Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January...

147
Ch12 - Fisheries ACIA Scientific Report External Review January 2004 12–1 DO NOT CIRCULATE OR CITE 13/01/04 Chapter 12. Fisheries Lead Authors: Hjálmar Vilhjálmsson and Alf Håkon Hoel Contributing Authors: Sveinn Agnarsson, Ragnar Arnason, James E. Carscadden, Arne Eide, David Fluharty, Geir Hønneland, Carsten Hvingel, Jakob Jakobsson, George Lilly, Odd Nakken, Vladimir Radchenko, Susanne Ramstad, William Schrank, Niels Vestergaard, and Thomas Wilderbuer 12.1 Introduction This chapter is concerned with identifying the possible effects of climate change on selected fish stocks and their fisheries in the Arctic. Arctic fisheries of selected species are described in the following regions: the northeast Atlantic consisting of the Barents and the Norwegian Seas; the waters around Iceland and Greenland; the waters off northeastern Canada; and the Bering Sea. The species discussed are: first, those few species, which are circumpolar (capelin, Greenland halibut, northern shrimp and polar cod); and, second, those additional species, which are of commercial importance in specific regions. These species include Atlantic cod, haddock, Alaska pollock, Pacific cod, snow crab and a number of other species. Because marine mammals play an important role in northern marine ecosystems and can be commercially important, they are also considered in this chapter. It is important to note that this chapter deals largely with the effects of climate change on commercial fisheries and the impacts on society as a whole. See Chapters 9, 10 and 11 on implications for indigenous peoples. Biological and model uncertainties/certainties It is not possible to provide precise forecasts of, either changes in the fish stocks 1 and fisheries or the effects on society. The sources of uncertainty can be conveniently grouped into three categories, namely, uncertainties in identifying the reasons for historical changes in fish biology, uncertainties in the predictions of possible changes in the ocean climate under the scenarios of climate change, and uncertainties relating to the socioeconomic effects of changes in fish stocks. There are many biological characteristics of fish that change due to natural variability of the physical environment. However, when fish stocks are heavily exploited, as many Arctic stocks have been, it has proven difficult to identify the relative importance of fishing and environment on observed changes in biology. Furthermore, many fish stocks are currently much lower in abundance than in the past, and exhibit extreme changes in population characteristics. Thus, even if historical observations of variability in fish biology could be associated with past changes in 1 There is an important distinction between fish stocks and fish species. A species is a subdivision of a genus, containing individuals that resemble one another and that may interbreed. A fish stock is a geographically distinct unit of a given species.

Transcript of Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January...

Page 1: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–1 DO NOT CIRCULATE OR CITE

13/01/04

Chapter 12. Fisheries

Lead Authors: Hjálmar Vilhjálmsson and Alf Håkon Hoel

Contributing Authors: Sveinn Agnarsson, Ragnar Arnason, James E. Carscadden, Arne Eide,David Fluharty, Geir Hønneland, Carsten Hvingel, Jakob Jakobsson, George Lilly, Odd Nakken,Vladimir Radchenko, Susanne Ramstad, William Schrank, Niels Vestergaard, and ThomasWilderbuer

12.1 Introduction

This chapter is concerned with identifying the possible effects of climate change on selected fishstocks and their fisheries in the Arctic. Arctic fisheries of selected species are described in thefollowing regions: the northeast Atlantic consisting of the Barents and the Norwegian Seas; thewaters around Iceland and Greenland; the waters off northeastern Canada; and the Bering Sea. Thespecies discussed are: first, those few species, which are circumpolar (capelin, Greenland halibut,northern shrimp and polar cod); and, second, those additional species, which are of commercialimportance in specific regions. These species include Atlantic cod, haddock, Alaska pollock,Pacific cod, snow crab and a number of other species. Because marine mammals play an importantrole in northern marine ecosystems and can be commercially important, they are also consideredin this chapter.

It is important to note that this chapter deals largely with the effects of climate change oncommercial fisheries and the impacts on society as a whole. See Chapters 9, 10 and 11 onimplications for indigenous peoples.

