangible sites (e.g. holy rivers and lakes) are under-rep-resented (Wild & McLeod 2008). While it may be that not all practices associated with sacred sites are in line with broader conservation objectives (e.g. conflicts be-tween traditional reindeer herding practices and conser-vation of large predators), sacred sites have the potential to provide an important link in habitat conservation and protected area networks.

One of the challenges to fully incorporating sacred sites into formal protected area networks may be reluctance on the part of indigenous peoples to have their sacred sites formally classified. There are issues with the pro-tection of cultural and intellectual property. The CAFF study also noted that the non-codified status of sacred sites, like traditional knowledge in general, leaves it open to abuse (CAFF 2004). Some of the key recom-mendations from that study include:• developing an action plan for the further integration

of sacred sites and indigenous territories of traditional nature use into broader protected areas networks;

• encouraging indigenous peoples to seek further reporting on sacred sites and their protection into national reporting on the implementation of the CBD;

• bringing to the attention of the World Intellectual Property Organization (WIPO) the need to accom-modate within its work the knowledge about indige-nous sacred sites, as this is a significant and important perspective for the intellectual property, traditional knowledge and genetic resources of indigenous peo-ples (CAFF 2004).

19.4.1.3. Potential for habitat shifts

One of the greatest challenges facing Arctic protected areas is climate change. The changes wrought by a warm-ing climate are raising a number of questions regarding the effectiveness of protected areas as a conservation tool in the future. The changes to the physical environment are already well documented and include such effects as reductions in snow and ice and changes in precipita-tion patterns. The associated impacts as a result of these changes include a northward shift in species, ‘greening’ of the Arctic, changes in timing of key life cycle events and changing migration patterns, to name just a few.

Changes in habitat type, in particular, pose significant threats to protection efforts. Already there is evidence of significant shifts in Arctic vegetation in recent decades, and this is expected to continue with further warm-ing (Henry & Elmendorf 2010). Henry & Elmendorf (2010) noted that treeline encroachment is threatening the southern margins of the tundra. According to some models, treelines may advance by as much as 500 km north over the next century with a resultant loss of 51% of tundra habitat (Callaghan et al. 2005). A study assessing changes in biome types in Canada’s protected areas networks under climate change found that the representation of northern biomes – tundra, taiga/tundra and boreal conifer forest – in protected ar-eas was projected to decrease (Lemieux & Scott 2005). The study projected that 38-79% fewer protected areas will still have part of tundra biomes, and 81-87% fewer protected areas will contain at least part of the taiga/tundra biome (Lemieux & Scott 2005). These decreases are the result of decreases in these biomes overall, so that a lower proportion of tundra overall translates into a lower proportion of tundra within protected areas.

Page 14: Arctic Biodiversity Assessment - Amazon S3€¦ · 628 Arctic Biodiversity Assessment Species protection has focused on preventing overharvest, which has historically been the largest

Chapter 19 • Disturbance, Feedbacks and Conservation 641

A similar and perhaps more dramatic change is occur-ring in the Arctic sea ice habitat. Changes are already be-ing seen in the extent and thickness of the sea ice, with thicker multi-year ice being replaced by thinner first-year ice (NSIDC 2010). Current predictions indicate that the Arctic Ocean could be nearly ice-free in late summer by the middle of this century or even sooner (Wang & Overland 2009). Sea ice represents a unique ecosystem in the Arctic providing habitat for numerous ice-associated species (see Josefson & Mokievsky, Chap-ter 8 and Michel, Chapter 14). Changes in sea ice can be expected to have impacts throughout the marine food web, from phytoplankton and zooplankton to seabirds and marine mammals. Indigenous peoples of the Arctic will also be affected by these changes as many use sea ice for transportation and hunting. In response to the changes occurring in this important ecosystem, CAFF is conducting an Arctic Sea Ice Associated Biodiversity Assessment which will summarize the current status and trends of sea ice-associated biodiversity and recommen-dation actions that might to mitigate these changes.

Further complicating these habitat shifts and changes is the associated issue of invasive species. As the climate warms and more human activity takes place across the Arctic, both northward range expansions and biologi-cal invasions (i.e. transport by humans, intentionally or otherwise) are likely to increase (Lassuy & Lewis 2010; Lassuy & Lewis, Chapter 16).

There is a further issue of habitat fragmentation outside of formally protected areas that may further reduce the ability of species and ecosystems to adapt to change. Al-though such patterns are most likely a long ways off for the Arctic, protected areas can become isolated islands in a broader sea of development (industrial, agricultural, etc.) inhibiting the movement of species to more suitable habitats. The scale of development in most of the cir-cumpolar Arctic is far below that seen in more southern regions, and some of the largest unaltered habitats are found here. The pressure to develop, however, is strong and will likely continue to grow as the Arctic becomes more accessible as a result of climate change.

19.4.1.4. Future prospects

The scale of environmental change facing the Arctic forces us to ask whether protected areas can continue to be an effective conservation tool in the future. The majority of protected areas are selected on the basis of ecosystem representation, where there is an underlying assumption that they will remain static, unchanged. The rapid changes occurring in the Arctic, however, show that this is not guaranteed. With habitat shifts resulting from climate change, it will be more difficult to define ‘natural’ in the future (Lemieux & Scott 2005).

The question remains how protected areas can be used to help ecosystems and species adapt to stressors, cli-mate change in particular. More systematic research and monitoring are needed to address the large uncertainties

facing protected areas in light of climate change. In ad-dition, more efforts are needed to place protected areas in the context of broader habitat conservation measures, i.e. conservation outside of protected areas. New tools will also be needed to help make sound management and policy decisions in a changing Arctic. The WWF project, Rapid Assessment of CircumArctic Ecosys-tem Resilience (RACER), is an example of a tool that has been developed to help identify and map places of importance in the Arctic, looking for areas of resilience that are likely to persist under the changes the Arctic is experiencing (Christie & Sommerkorn 2012).

While protected areas are facing clear threats as a result of climate change, they can also help mitigate some climate change impacts (e.g. carbon sequestration, flood control). They can also provide areas where natural processes can continue and potentially adapt to the impacts of climate change. For these reasons, protected areas will continue to be vital to habitat and biodiversity conservation efforts in the future (e.g. Livingston 2011).

19.4.2. Species protection“A recurring theme in wildlife and fisheries manage-ment over the centuries is that numerical abundance is not always a hedge against declines… We only have to think of salmon, northern cod, (and) bison... What determines persistence is rate of change not the size of the starting population. But numerical abundance carries the risk of over-confidence – ‘there’s still lots of caribou.’ Another contribution to over confidence among users is that the caribou, being cyclic in their abundance, have been low in number before and have come back. However, given changing environmental conditions, the past may not be a secure guide to the future” (CARMA 2010).

The Arctic embraces a wide variety of species of global importance. Almost a quarter of the world’s shorebird species are endemic to the Arctic, and all but three of the world’s 17 Arctic and sub-Arctic goose species have populations numbering in the hundreds of thousands or millions (Ganter & Gaston, Chapter 4). It is also home to several million reindeer and caribou and many unique marine mammals such as the polar bear Ursus maritimus, walrus Odobenus rosmarus and narwhal Monodon monoceros. Seasonal changes are extreme, with dark winters, snow, ice and temperatures plummeting to –50 °C. Summers bask in 24 hour daylight with temperatures soaring to above 20 °C. During this brief summer, several millions of birds and many thousands of terrestrial and marine mammals migrate into the area to breed and take advan-tage of the brief rich feeding grounds.

Historically, these dependable – albeit extreme – condi-tions helped protect Arctic species by limiting physi-cal access. This in turn helped to reduce disturbance, slowed habitat fragmentation and generally limited other human activities. History, however, may no longer be an adequate guide to the future of the Arctic.

Page 15: Arctic Biodiversity Assessment - Amazon S3€¦ · 628 Arctic Biodiversity Assessment Species protection has focused on preventing overharvest, which has historically been the largest

642 Arctic Biodiversity Assessment

19.4.2.1. Challenges

The Arctic is one of the most rapidly changing regions on Earth. Increasing interest in developing natural resources there, coupled with rapid warming, will radically change this area once protected by its inaccessibility and the higher costs associated with extractive industrial devel-opment. These cumulative pressures create significant hurdles for conserving biodiversity (Fig. 19.3). The com-bination of rapid climatic warming and increasing human activities will require the development of new manage-ment tools, investment in basic scientific monitoring, and new governance agreements across the Arctic.