Biological and model uncertainties/certainties

It is not possible to provide precise forecasts of, either changes in the fish stocks1 and fisheries orthe effects on society. The sources of uncertainty can be conveniently grouped into threecategories, namely, uncertainties in identifying the reasons for historical changes in fish biology,uncertainties in the predictions of possible changes in the ocean climate under the scenarios ofclimate change, and uncertainties relating to the socioeconomic effects of changes in fish stocks.

There are many biological characteristics of fish that change due to natural variability of thephysical environment. However, when fish stocks are heavily exploited, as many Arctic stockshave been, it has proven difficult to identify the relative importance of fishing and environment onobserved changes in biology. Furthermore, many fish stocks are currently much lower inabundance than in the past, and exhibit extreme changes in population characteristics. Thus, evenif historical observations of variability in fish biology could be associated with past changes in

1 There is an important distinction between fish stocks and fish species. A species is a subdivision of a genus,containing individuals that resemble one another and that may interbreed. A fish stock is a geographically distinct unitof a given species.

Page 2: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–2 DO NOT CIRCULATE OR CITE

ocean climate, it is not known whether the present populations would respond in a manner similarto the historical response.

Some uncertainties surrounding the response of the ocean to the changes in global climate,resulting from the scenarios in this assessment, have been identified in Chapter 8. One of the mostimportant aspects of the Arctic climate is the thermohaline circulation. Possible changes in thisand its consequences cannot be evaluated with current models. In addition, the current climatemodels cannot assess changes in winds, with the result that there is considerable uncertainty in thechanges of stratification in the Arctic oceans. The current climate models can make predictions ona global scale but cannot provide reliable predictions of climate change at the regional level. Thisresults in uncertainty in the evaluation of the potential effects of climate change on the largemarine ecosystems considered in this chapter.

However, there are conclusions in Chapter 8, which reflect a high degree of certainty aboutchanges in the ocean climate in Arctic seas. Although Chapter 8 did not identify regional changes,that chapter concludes that “most Arctic seas will very likely experience a temperature increase”that will “very likely lead to a northward shift of many fish species, changes in the timing of theirmigration, possible extension of their feeding areas and increased growth rates”. Furthermore, theincrease in sea temperature will result in reductions of sea ice cover, which in turn will “verylikely enhance primary production, leading to likely increases in zooplankton production andpossibly increased fisheries production”. In addition, the reduction in sea ice cover will likely leadto an expansion of the feeding grounds of whales, presumably leading to an increase in theirnumbers. Thus, although these conclusions are global in scale and do not identify specific changesin the large marine ecosystems assessed in this chapter, the conclusions proffered by Chapter 8,with a high degree of probability, provide a basis for this chapter to consider those conclusions inthe context of the fish stocks, fisheries and possible effects on human societies by anticipatedchanges of the large marine ecosystems considered.

Societal uncertaintiesOnce the fish population changes have been evaluated, it becomes necessary to relate thosechanges to changes in society. This raises new difficulties. Even where the changes in fishpopulations are predictable to a high degree of accuracy, there is no deterministic relationshipbetween these changes and those in society. Social change is driven by a number of differentforces, where climate change is but one of a number of natural factors. In addition, humans arealso important drivers of change, through economic and political activities. It is extremely difficultto isolate the relative impact of the various drivers of change. In addition, societies have thecapacity to adapt to change. Changes in fish stocks, for example, are met by adjustments infisheries management practices and the way fisheries are performed.

The result of these uncertainties is that in this chapter there are few concrete predictions. Instead,changes of potential effects and likely outcomes are considered.

The global framework for managing living marine resourcesOver recent decades, a global framework for the management of living marine resources has beendeveloped, providing coastal states with extended jurisdiction over natural resources. The 1973-82United Nations Law of the Sea Conference (UNCLOS) resulted in the 1982 Law of the Sea

Page 3: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–3 DO NOT CIRCULATE OR CITE

Convention, which lays down the rules and principles for the use and management of the naturalresources in the ocean. The most important elements here are the provisions that enable coastalstates to establish Exclusive Economic Zones (EEZs) up to 200 nautical miles (360 kilometres)from their coastal baselines. In the EEZs coastal states have sovereign rights over the naturalresources. The Convention also mandates that coastal states manage resources in a sustainablemanner and that they be used optimally. Where fish stocks are shared among countries, they shallseek to cooperate on their management.

A country’s authority to manage fish stocks is defined by its 200 mile Exclusive Economic Zone(EEZ). Within its EEZ, a coastal state has sovereign rights over the natural resources, andtherefore the authority to manage the living marine resources there.

During the 1980’s it became evident that the framework provided by the Law of the SeaConvention was inadequate to cope with two major developments in fisheries worldwide: i.e. thedramatic increase in fishing at the high seas beyond the EEZs; and a corresponding increase incatches within the EEZs. Both developments were driven by rapidly growing fishing capacity. Theconsequence was that many stocks were overfished. Therefore, a treaty was negotiated under UNauspices to supplement the 1982 Convention, seeking to provide a legal basis for restrictingfisheries on the high seas and introducing more restrictive management principles, enhancedinternational cooperation in management, and improved enforcement of management measures.The 1995 UN Convention on Straddling Fish Stocks and Highly Migratory Fish Stocks (The UNFish Stocks Agreement) thus mandates the application of a precautionary approach to fisheriesmanagement. It also emphasises the need for cooperation between countries at a regional level inthis respect. These two elements have proved crucial in the development of international fisheriesconservation and management policies since the mid-1990s, not least in the Arctic region. Existingregional arrangements there have been improved upon, in order to implement the agreements. Thisapplies to the Northwest Atlantic Fisheries Organization (NAFO), which covers the NorthwestAtlantic, and the Northeast Atlantic Fisheries Commission (NEAFC), which covers theinternational waters in the Northeast Atlantic. An agreement placing a moratorium on fishing onthe high seas in the Bering Sea has been in force since 1992.

The development of this global framework for fisheries management has been accompanied by acorresponding development of fisheries management regimes in individual countries. The actualdesign and performance of such regimes are crucial to the fate of fish stocks. At the global level,the major challenges to fisheries management are related to the need to reduce a substantialovercapacity in the world’s fish fleets, and the need to introduce more sustainable managementpractices. To achieve the latter, countries are introducing precautionary approaches to fisheriesmanagement – a crucial requirement of the 1995 Fish Stocks Agreement. In addition, ecosystem-based approaches to the management of living marine resources, where natural factors as climatechange are taken into account in decision-making, are in the process of being developed. The 2002World Summit on Sustainable Development stated in its implementation plan that such eco-system based approaches to management are to be in place by 2010.

All Arctic countries with significant fisheries have well established resource management regimeswith comprehensive systems for producing the knowledge base required for management, thepromulgation of regulations to govern fishing activities, and arrangements to ensure compliance

Page 4: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–4 DO NOT CIRCULATE OR CITE

with regulations. While the various regimes vary considerably with regard to the actual design ofmanagement policies, the challenges they confront in attempting to reduce overcapacity and inintroducing precautionary approaches to fisheries are similar.

For marine mammals there exists one international body at the global scale, and several regionalones. At the global scale the 1946 International Convention for the Regulations of Whalingmandates an International Whaling Commission (IWC) to regulate the harvest of great whales. Amoratorium on commercial whaling was adopted in 1982. A number of countries, among themNorway and Russia, availed themselves of their right under the convention not to be bound by thisdecision. Canada and Iceland left the Commission due to the preservationist developments there.Iceland later rejoined the Commission in 2003. The North Atlantic Marine Mammals Commission(NAMMCO) is tasked with the management of marine mammals in the North Atlantic.

Chapter organizationMost of the remainder of the chapter is presented in sections relating to each of the four regionsdescribed above. The discussion of each region follows a standard format:

1) Introduction2) Ecosystem essentials3) Fish stocks and fisheries4) Past climatic variations and their impact on commercial stocks5) Possible impacts of global warming on fish stocks6) The economic and social importance of fisheries7) Past economic and social impacts of climate change on fisheries8) Economic and social impacts of global warming: possible scenarios9) Ability to cope with change10) References

The chapter concludes with two sections. One presents a synthesis of the regional assessments ofthe impacts of climate change on Arctic fisheries and societies and the second presents researchrecommendations.