An entire marine ecosystem – from phytoplankton to polar bears and bowhead whales Balaena mysticetus – depends on the continued existence of Arctic sea ice. As temperatures increase and sea ice continues to decline, ice-associated species such as the ringed seal Pusa his-pida, walrus and polar bear will find it more and more difficult to survive within historical ranges or current abundance. On land, thawing permafrost and shifting biophysical drivers will fundamentally alter current ter-restrial ecosystems. Disturbance and fragmentation of habitats through increased human activities (primarily resource extraction) will further complicate conditions in this once relatively undisturbed region.

The conservation of species in the Arctic has tradition-ally focused on large mammals which are, or have been,

commercially harvested. Examples are bowhead whales, walrus, caribou and polar bears. However, the basis for the rich marine and estuarine food webs has rarely, if ever, been acknowledged and protected. For example, the increasing loss of sea ice due to climate warming will have a dramatic impact on the plankton community liv-ing under the ice (see Michel, Chapter 14). That will alter the Arctic ecosystem as we know it today by affecting fish assemblages, ice-associated seals and polar bears (see Reid et al., Chapter 3). The only long-term solutions for pro-tecting this ecosystem and the species that have evolved with it are to decrease global greenhouse gas emissions and to manage human activities inside the Arctic.

In 2006, the polar bear was added to the IUCN’s Red List as a vulnerable species, largely due to predicted impacts from climate change and the expected loss of sea ice habitat. Hudson Bay and southern Beaufort Sea polar bear subpopulations have shown significant declines (Regehr 2007) or metrics of pending decline including decreased adult size and decreased cub survival (Regehr 2010) over the last two decades. These changes in survival and condition have been directly linked to a decrease in summer sea ice habitat as a result of climate warming and are expected to affect polar bears across their range if warming trends continue unabated (Wiig et al. 2008; see also Reid et al., Chapter 3).

While polar bears are often the face of Arctic warming, many species are or will become negatively affected by

Changes in vegetation

Decrease in populationsEcosystem state change

ExtinctionsDecrease in spatial

distribution

Increasedvulnerability

Fragmentation of habitat

Loss of habitat

Decreased habitat quality Invasive species

Population declines

PR

ES

SU

RE

SIM

PA

CT

S

Melting sea ice

Decreasedsnow cover

Changes inprecipitation patterns

Permafrostthawing

Habitat loss alongmigration routes

Depletion of food sourcesalong migration routes

Loss of winteringgrounds outside the Arctic

Long range transportof contaminants

Mineral exploration,extraction and development

Oil spills

Increased human activity(tourism, shipping, development)

Unsustainableharvest

Human Activities

C h a n g e i n N a t i v e B i o d i v e r s i t y

In the ArcticWorldwide

Climate change

Figure 19.3. Currently, the environment in the circumpolar North represents one of the areas of the world with the least human impact. Pressures that exist and are on the increase are climate change, infrastructure development and the development of mineral resources – such as oil and gas and increased economic activity (CAFF 2010).

Page 16: Arctic Biodiversity Assessment - Amazon S3€¦ · 628 Arctic Biodiversity Assessment Species protection has focused on preventing overharvest, which has historically been the largest

Chapter 19 • Disturbance, Feedbacks and Conservation 643

climate change. Hundreds of endemic Arctic species, from small beetles, ice associated algae and plankton, mushrooms, lichens, flowers and lemmings, to large mammals such as caribou, walrus and narwhal face an uncertain future. Barren Ground caribou numbers have dropped across their range in recent years, and experts suspect climate change is a significant contributing fac-tor, though there are signs that the declining trend is re-versing (Reid et al., Chapter 3). Impacts include possible changes in the timing and availability of peak forage in the early summer and increased freezing rain events dur-ing winter that cover the vegetation in ice and decrease availability (Hummel & Ray 2008).

Many Arctic species are migratory and spend most of the year in much lower latitudes. For conservation, this often means protection of an Arctic species has to take place in areas far from the Arctic (Scott 1998). An example is the spoon-billed sandpiper Calidris pygmeus which breeds in the low hundreds in Chukotka, Rus-sia and passes through coastal wetlands in China while migrating to overwintering grounds in SE Asia. During the last 30 years, the number of spoon-billed sandpipers has decreased dramatically from about 6,000 breeding pairs to just a few hundred pairs. Besides loss of habitat in staging and wintering areas, the most eminent cause of the decline appears to be indiscriminate hunting in Myanmar where the birds are caught in mist nets and sold to local markets as food (Zöckler et al. 2010). The only way to save this migrating Arctic species from extinction is to protect it outside the Arctic while safe-guarding critical summer nesting areas within the Arctic (see Ganter & Gaston, Chapter 4).

Diminishing summer sea ice will also lead to an increase of human activities such as shipping, fishing, mining and oil and gas exploration. There is an urgent need for circumarctic management and governance that ensure a stewardship-first approach to these increasing demands for once inaccessible resources (e.g. Chapin et al. 2009a, 2009b). The Arctic remains one of the largest largely intact ecosystems on earth. Careful planning that incor-porates future change and cumulative impact assessment prior to activities proceeding could reduce additional stressors to an already strained system (e.g. Meek 2011).

19.4.2.2. Management and regulation

The most effective way to protect the vast majority of species is to safeguard habitat, which often conserves representative ecosystems (see Section 19.4.1). More typically, species protection has focused on prevent-ing overharvest, which has historically been the largest threat to Arctic biodiversity (e.g. Meltofte et al., Chapter 1, Reid et al., Chapter 3, Ganter & Gaston, Chapter 4, Christiansen & Reid, Chapter 6). Species protection in the Arctic is regulated at different levels. For example, The Agreement on the Conservation of Polar Bears from 1973 sets forth standards for polar bear conservation across its range. This landmark ‘range state’ agreement dramatically improved harvest management of polar

bears and set up a framework to better coordinate and communicate scientific research and circumarctic man-agement (Larsen & Stirling 2009).

Most species protection falls under national legislation, or a mix of national legislation, and bilateral and inter-national agreements. In Greenland for example, the International Whaling Commission (IWC) sets national subsistence quotas for minke whales Balaenoptera acu-torostrata. Narwhals and beluga whales Delphinapterus leucas, however, are regulated by the Canada-Greenland Joint Commission on the Conservation and Management of Narwhal and Beluga (DFO 2008).

As noted above, national parks and nature reserves are often created to protect a certain habitat for individual species. However, rapid climate change means the condi-tions for keeping these specific habitats (and inhabitants) within the reserve boundaries may be altered. In many cases the protected habitat itself will change over time. For example, the tundra in northernmost Scandinavia will shift to forest as the tree line moves north due to increasing temperatures (Heiskanen et al. 2008).

National legislation for typical nature reserves and pro-tected areas is established to preserve what currently ex-ists, but doesn’t address what will happen in the future in the context of a rapidly changing world. Similar chal-lenges exist with conservation constructs at the inter-national level. While the IUCN’s mission, for example, preserves the ‘now’, assuming stable conditions, it does not take into account future changes under unstable sys-tems. Many national parks and nature reserves will not be able to meet the goals they set in terms of protecting viable populations of specific species or unique habitats. New tools and adaptive management strategies will be required as we move into uncharted territory.

19.4.2.3. Future prospects

The changing Arctic environment will put pressure on species as well as entire ecological processes. It is ex-pected that high Arctic species, such as red knots Calidris canutus, will have fewer options in a changing environ-ment, since the high Arctic zone in particular, will be ‘squeezed in’ between the northward expanding low Arctic biome and the Arctic Ocean (Meltofte et al. 2007; see also Ganter & Gaston, Chapter 4). Southern species, such as the red fox Vulpes vulpes, may see range expansions putting Arctic species under pressure (see Reid et al., Chapter 3 and Ims & Ehrich, Chapter 12). Wildlife can try to adapt (an unlikely option given the current and ex-pected rates of change), migrate or face a very uncertain future. Species that today are considered sentinel may be marginalized as ecosystems cross significant thresholds and shift into new phases. Single species protection will still be important, but it will likely become more impor-tant to preserve ecological processes over time.

A variety of regional, national and international legal mechanisms exist to help manage at-risk species such

Page 17: Arctic Biodiversity Assessment - Amazon S3€¦ · 628 Arctic Biodiversity Assessment Species protection has focused on preventing overharvest, which has historically been the largest

644 Arctic Biodiversity Assessment

as the IUCN Red List, the Convention on International Trade in Endangered Species of Wild Flora and Fauna (CITES) and the various national legislation and regu-lations. Current and predicted impacts from rapidly warming climate scenarios have led to an almost pre-emptive push to list species as endangered or threatened at the regional, federal and international levels and often across species ranges. Listing of species under various legal articles provides increased public awareness of spe-cies status and threats, generally increases legal protec-tions and can boost basic research and monitoring efforts as was seen following the 2007 US listing of the polar bear as threatened.

However, initiating increased protections indicates a failure to manage on other fronts and can have unin-tended consequences. The addition of species to higher categories of risk under constructs like the IUCN Red List is nothing to strive for or to celebrate. It measures the continued loss of biodiversity and societal or soci-ety’s lack of understanding, will or ability to successfully manage the challenges facing species today. Range-wide listing decisions, as seen with both the polar bear and ice-associated species of seal listings in the US, may not adequately account for the varying rates of anticipated change across dramatically differing habitats within the Arctic. Existing legal structures, at the international and national levels, were not developed for pervasive, long term threats like climate change and often lack flexibil-ity once enacted.

Polar bears provide one example. There are 19 subpopu-lations, or management units, of polar bears in the Arc-tic inhabiting a range of very different habitats. We are likely to see up to 19 different stories unfold as warming affects different areas at different times and in different ways. While scientists are already noting population de-clines or indices suggesting decline in the most southerly of polar bear populations, it is expected that populations in the higher Arctic will remain more resilient to change in the near term (Wiig et al. 2008; see also Reid et al., Chapter 3). Prescriptive, one-size-fits-all solutions will challenge the acceptance of stakeholders and the people who rely on these species for their own survival both culturally and economically (Dowsley 2010).

Successful management of Arctic species will require new management tools and greater flexibility. The overarching threat posed by rapid climate warming will challenge our best efforts and existing legal mechanisms. It must also be recognized that people live in the Arctic and rely on its wildlife. Any plan to protect Arctic spe-cies must involve the people who live with them. It must understand the food and economic security challenges that come with increased legal protective status, and potential clashes with established indigenous rights. The situation is complex and demands well thought out and complex responses to the threats of today and the chal-lenges of tomorrow.

19.4.3. Conservation through community involvement

The last several decades have seen continued interest in natural resource monitoring that involves both scientists and local stakeholders (Gofman 2010, Huntington 2011). This partnership, often referred to as community based monitoring (CBM), or community-based observations, continues to evolve and exert increased influence on decision making and resource management (Gofman 2010). The scope of CBM is diverse and complex and continues to develop as experiences of integration are shared. Moreover, the overwhelming connection of Arc-tic peoples to the land provides opportunities for strong conservation partnerships, for example initiatives related to ecological monitoring, food security or sacred sites.

In essence, CBM seeks to improve the ability to share observations and understanding of local changes that are occurring in a vast and remote region through the eyes of Arctic residents. The idea is that intimate and multi-generational knowledge held by local stakeholders can help governments and local organizations identify and address serious environment and development chal-lenges at early stages (Harremoës et al. 2001).

19.4.3.1. Monitoring approaches

Monitoring approaches in all Arctic countries have some level of local involvement, and examples of CBM exist throughout the Arctic. These monitoring approaches range from programs involving local stakeholders only in data collection (citizen science) with the design, analysis and interpretation undertaken by professional research-ers, to entirely autonomous monitoring schemes run by local people (see Gofman 2010 for full discussion).

The level of involvement by local peoples beyond pro-ject development and planning to include analysis can contribute to longer-term capacity and implementa-tion benefits beyond just the collected data (Tab. 19.3). Although local residents can unquestionably monitor and report on certain observed changes, their interpretation of the changes and any policy implications they may have are sometimes left aside. However, this is not a problem limited to CBM. From a policy implementation perspec-tive, opportunities to involve Arctic peoples in know-ledge production, in an open and transparent manner, is critical when considering managing individual and commercial activities in the North.

19.4.3.2. Validity of CBM data

The struggle to break through the perceived limitations surrounding CBM is often linked to the approaches and skepticism at the heart of western approaches to knowledge production. Scientists have documented Arctic community members’ detailed knowledge of key components of their environment, such as sea-ice (Laid-

Page 18: Arctic Biodiversity Assessment - Amazon S3€¦ · 628 Arctic Biodiversity Assessment Species protection has focused on preventing overharvest, which has historically been the largest

Chapter 19 • Disturbance, Feedbacks and Conservation 645

ler 2006), weather patterns (Weatherhead et al. 2010) and caribou (Ferguson et al. 1998, Russell et al. 2013). Nevertheless, we know of no studies that have examined the accuracy of community-based monitoring of natural resources in the Arctic. Studies from other parts of the world provide cautious support for the idea that moni-toring by community members can yield results that can be as reliable as those derived from professional, scien-tist-conducted monitoring (e.g. Danielsen et al. 2005, Jones et al. 2008, Rist et al. 2010).

Whereas scientists aspire to be impartial (Beardsley 2010), some fishermen, hunters and environmentally interested people may have a conflict of interest in their assessment of the status of those resources on which they depend for their livelihoods or that they are otherwise interested in (Root & Alpert 1994). For instance, a special local interest in certain resources or a preoccu-pation with certain challenges to resource management may influence which attributes are recorded, when and where. The community perspective is relevant too. Indigenous communities often view scientific initiatives with suspicion, if the scientists do not possess social and cultural skills to appreciate context and locality, creating a need to establish credibility in both directions.

Many of the potential limitations of CBM can be over-come by careful planning, by explicit consideration of likely biases, and by thorough training and supervision of the participants (Danielsen et al. 2009, Gofman 2010, Luzar et al. 2011). It is a challenge, however, that com-munity monitoring can superficially appear low-tech and therefore primitive in a high-tech world. There remains a huge unexplored potential for strengthening monitoring efforts across the Arctic by engaging more communities and encouraging linkages with scientific monitoring pro-grams (Huntington 2008). Often, an investment to build capacity to collect, interpret and manage data are central to maximizing such monitoring efforts (Gofman 2010).

19.4.3.3. Challenges

As the CBM record evolves and demonstrates contin-ued improvement of accessible information on Arctic biodiversity, it is anticipated that there will be a delay between information production and use, accessibility and integration. In northwestern Canada and northeast-ern Alaska, for example, the reporting by the Arctic Borderlands Ecological Knowledge Co-op of CBM data on population health and body condition of the Porcu-pine caribou herd were largely dismissed and underval-ued in favor of scientific models projecting substantive

Table 19.3. Arctic and sub-Arctic natural resource monitoring schemes across a spectrum of possible monitoring approaches based on the relative participation of different actors (modified from Danielsen et al. 2009).

Category of monitoring

Arctic examples Description

Fully autonomous local monitoring

Customary conservation regimes, e.g. in Canada (Ferguson et al. 1998, Moller et al. 2004)

The whole monitoring process – from design, to data collection, to analysis, and finally to use of data for management decisions – is carried out autono-mously by local stakeholders

Inc

re

as

ing

ro

le o

f l

oc

al

st

ak

eh

old

er

sCollaborative moni-toring with local data interpretation

Arctic Borderlands Ecological Knowledge Co-op, Canada (Eamer 2006, Russell et al. 2013); Commu-nity-based monitoring by Inuvialuit Settlement region, Canada (Huntington 2011); Opening Doors to the Native Knowledge of the Indigenous Peoples of the Nenet Autonomous Okrug, Russia (The Asso-ciation of the Nenets People Yasavey and RAIPON); Piniakkanik Sumiiffinni Nalunaarsuineq, Greenland (Danielsen et al. in press)

Locally based monitoring involving local stakehold-ers in data collection, interpretation or analysis, and management decision making, although external scientists may provide advice and training. The original data collected by local people remain in the area being monitored, but copies of the data may be sent to professional researchers for in-depth or larger-scale analysis

Collaborative moni-toring with external data interpretation

Community Moose Monitoring Project, Canada (Gofman 2010); Integrated Ecosystem Management (ECORA), Russia (Larsen et al. 2011)

Local stakeholders involved in data collection and monitoring-based management decision making, but the design of the scheme and the data analysis and interpretation are undertaken by external scientists

Externally driven monitoring with lo-cal data collectors

Bering Sea Sub Network, Alaska and Russia (Gofman & Smith 2009); Environmental Observations of Seal Hunters, Finland (Gofman 2010); Fávllis Network, Norway (Gofman 2010); Monitoring of breeding common eiders, Greenland (Merkel 2010); The Pini-arneq fisheries catch and hunting report database, Greenland

Local stakeholders involved only in data collec-tion stage, with design, analysis and interpretation of monitoring results for decision-making being undertaken by professional researchers, generally far from the site

Externally driven, researcher executed monitoring

Multiple scientist-executed natural resource moni-toring schemes with no involvement of the local stakeholders

Design and implementation conducted entirely by professional scientists who are funded by external agencies and generally reside elsewhere

Page 19: Arctic Biodiversity Assessment - Amazon S3€¦ · 628 Arctic Biodiversity Assessment Species protection has focused on preventing overharvest, which has historically been the largest

646 Arctic Biodiversity Assessment

population declines (Gofman 2010, Russell et al. 2013). Moreover, such projected declines prompted govern-ment and decision makers to push for and build harvest regimes that limited northern residents’ ability to har-vest. Indeed in 2012, several years after the CBM results were released, scientific population surveys revealed record numbers of caribou actually existed. In this case, CBM would have limited harvest concerns and sup-ported improved access to northern food. However, the combination of the potential for conflict of interest and the lack of demonstrable validation capacity may have contributed to placing limited value on the information from this source.