Page 5: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–5 DO NOT CIRCULATE OR CITE

12.2 The Northeast Atlantic

12.2.1 IntroductionThis section addresses potential impacts from climate change on the fisheries in the Arctic area inthe Northeast Atlantic.2 The area comprises the northern and eastern parts of the Norwegian Sea tothe south, and the North Norwegian and Northwest Russian coasts and the Barents Sea to the eastand North. The fisheries in this region take place in areas under Norwegian and Russianjurisdictions as well as in international waters. The total fisheries in the area amounted to some 2.1million t in 2001.3 Aquaculture production is dominated by salmon and trout and amounted to 86000 t in 2001 (Fiskeridirektoratet 2002a).

The legal and political setting of the fisheries in the Northeast Atlantic is complex. Norway andRussia established 200 nautical mile Exclusive Economic Zones (EEZs) in 1977, as a consequenceof developments in international ocean law at the time. The waters around Svalbard come under aFisheries Conservation Zone set up by Norway, which according to the 1920 Svalbard Treatyholds sovereignty over the Svalbard archipelago. The waters around the Norwegian island of JanMayen, north of Iceland, are covered by a Fisheries Zone. Two areas are on the high seas beyondthe EEZs: in the Barents Sea a so called “Loophole” and in the Norwegian Sea a “Herring hole”(Fig. 12.2.1). Norway and Russia have long traditions for cooperating both in trade andmanagement issues. In the 18th century, Norwegian fishermen in the north traded cod forcommodities from Russian vessels - the so-called "Pomor-trade" (Berg 1995). Joint managementof the Barents Sea fish stocks havs been negotiated since 1975. Since then, a comprehensiveframework for managing the living marine resources in the area has been developed, including thehigh seas. The resources in the area are exploited with vessels from Norway and Russia, as well asfrom third countries.

Northern Norway includes three counties: Finnmark, Troms and Nordland, and covers an area of110 000 square kilometers - about the same size as Great Britain. The total population is 460 000(Norwegian Central Statistical Bureau, 2001). Due to the influence of the North Atlantic Current,the climate in this region is several degrees warmer than the average in areas at the same latitude.While the Norwegian fishing industry is found in numerous communities all along the northerncoast, the northwest Russian fishing fleet is concentrated in large cities, primarily in Murmansk. Inaddition to the Murmansk Oblast, Russia's ”northern fishery basin” comprises ArkhangelskOblast, the Republic of Karelia and Nenets Autonomous Okrug. There is no significantcommercial fishing activity east of these until one reaches the far eastern fishery basin in the NorthPacific. As of 1 January 2002, the population in the four federal subjects constituting Russia’snorthern fishery basin was 3,2 million people.

Insert Figure 12.2.1 here. Map of the Northeast Atlantic (Change ??)

2 See chapter 1 of this report for definitions of the Arctic.3 This figure is arrived at by using information in Michalsen, K. (ed.) 2003: Fisken og Havet, særnummer 1.

Page 6: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–6 DO NOT CIRCULATE OR CITE

12.2.2 Ecosystem essentialsAs described in chapter 8, seasonal variations in the hydrographic conditions of the Barents Seaare large in the upper water layers. The spring bloom starts in the southwestern areas andcontinues north- and eastward along with the retraction of ice as it melts. Fish and mammals havesimilarly directed migrations; spawning migrations south- and westwards in late autumn andwinter, and feeding migrations north- and eastwards in late spring and summer.

Relatively few species and stocks make up the bulk of the biomass at the various trophic levels.Fifteen to twenty species of whales and seals forage regularly in the area. Harp seals (Phocagroenlandica) and minke whales (Balaenoptera acutorostrata) are the two most importantpredators in the pelagic ecosystem. The stock of harp seals has its breeding areas in thesoutheastern parts of the Barents Sea, i.e. in the White Sea, and feed close to the ice edge, mainlyon amphipods and capelin. In periods of low capelin (Mallotus villosus) abundance, harp sealsfeed on other fish, like cod (Gadus morhua), haddock (Melanogrammus aeglefinus) and saithe(Pollachius virens), and migrate southwards along the Norwegian coast (Nilssen 1995). Minkewhales feed on various species of fish in most of the area from May to September (Nordøy et al.1995). During the winter the whales inhabit areas further south in the Atlantic Ocean.