Such examples suggest that efforts to emphasize analysis and integration between the two knowledge produc-tion approaches should continue. Indeed, more recent biodiversity monitoring planning processes are propos-ing ways of integrating and coordinating the methods for knowledge co-production (Gofman 2010, Vongraven et al. 2012). The Circumpolar Biodiversity Monitoring Program’s (CBMP) strategy for bridging some of the structural challenges over the next few years includes improving the access to CBM data via improved provi-sion of and access to metadata, modeling and demon-strating integration examples of CBM with scientific monitoring processes (Gill et al. 2011, Culp et al. 2012).

19.4.3.4. Contributions to biodiversity monitoring

Full participation in biodiversity monitoring programs continues to be a challenge for many Arctic peoples. Greenland’s effort to increase involvement of CBM with management provides one of the promissong stories becoming more common in the Arctic. The Greenland government is piloting a natural resource monitoring system whereby local people and local authority staff are directly involved in data collection, interpretation and resource management. The scheme is called Pini-akkanik sumiiffinni nalunaarsuineq (Opening Doors to Native Knowledge). Four communities in Disko Bay and Umanak/Uummannaq Fjord are involved: Akunnaaq, Kitsissuarsuit, Qaarsut and Jakobshavn/Ilulissat.

As in other parts of the Arctic, the communities in Greenland are widely distributed over a vast territory, and the opportunities for environmental monitoring and for implementing hunting and fishing regulations on the ground are limited. It has long been a priority of the Greenland government to increase the involve-ment of local citizens in the decision-making process related to natural resources (Greenland Government 1999, Haaland et al. 2005). However, there is limited funding available for monitoring Greenland’s resources, and many species and populations are thus monitored infrequently or not at all (Nielsen 2009). There is there-fore insufficient knowledge available about some wildlife populations to guide government decision making and consequently a need to supplement the existing scientist-led monitoring programs with low-cost monitoring, for example through CBM.

The following are examples of how the influence and impact of the data are increasing when it comes to Arctic resource management. In each of the examples, local community observations were central to effecting changes to management regimes.

Conservation of marine habitat: In Akunnaaq, Green-land, the Natural Resource Committee (NRC) recorded trawlers fishing for shrimp in a shallow sea area adjacent to their village on a daily basis. There were 4-5 vessels almost every day throughout April and May 2010. This number was the same as in 2009 but higher than in previous years. Moreover, the vessels were larger and used heavier fishing gear. The NRC in Akunnaaq was worried that potential degradation of the seafloor might affect the breeding and production of Atlantic wolf-fish Anarhichas lupus. The NRC therefore proposed that the municipality should issue an ordinance to restrict the size of vessels in the area.

Influencing marine harvest techniques: One of the attri-butes recorded by Qaarsut NRC concerned their catch of Greenland halibut Reinhardtius hippoglossoides in Umanak/Uummannaq Fjord. On the basis of their catch-and-ef-fort data from long-line fishery, they estimated that the local Greenland halibut population was the same in May 2010 but higher in June-September 2010 than in the same months of 2009. Nevertheless, the NRC was con-cerned that many nets were being set over their longlines and that some nets were left at sea when the sea froze over. This resulted in many rotting fish, which attract-ed Greenland sharks Somniosus microcephalus. The NRC therefore proposed that the municipality should issue an ordinance to restrict net fishing in Umanak/Uumman-naq Fjord. The fisheries legislation in Greenland allows municipalities – subject to ministerial approval – to pro-hibit the use of certain vessels and equipment in specific areas (Greenland Government 1996).

Influencing goose harvest pressure: Members of the Qaarsut NRC have observed that, over the past decade, the population of Canada goose Branta canadensis has risen sharply. Canada goose may out-compete the threatened Greenland white-fronted goose Anser albifrons flavirostris (Boyd & Fox 2008 versus Raundrup et al. 2012). Hunt-ing seasons in Greenland are decided by the Ministry of Fisheries, Hunting and Agriculture on the basis of advice from scientists and from public input during a hearing process. The current hunting season for Canada goose is 15 August to 15 October (Department of Fisheries, Hunting and Agriculture 2011). The NRC proposed that the municipality should suggest to the Ministry that the hunting season for Canada goose be extended, for example by two weeks, to help keep the population from expanding further. However, a recent study has not found such competition between Canada geese and Greenland white-fronts during molt (Raundrup et al. 2012).

In all three examples, it is noteworthy that the proposals if implemented will benefit the people having put them forward. International experiences however suggest that

Page 20: Arctic Biodiversity Assessment - Amazon S3€¦ · 628 Arctic Biodiversity Assessment Species protection has focused on preventing overharvest, which has historically been the largest

Chapter 19 • Disturbance, Feedbacks and Conservation 647

CBM also often leads to people suggesting restrictions in their own take of resources (Danielsen et al. 2007). CBM encourages people to take a long term perspective on the use of resources through facilitating agreements at community and municipal level to increase or reduce the use of resources.

19.4.3.5. Future prospects

The Arctic environment is rapidly changing (e.g. Hinz-man et al. 2005, CAFF 2010) and there is increasing pressure on its natural resources. There is therefore also an increased need for monitoring. To date, many examples exist of Arctic peoples describing the changes they witness related to climate, sea ice and especially to harvested wildlife species. There is a persistent need for more CBM that can detect change, interpret and integrate results, and lead to prompt decision-making to help tackle environmental challenges at operational lev-els of resource management (Huntington & Fox 2005, Danielsen et al. 2010).

Representatives of indigenous communities practice wildlife management guided by their indigenous knowl-edge, realizing that indigenous knowledge and western scientific knowledge are based on different knowledge generation systems or epistemologies (e.g. Agrawal 1995, Huntington et al. 2004). Through CBM, however, it may be possible to find a suitable means of cooperation and collaboration in which monitoring can be based on local observations and knowledge (Pulsifer et al. 2010, van der Velden 2010) and, at the same time, follow principles of data handling and data management in ac-cordance with Western concepts of scientific accuracy (Yoccoz et al. 2001), which is what national government agencies and international conventions require. Sev-eral Arctic programs (including the CBMP) and Arctic peoples have already started to implement strategies to bridge this gap by building structures such as inventories and metadatabases to better access, use and integrate CBM knowledge in the arctic (e.g. Pulsifer et al. 2012).

In combination, the increased need for data and the necessity of promoting locally relevant knowledge and management actions suggest that there are substantial prospects in the coming decades for more CBM around the Arctic, and that such an increase will contribute to effective local conservation actions.

19.5. DISCUSSION AND CONCLUSIONS

The sections of this chapter have addressed a wide range of topics, quantitatively where possible and qualitatively otherwise. Evaluating the status and likely trends of dis-turbances, feedbacks and conservation efforts is not easy (see Tab. 19.4). For example, an increase in the num-ber of species listed as threatened or endangered may indicate greater commitment to species protection, or it

may indicate a greater number of species at risk. More extensive habitat protection will benefit biodiversity, but what occurs outside of protected areas may ultimately be more important, since protected areas are unlikely to cover a majority of the Arctic.

Community involvement offers a number of clear ben-efits, but should not replace national and other monitor-ing and conservation efforts, since community practices may not always be consistent with the protection of biodiversity (see Huntington, Chapter 18). Disturbance is equally clearly a negative outcome of human-ecosys-tem interactions, though the causes vary from industrial exploitation of petroleum and minerals, heavy graz-ing and trampling, and the impacts of climate change. Determining how to address disturbance is thus not always straightforward, especially where large financial interests are at stake. The potential for climate feedbacks to magnify warming trends is worrisome, pointing to the need for global action to address threats with global causes. Action within the Arctic will not always be suf-ficient to conserve Arctic biodiversity.

To monitor trends in these indicators of human actions that affect biodiversity, a set of quantitative indicators should be developed. Other types of disturbance, feed-backs and conservation measures should also be con-sidered. Noise and chemical pollution, including ocean acidification, may disturb the metabolism or behavior of many animals. The Arctic hydrological cycle, includ-ing the potential for sea level rise from melting of ice caps, has feedbacks to the global climate system, and the well-being of migratory species depends on the interre-lationship of Arctic conditions with conditions elsewhere in the annual journeys of those species. Conservation outside of protected areas, the regulation of fishing and hunting, human population growth and the rate of con-sumption of non-renewable resources are all relevant to the success of biodiversity conservation generally.

Indicator Trend Notes

Disturbance Increasing Roads and industrial activity are expanding, creating more potential for disturbance

Feedbacks to ecosystems and climate

? Positive feedbacks, especially to climate, are exacerbating feedbacks from the Arctic to the globe

Habitat protection

Increasing Parks and protected areas are more numerous, but additional designa-tions may become harder; marine protection is nearly absent

Species protection

Increasing Protective measures are increasing, but perhaps reflecting more species in need of protection

Conservation through commu-nity involvement

Increasing Interest in this approach is grow-ing, though the creation of new programs is slow

Table 19.4. Trends in the five indicators considered in this chapter. Note that ‘Increasing’ may be regarded as positive or negative depending on the indicator.

Page 21: Arctic Biodiversity Assessment - Amazon S3€¦ · 628 Arctic Biodiversity Assessment Species protection has focused on preventing overharvest, which has historically been the largest

648 Arctic Biodiversity Assessment

Tracking all potential indicators is not possible, but a robust set of measures against which progress or de-cline can be monitored would greatly help in providing the public and policy makers with a means of assessing whether Arctic communities, Arctic countries and the world as a whole are contributing to the conservation of Arctic biodiversity or the opposite. Without timely and unambiguous measures of performance, uncertainty will provide an excuse for inaction or for accepting greater levels of risk than are consistent with a commitment to protecting the future of Arctic ecosystems and those who use them.

ACKNOWLEDGEMENTSWe thank the US Fish and Wildlife Service for provid-ing core funding for the preparation of this chapter. We thank the many agencies and organizations that have supported the research that is the basis for the sections of this chapter, along with those who have supported the preparation of the sections themselves. And we thank the reviewers, who provided constructive comments that improved the chapter.

REFERENCESAgrawal, A. 1995. Dismantling the divide between indigenous and

scientific knowledge. Dev. Change 26: 413-439.AMAP 1998. Arctic Pollution Issues: a State of the Arctic Envi-

ronment Report. Arctic Monitoring and Assessment Pro-gramme, Oslo.

AMAP 2007. Arctic Oil and Gas 2007. Arctic Monitoring and As-sessment Program, Oslo.

AMSA 2009. Arctic Marine Shipping Assessment 2009 Report. Protection of the Arctic Marine Environment, Ottawa, Canada.

Barry, T. & McLennan, D. 2010. Changes in protected areas. In: CAFF. Arctic Biodiversity Trends 2010 – Selected Indicators of Change, pp 96-98. CAFF International Secretariat, Akureyri.

Beardsley, T.M. 2010. Emotions and engagement. BioScience 60: 775.

Bellamy, D., Radforth, J. & Radforth, N.W. 1971. Terrain, traffic and tundra. Nature 231: 429-432.

Betts, R.A. 2000. Offset of the potential carbon sink from boreal forestation by decreases in surface albedo. Nature 408: 187-190.

Bliss, L.C. 1970a. A biologist explains why we must plan now to protect the Arctic. Science Forum 3(8): 3-8.

Bliss, L.C. 1970b. Oil and the ecology of the Arctic. In: The tun-dra environment. Transactions of the Royal Society of Canada, Series IV, 7: 1-12.

Boyd, H. & Fox, A.D. 2008. Effects of climate change on the breeding success of White-fronted Geese Anser albifrons flaviro-stris in west Greenland. Wildfowl 58: 55-70.

CAFF 2004. The conservation value of sacred sites of indigenous peoples of the Arctic: a case study in Northern Russia. Report on the state of sacred sites and sanctuaries. CAFF Technical Report No. 11. CAFF International Secretariat, Akureyri.

CAFF 2010. Arctic Biodiversity Trends 2010 – selected indicators of change. CAFF International Secretariat, Akureyri.

Callaghan, T.V., Björn, L.O., Chapin III, F.S., Chernov, Y., Chris-tensen, T.R., Huntley, B. et al. 2005. Arctic tundra and polar desert ecosystems. In: ACIA. Arctic Climate Impact Assess-ment, pp 243-352. Cambridge University Press, Cambridge.

Cameron, R.D., Smith, W.T., White, R.G. & Griffith, B. 2005. Central Arctic caribou and petroleum development: distribu-tional, nutritional and reproductive implications. Arctic 58: 1-10.

Canadell, J.G. & Raupach, M.R. 2009. Land carbon cycle feed-backs. In: M. Sommerkorn & S.J. Hassol (eds.). Arctic Climate Feedbacks: Global Implications, pp 69-80. WWF International Arctic Programme, Oslo.

CARMA 2010. Threats to Caribou. www.carmanetwork.com/display/public/Threats [accessed 12 December 2010]

Chapin, F.S. III, Sturm, M., Serreze, M.C., McFadden, J.P., Key, J.R., Lloyd, A.H. et al. 2005. Role of land-surface changes in arctic summer warming. Science 310: 657-660.

Chapin, F.S. III, Carpenter, S.R., Kofinas, G.P., Folke, C., Abel, N., Clark, W.C. et al. 2009a. Ecosystem stewardship: sustain-ability strategies for a rapidly changing planet. Trends in Ecol-ogy and Evolution 24: 241-249.

Chapin, F.S. III, Kofinas, G.P. & Folke, C. (eds.) 2009b. Principles of Ecosystem Stewardship Springer, New York.

Christie, P. & Sommerkorn, M. 2012. RaCeR: Rapid assessment of circum-arctic ecosystem resilience, 2nd ed. WWF Global Arctic Programme, Ottawa.

Crawford, R.M.M. (ed.) 1997. Disturbance and Recovery in Arctic Lands: An Ecological Perspective. Kluwer Academic, Dordrecht.

Culp, J.M., Lento, J., Goedkoop, W., Power, M., Rautio, M., Christoffersen, K.S. et al. 2012. Developing a circumpolar monitoring framework for Arctic freshwater biodiversity. Biodiversity 13: 215-227.

Danielsen, F., Burgess, N.D. & Balmford, A. 2005. Monitoring matters: examining the potential of locally-based approaches. Biodiv. Conserv. 14: 2507-2542.

Danielsen, F., Mendoza, M.M., Tagtag, A., Alviola, P.A., Balete, D.S., Jensen, A.E. et al. 2007. Increasing conservation man-agement action by involving local people in natural resource monitoring. Ambio 36: 566-570.

Danielsen, F., Burgess, N.D., Balmford, A., Donald, P.F., Funder, M., Jones, J.P.G. et al. 2009. Local participation in natural re-source monitoring: a characterization of approaches. Conserv. Biol. 23: 31–42.

Danielsen, F., Burgess, N.D., Jensen, P.M. & Pirhofer-Walzl, K. 2010. Environmental monitoring: the scale and speed of implementation varies according to the degree of peoples involvement. J. Appl. Ecol. 47: 1166-1168.

Danielsen, F., Topp-Jørgensen, E., Levermann, N., Løvstrøm, P., Schiøtz, M., Enghoff, M. et al. in press. Counting what counts: using local knowledge to improve Arctic resource manage-ment. Polar Geography.

Department of Fisheries, Hunting and Agriculture 2011. dk.nanoq.gl/Emner/Erhverv/Erhvervsomraader/Fangst_og_Jagt/Jagttider.aspx [accessed 20 January 2012]

DFO 2008. Canada/Greenland Joint Commission on Narwhal and Belugas. www.dfo-mpo.gc.ca/CSAS/Csas/Publications/SAR-AS/2008/SAR-AS2008_035_E.pdf [accessed 14 June 2011]

Dowsley, M. 2010. The value of a polar bear: evaluating the role of a multiple-use resource in the Nunavut mixed economy. Arctic Anthropology 47: 39-56.

Dudley, N. (ed.) 2008. Guidelines for Applying Protected Area Management Categories. IUCN, Gland.

Eamer, J. 2006. Keep it simple and be relevant: the first nine years of the Arctic Borderlands Ecological Knowledge Co-op. In: W.V. Reid, F. Berkes, T. Wilbanks & D. Capistrano (eds.). Bridging Scales and Knowledge Systems, pp 185-206. Island Press, Washington DC.

Euskirchen, E.S., McGuire, A.D. & Chapin, F.S. III 2007. Energy feedbacks of northern high-latitude ecosystems to the climate system due to reduced snow cover during 20th century warm-ing. Global Change Biology 13: 2425-2438

Euskirchen, E.S., McGuire, A.D., Chapin, F.S. III, Yi, S. & Thompson, C.C. 2009. Changes in plant communities in northern Alaska under scenarios of climate change 2003-2100:

Page 22: Arctic Biodiversity Assessment - Amazon S3€¦ · 628 Arctic Biodiversity Assessment Species protection has focused on preventing overharvest, which has historically been the largest

Chapter 19 • Disturbance, Feedbacks and Conservation 649

Implications for climate feedbacks. Ecological Applications 19: 1022-1043.

Euskirchen, E.S., McGuire, A.D., Chapin, F.S. III & Rupp, T.S. 2010. The changing effects of Alaska boreal forests on the climate system. Canadian Journal of Forest Research 40: 1336-1346.

Ferguson, M.A.D., Williamson, R.G. & Messier, F. 1998. Inuit Knowledge of Long-term Changes in a Population of Arctic Tundra Caribou. Arctic 51: 201-219.

Forbes, B.C. 1995. Tundra disturbance studies. III. Short-term effects of aeolian sand and dust, Yamal Region, northwest Sibe-ria, Russia. Environmental Conservation 22: 335-344.

Forbes, B.C. 1997. Tundra disturbance studies. IV. Species estab-lishment on anthropogenic primary surfaces, Yamal Peninsula, Northwest Siberia, Russia. Polar Geography 21: 79-100.

Forbes, B.C. & Kumpula, T. 2009. The ecological role and geogra-phy of reindeer (Rangifer tarandus) in northern Eurasia. Geog-raphy Compass 3/4: 1356-1380.

Forbes, B.C. & McKendrick, J.D. 2002. Polar tundra. In: M. Perrow & A.J. Davy (eds.). Handbook of Ecological Resto-ration, Vol. 2: Restoration in Practice, pp 355-375. Cambridge University Press, Cambridge.

Forbes, B.C., Ebersole, J.J. & Strandberg, B. 2001. Anthropogenic disturbance and patch dynamics in circumpolar arctic ecosys-tems. Conservation Biology 15: 954-969.

Forbes, B.C., Stammler, F., Kumpula, T., Meschtyb, N., Pajunen, A. & Kaarlejärvi, E. 2009. High resilience in the Yamal-Nenets social-ecological system, West Siberian Arctic, Russia. Pro-ceedings of the National Academy of Sciences 106: 22041-22048.

Forbes, B.C., Macias-Fauria, M. & Zetterberg, P. 2010. Russian Arctic warming and ‘greening’ are closely tracked by tundra shrub willows. Global Change Biology 16: 1542-1554.

Forster, P., Ramaswamy, V., Artaxo, P., Berntsen, T., Betts, R., Fahey, D.W. et al. 2007. Changes in Atmospheric Constituents and in Radiative Forcing. IPCC Fourth Assessment Report WG1. Cambridge University Press, Cambridge.

Gilders, M.A. & Cronin, M.A. 2000. North Slope oil field de-velopment. In: J.C. Truett & S.R. Johnson (eds.). The Natural History of an Arctic Oil Field – Development and the Biota, pp 15-33. Academic Press, San Diego.

Gill, M., Crane, K., Hindrum, R., Arneberg, P., Bysveen, I., Denisenko, N.V. et al. 2011. Arctic Marine Biodiversity Monitoring Plan (CBMP-MARINE PLAN), CAFF Monitoring Series Report No.3, April 2011, CAFF International Secretariat, Akureyri, Iceland.

Gofman, V. 2010. Community based monitoring handbook: lessons from the Arctic. CAFF CBMP Report No. 21. CAFF, Akureyri.

Gofman, V. & Smith, M. 2009. Bering Sub-Sea Network Pilot Phase Final Report. CAFF Monitoring Series Report No. 2. CAFF, Akureyri.

Goldenberg, S. 2011. BP’s Gulf of Mexico oil spill was ‘caused by cost-cutting’. The Guardian Weekly 14 January 2011, p 17.

Graham, B. & Reilly, W.K. 2011. Deepwater: The Gulf Oil Disas-ter and the Future of Offshore Drilling. Report to the Presi-dent. National Commission on the BP Deepwater Horizon Oil Spill and Offshore Drilling, Washington.

Greenland Government 1996. Landstingslov nr. 18 af 31. oktober 1996 om fiskeri. www.lovgivning.gl/gh.gl-love/dk/1996/Ltl/ltl_nr_18-1996_dk.htm [accessed 20 January 2012]

Greenland Government 1999. Landstingslov nr. 12 af 29. oktober 1999 om fangst og jagt, §2 stk. 3. dk.nanoq.gl [accessed 20 February 2012]

Haaland, H., Skogen, K., Landa, A., Loeks, D., Andersen, O., Aastrup, P. et al. 2005. Uddelegering af forvaltningsansvar – levende ressourcer i Grønland. NINA Rapport 6.

Harremoës, P., Gee, D., MacGarvin, M., Stirling, A., Keys, J., Wynne, B. et al. (eds.) 2001. Late Lessons from Early Warn-ings: The Precautionary Principle 1896-2000. Environmental Issue Report 22. European Environment Agency, Copenhagen.

Haskell, S.P., Nielson, R.M., Ballard, W.B., Cronin, M.A. & McDonald, T.L. 2006. Dynamic responses of calving caribou to oilfields in northern Alaska. Arctic 59: 179-190.

Heiskanen, J., Nilsson, B., Mäki, A-H., Allard, A., Moen, J., Holm, S. et al. 2008. Aerial photo interpretation for change detection of treeline ecotones in the Swedish mountains. www.resgeom.slu.se/resana/NILS/Publikationer/arb_rapp_242.pdf [accessed 15 June 2011]

Helldin, J.O., Jung, J., Neumann, W., Olsson, M., Skarin, A. & Widemo, F. 2012. The impacts of wind power on terrestrial mammals: a synthesis. Report 6510. Swedish Environmental Protection Agency, Stockholm.

Henry, G. & Elmendorf, S. 2010. Greening of the Arctic. In: CAFF. Arctic Biodiversity Trends 2010: Selected Indicators of Change, pp 62-64. CAFF International Secretariat, Akureyri.

Hinzman, L.D., Bettez, N.D., Bolton, W.R., Chapin, F.S., Dy-urgerov, M.B., Fastie, C.L. et al. 2005. Evidence and implica-tions of recent climate change in Northern Alaska and other Arctic regions. Clim. Change 72: 251-298.

Hu, F.S., Higuera, P.E., Walsh, J.E., Chapman, W.L., Duffy, P.A., Brubaker, L.B. & Chipman, M.L. 2010. Tundra burning in Alaska: Linkages to climatic changes and sea ice retreat. Jour-nal of Geophysical Research 115: 1-8.

Hummel, M. & Ray, J. 2008. Caribou and the North: A Shared Future. Dundurn Press, Toronto.

Huntington, H. 2008. A Strategy for Facilitating and Promoting Community-Based Monitoring Approaches in Arctic Biodiver-sity Monitoring. CAFF CBMP Report No. 13.

Huntington, H.P. 2011. Arctic science – the local perspective. Nature 478: 182-183.

Huntington, H. & Fox, S. 2005. The changing Arctic: indigenous perspectives. In: ACIA. Arctic Climate Impact Assessment, pp 61-98. Cambridge University Press, Cambridge.

Huntington, H., Callaghan, T., Fox, S. & Krupnik, I. 2004. Match-ing traditional and scientific observations to detect environ-mental change: A discussion on Arctic terrestrial ecosystems. Ambio 13: 18-23.

Jernelöv, A. 2010. How to defend against future oil spills. Nature 466: 182-183.

John, E. & Turkington, R. 1997. A 5-year study of the effects of nutrient availability and herbivory on two boreal forest herbs. Journal of Ecology 85: 419-430.

Jokinen, M. (ed.) 2005. Poronhoidon ja suojelun vaikutukset Mal-lan luonnonpuistossa. Metsäntutkimuslaitoksen tiedonantoja [Finnish Forest Research Institute Research Papers] 941.

Jones, J.P.G., Andriamarovolona, M.M., Hockley, N.J., Gibbons, J.M. & Milner-Gulland, E.J. 2008. Testing the use of inter-views as a tool for monitoring trends in the harvesting of wild species. J. Appl. Ecol. 45: 1205-1212.

Jorgenson, M.T., Racine, C.H., Walters, J.C. & Osterkamp, T.E. 2001. Permafrost degradation and ecological changes associat-ed with a warming climate in central Alaska. Climate Change 48: 551-579.

Kevan, P.G., Forbes, B.C., Kevan, S.M. & Behan-Pelletier, V.M. 1995. Vehicle tracks on high Arctic tundra: their effects on the soil, vegetation, and soil arthropods. Journal of Applied Ecology 32: 655-667.

Key, S.L. & Stohl, A. 2007. Arctic air pollution: Origins and im-pacts. Science 315: 1537-1540.

Khitun, O. 1997. Self-recovery after technogenic and natural dis-turbances in the central part of the Yamal Peninsula (Western Siberian Arctic). In: R.M.M. Crawford (ed.). Disturbance and Recovery in Arctic Lands: An Ecological Perspective, pp 531-562. Kluwer, Dordrecht.

Khitun, O. & Rebristaya, O. 2002. Anthropogenic impacts on hab-itat structure and species richness in the West Siberian Arctic. In: A.E. Watson, L. Alessa & J. Sproull (eds.). Wilderness in the Circumpolar North, pp 85-95. USDA-FS, Rocky Mountain Research Station, Ogden.

Kumpula, T., Pajunen, A., Kaarlejärvi, E.M., Forbes, B.C. & Stammler, F. 2011. Land use and land cover change in arctic Russia: ecological and social implications of industrial develop-ment. Global Environmental Change 21: 550-562.

Page 23: Arctic Biodiversity Assessment - Amazon S3€¦ · 628 Arctic Biodiversity Assessment Species protection has focused on preventing overharvest, which has historically been the largest

650 Arctic Biodiversity Assessment

Kumpula, T., Forbes, B.C., Stammler, F. & Meschtyb, N. 2012. Dynamics of a coupled system: multi-resolution remote sens-ing in assessing social-ecological responses during 25 years of gas field development in Arctic Russia. Remote Sens. 4: 1046-1068.

Laidler, G.J. 2006. Inuit and scientific perspectives on the rela-tionship between sea ice and climate change: the ideal comple-ment? Climate Change 78: 407-444.

Larsen T. & Stirling, I. 2009. The Agreement on the Conservation of Polar Bears – its History and Future. Rapportserie Nr. 127, Norsk Polarinstitutt, Tromsø.

Larsen, T.S., Kurvits, T. & Kuznetsov, E. 2011. Lessons learned from ECORA – An integrated ecosystem management approach to conserve biodiversity and minimize habitat fragmentation in the Russian Arctic. CAFF Strategy Series Report No. 4.

Lassuy, D. & Lewis, P.N. 2010. Invasive species (human-induced). In: CAFF. Arctic Biodiversity Trends 2010: Selected Indicators of Change. CAFF International Secretariat, Akureyri.

Lawrence, D.M., Slater, A.G., Tomas, R.A., Holland, M.M. & Deser, C. 2008. Accelerated Arctic land warming and permafrost degradation during rapid sea ice loss. Geophysical Research Letters 35: L11506

Lemieux, C.J. & Scott, D.J. 2005. Climate change, biodiversity conservation and protected area planning in Canada. Canadian Geographer 49: 384-399.

Light, B., Grenfell, T.C. & Perovich, D.K. 2008. Transmission and absorption of solar radiation by Arctic sea ice during the melt season. Journal of Geophysical Research 113: 1-19.

Livingston, D. 2011. Circumpolar Protected Areas Monitoring. Arctic Protected Areas Monitoring Scheme Background Paper. CAFF International Secretariat, CAFF Monitoring Series Report Nr. 5.

Lloyd, A.H., Rupp, T.S., Fastie, C.L. & Starfield, A.M. 2003. Patterns and dynamics of treeline advance on the Seward Pen-insula, Alaska. Journal of Geophysical Research 108: 1-15.

Luzar, J.B., Silvius, K.M., Overman, H., Giery, S.T., Read, J.M. & Fragoso, J.M.V. 2011. Large-scale environmental monitor-ing by indigenous people. BioScience 61: 771-781.

Macias-Fauria, M., Forbes, B.C., Zetterberg, P. & Kumpula, T. 2012. Eurasian Arctic greening reveals teleconnections and the potential for structurally novel ecosystems. Nature Climate Change 2: 613-618.

Margesin, R. & Schinner, F. 1999. Biotechnological Applications of Cold-Adapted Organisms. Springer-Verlag, Heidelberg.

McGuire, A.D., Chapin, F.S. III, Walsh, J.E. & Wirth, C. 2006. Integrated regional changes in Arctic climate feedbacks: Implications for the global climate system. Annual Review of Environmental Resources 31: 61-91.

McGuire, A.D., Anderson, L.G., Christensen, T.R., Dallimore, S., Guo, L., Hayes, D.J. et al. 2009. Sensitivity of the carbon cycle in the Arctic to climate change. Ecological Monographs 79: 523-555.

McGuire, A.D., Chapin, F.S. III & Ruess, R.W. (eds.) 2010. Dynamics of change in Alaska’s boreal forests: resilience and vulnerability in response to climate warming. Can. J. For. Res. 40(7).

Meek, C.L. 2011. Putting the US polar bear debate into context: the disconnect between old policy and new problems. Journal of Marine Policy 35: 430-439.

Meltofte, H., Piersma, T., Boyd, H., McCaffery, B., Ganter, B., Golovnyuk, V.V. et al. 2007: Effects of climate variation on the breeding ecology of Arctic shorebirds. – Meddr. Grønland, Biosci. 59: 1-48.

Merkel, F. 2010. Evidence of recent population recovery in com-mon eiders breeding in Western Greenland. Journal of Wildlife Management 74: 1869-1874.

Meschtyb, N.A. 2008. Aleuts. In: L. Sillanpää (ed.). Awakening Siberia – From Marginalization to Self-Determination: the Small Indigenous Nations of Northern Russia on the Eve of the Millennium. Acta Politica No. 33. Department of Political Science, University of Helsinki, Helsinki.

Moller, H., Berkes, F., Lyver, P.O. & Kislaioglu, M. 2004. Com-bining science and traditional ecological knowledge: monitor-

ing populations for co-management. Ecology and Society 9(3), article no. 2.

Mustonen, T. 2009. Karhun väen ajast-aikojen avartuva avara. Tutkimus kolmen euraasialaisen luontaistalousyhteisön pai-kallisesta tiedosta pohjoisen ilmastonmuutoksen kehyksessä. University of Joensuu.

Mustonen, T. 2012. Metsäveri – Aslak Ola Aikion elämää ja tari-noita. Snowchange Cooperative.

Myers-Smith, I.H., Arnesen, B.K., Thompson, R.M. & Chapin, F.S. III. 2006. Cumulative impacts on Alaskan Arctic tundra of a quarter century of road dust. Écoscience 13: 503-510.

Myhre, G. 2009. Consistency between satellite-derived and modeled estimates of the direct aerosol effect. Science 325: 187-190.

National Research Council 2003. Cumulative Environmental Effects of Oil and Gas Activities on Alaska’s North Slope. Na-tional Academies Press, Washington.

Nellemann, C. Kullerud, L., Vistnes, I., Forbes, B.C., Husby, E., Kofinas, G.P. et al. 2001. GLOBIO Global methodology for mapping human impacts on the biosphere: the Arctic 2050 sce-nario and global application. UNEP/DEWA Technical Report 3. United Nations Environment Programme, Nairobi.

Nielsen, M.R. 2009. Is climate change causing the increasing nar-whal (Monodon monoceros) catches in Smith Sound, Greenland? Polar Research 28: 238-245.

Novikova,N.I. 2008. Eskimos. In: L. Sillanpää (ed.). Awakening Siberia – From Marginalization to Self-Determination: the Small Indigenous Nations of Northern Russia on the Eve of the Millennium. Acta Politica No. 33. Department of Political Science, University of Helsinki.

NSIDC (National Snow and Ice Data Center) 2010. nsidc.org [ac-cessed 12 January 2010]

NY Times 2011a. Failure in the Gulf. New York Times 7 January 2011: A22.

NY Times 2011b. The verdict on the spill. New York Times 11 January 2011: A22.

Olofsson, J., Oksanen, L., Callaghan, T., Hulme, P.E., Oksanen, T. & Suominen, O. 2009. Herbivores inhibit climate-driven shrub expansion on the tundra. Global Change Biology 15: 2681-2693.

Oozeva, C., Noongwook, C., Noongwook, G., Alowa, C. & Krupnik, I. 2004. Watching ice and weather our way. Arctic Studies Center, Smithsonian Institution, Washington, DC.

Pedersen, C.A., Roeckner, E., Lüthje, M. & Winther, J.-G. 2009. A new sea ice albedo scheme including melt ponds for ECHAM5 general circulation model. J. Geophys. Res. 114, D08101.

Perovich, D., Meier, W., Tschudi, M., Gerland, S. & Richter-Menge, J. 2012. Arctic report card: update for 2012: sea ice. www.arctic.noaa.gov/reportcard/sea_ice.html [accessed 7 December 2012]

Post, E. & Pedersen, C. 2008. Opposing plant community re-sponses to warming with and without herbivores. PNAS 105: 12353-12358.

Pulsifer, P.L., Laidler, G.J., Taylor, D.R.F. & Hayes, A. 2010. To-wards an Indigenist data management program: reflections on experiences developing an atlas of sea ice knowledge and use. Can. Geog. 55: 108-124.

Pulsifer, P., Gearheard, S., Huntington, H.P., Parsons,M.A., Mc-Neave, C. & McCann, H.S. 2012. The role of data management in engaging communities in Arctic research: overview of the Exchange for Local Observations and Knowledge of the Arctic (ELOKA). Polar Geography 35: 271-290.

Rasmussen, R.O, & Koroleva, N.E. (eds.) 2003. Social and Envi-ronmental Impacts in the North. Kluwer, Dordrecht.

Raundrup, K., Levermann, N. & Poulsen, M. 2012. Overlap in diet and distribution of two goose species suggests potential for competition at a common moulting area in West Green-land. Dansk. Orn. Foren. Tidsskr. 106: 93-100.

Regehr, E.V., Lunn, N.J., Amstrup, S.C. & Stirling, I. 2007. Effects of earlier sea ice breakup on survival and population size of polar bears in Western Hudson Bay Journal of Wildlife Management 71: 2673-2683.

Page 24: Arctic Biodiversity Assessment - Amazon S3€¦ · 628 Arctic Biodiversity Assessment Species protection has focused on preventing overharvest, which has historically been the largest

Chapter 19 • Disturbance, Feedbacks and Conservation 651

Regehr, E.V., Hunter, C.M., Caswell, H., Amstrup, S.C. & Stirling, I. 2010. Survival and breeding of polar bears in the Southern Beaufort Sea in relation to sea ice. Journal of Animal Ecology 79: 117-127.

Riordan, B., Verbyla, D. & McGuire, A.D. 2006. Shrinking ponds in subarctic Alaska based on 1950-2002 remotely sensed imag-es. J. Geophys. Res. 111: 1-11.

Rist, J., Milner-Gulland, E.J., Cowlishaw, G. & Rowcliffe, M. 2010. Hunter Reporting of Catch Per Unit Effort as a Moni-toring Tool in a Bushmeat-Harvesting System. Conserv. Biol. 24: 489-499.

Root, T. & Alpert, P. 1994. Volunteers and the NBS. Science 263: 1205.

Rouse, W.R., Oswald, C.J., Binyamin, J., Spence, C., Schertzer, W.M., Blanken, P.D. et al. 2005. The role of northern lakes in a regional energy balance. J. Hydrometeorology 6: 291-305.

Russell, D.E., Svoboda, M., Arokium, J. & Cooley, D. 2013. Arc-tic Borderlands Ecological Knowledge Cooperative: can local knowledge inform caribou management? Rangifer 33(21): 71-78.

Schuur, E.A.G, Bockheim, J., Canadell, J., Euskirchen, E.S., Field, C., Goryachkin, S. et al. 2008. The vulnerability of per-mafrost carbon to climate change: implications for the global carbon cycle. Bioscience 58: 701-714.

Scott, D.A. 1998. Global Overview of the Conservation of Arctic Migratory Breeding Birds outside the Arctic – 1998. Wetlands International Publication No. 45. CAFF Technical Report No. 4. CAFF, Akureyri.

Shindell, D. & Faluvegi, G. 2009 Climate response to regional radiative forcing during the twentieth century. Nature Geosci-ence 2: 294-300.

Smith, L.C., Sheng, Y., MacDonald, G.M. & Hinzman, L.D. 2005. Disappearing arctic lakes. Science 308: 1429.

Stammler, F. & Forbes, B.C. 2006. Oil and gas development in the Russian Arctic: West Siberia and Timan-Pechora. IWGIA Newsletter Indigenous Affairs, Arctic Oil and Gas Develop-ment 2-3/06: 48-57.

Stroeve, J., Holland, M.M., Meier, W., Scambos, T. & Serreze, M. 2007. Arctic sea ice decline: Faster than forecast. Geophysical Research Letters 34: 1-5.

Sturm, M., Racine, C.R. & Tape, K. 2001. Increasing shrub abun-dance in the Arctic, Nature 411: 546-547.

Sturm, M., Douglas, T., Racine, C. & Liston, G.E. 2005. Chang-ing snow and shrub conditions affect albedo with global impli-cations. Journal of Geophysical Research 110, G01004.

Tarnocai, C., Canadell, J.C., Schuur, E.A.G., Kuhry, P., Mazhi-tova, G. & Zimov, S. 2009. Soil organic carbon pools in the northern circumpolar permafrost region. Global Biogeochem-ical Cycles 23: 1-11.

van der Velden, M. 2010. Design for the contact zone. In: F. Sudweeks, H. Hrachovec & C. Ess (eds.) Proceedings Cultural Attitudes Towards Communications and Technology 2010, pp 1-18. Murdoch University, Australia.

Vistnes, I. & Nellemann, C. 2008. The matter of spatial and temporal scales: a review of reindeer and caribou response to human activity. Polar Biology 31: 399-407.

VNIPIGazdovycha 2005. Baseline for Investments in the Devel-opment of Bovanenkovo Deposit on the Yamal Peninsula and Gas Transport, Vols. 3 and 7 (in Russian). VNIPIGazdovycha, Saratov and VNIIGaz, Moscow.

Vongraven, D., Aars, J., Amstrup, S., Atkinson, S., Belikov, S., Born, E. et al. 2012. A circumpolar monitoring plan for polar bears. Ursus 23(2):1-66.

Walker, D.A. 1996. Disturbance and recovery of arctic Alaskan vegetation. Ecological Studies 120: 35-71.

Walter, K.M., Smith, L.C. & Chapin, F.S. III. 2007. Methane bub-bling from northern lakes: present and future contributions to the global methane budget. Philos. Trans. Royal Society 365: 1657-1676.

Wang, M. & Overland, J. 2009. A sea ice free summer Arctic within 30 years? Geophys. Res. Lett. 36: L07502.

Weatherhead, E., Gearheard, S. & Barry, R.G. 2010. Changes in weather persistence: Insight from Inuit knowledge. Global Environ. Change 20: 523-528.

Wiig, Ø., Aars, J. & Born, E.W. 2008. Effects of climate change on polar bears. ScienceProgress 91: 151-173.

Wild, R. & McLeod, C. (eds.) 2008. Sacred Natural Sites: Guide-lines for Protected Area Managers. IUCN, Gland.

Yoccoz, N.G., Nichols, J.D. & Boulinier, T. 2001. Monitoring of biological diversity in space and time. Trends Ecol. Evol. 16: 446-453.

Zhang, T., Frauenfeld, O.W., Serreze, M.C., Etringer, A., Oelke, C., McCreight, J. et al. 2005. Spatial and temporal variability in active layer thickness over the Russian Arctic drainage basin. Journal of Geophysical Research 110, D16101.

Zhuang, Q., Melillo, J.M., Sarofim, M.C., Kicklighter, D.W., McGuire, A., Felzer, B.S. et al. 2006. CO2 and CH4 exchanges between land ecosystems and the atmosphere in northern high latitudes over the 21st century. Geophysical Research Letters 33: L17403.

Zöckler, C., Hla, T.H., Clark, N. & Morozov, V. 2010. Spoon-billed Sandpiper. Recovery Team News Bulletin No 4 (May). www.shorebird-network.net/PDFs/Sbs-news-bulletin-May-2010.pdf [accessed 15 June 2011]


Recommended