The spawning grounds of most species are situated along the coast of Norway and Russia.Normally spawning takes place in winter and spring (February to May) and egg and larval driftroutes are towards the north and east. Juveniles and adults feed in the area; polar cod (Boreogadussaida) in the north- and northeasternmost parts, saithe and herring (Clupea harengus) in thesouthwest, as well as the easternmost Norwegian Sea and off the Norwegian coast. Capelin residemainly on the Atlantic side of the Polar Front during winter, but utilize the zooplanktonproduction in the large Arctic but ice-free areas north of the front in summer and autumn. Cod hasthe most extensive distribution. Adult cod spawn in Atlantic water far south along the coast ofNorway in March to April, feed along the Polar Front and even far into Arctic water masses duringsummer and autumn. All species show seasonal migrations, which coincide with the formation andmelting of ice; north- and eastwards during spring and summer, south- and westward duringautumn and winter.

Cod, saithe, haddock (the gadoids) and redfish (Sebastes marinus and S. mentella) have their mainspawning fields on the coastal banks and off the shelf edge (redfish) of Norway between 62°N and70°N and return to the Barents Sea area after spawning. Herring migrate out of the Barents Seabefore maturing, feed as adults in the Norwegian Sea and have their main spawning fields farthersouth along the Norwegian coast, between about 59°N and 68°N. Capelin spawn in the northerncoastal waters mainly between 20°E and 35°E, while polar cod has two main spawning areas; onein Russian waters in the southeastern part of the Barents Sea and another in the northwest, close tothe Svalbard archipelago. The capelin spawning schools are followed by predating immature cod,4-6 years old. Adult Greenland halibut (Hippoglossus reinhardtii) inhabit the slope waters atdepths between 400 and 1000 m in the entire area. Northern shrimp, (Pandalus borealis), is foundin most of the area with bottom depths between 100 and 700 m on the ”warm” side of the PolarFront. Individuals are 4-7 years of age when they change sex from male to female and spawning(hatching of eggs) occurs in summer - autumn in most of the area.

Page 7: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–7 DO NOT CIRCULATE OR CITE

From simulations of interactions between the species capelin, herring, cod, harp seal and minkewhale, Bogstad et.al. (1997) found that the stock of herring is sensitive to changes in the minkewhale abundance because whale predation in the Barents Sea affects the number of recruits to themature stock of herring. They also found that an increasing stock of harp seal will reduce thecapelin and cod stocks, so that an unexploited seal population would lead to a substantial loss ofcatch in the cod fishery.

Cod, capelin and herring are considered key fish species in the ecosystem and interactionsbetween them generate changes, which also largely affect other fish stocks as well as mammalsand birds (Bogstad et.al., 1997). Recruitment of cod and herring is enhanced by inflows ofAtlantic water carrying large amounts of suitable food (especially the “redfeed“ copepod Calanusfinnmarchius) for larvae and fry of these species. Consequently, survival increases, so thatjuvenile cod and herring become abundant in the area. However, since young and juvenile herringprey on capelin larvae in addition to zooplankton, capelin recruitment might be negatively affectedand thus cause a temporal decline in the capelin stock, a matter that affects most species in thearea (fish, birds and mammals) since capelin is their main forage fish. Predators will then prey onother small sized fish and shrimps. In particular, cod cannibalism may increase and thus affectfuture recruitment of cod to the fishery (Hamre, 2003).

It has become evident that in periods of low abundance or absence of capelin and/or herring, thetop predators will have to feed somewhere else or shift to prey in the zooplankton group. For codsuch shifts have been observed twice during the past 15 years and were related to collapses of thecapelin stock in 1986-1988 and 1993-1994.

12.2.3 Fish stocks and fisheriesFor the past thousand years, fishing for cod and herring has been important for coastalcommunities in Norway and northern Russia (Solhaug 1976). Throughout the centuries fishingwas purely coastal and seasonal and based on the large amounts of adult cod and herring migratinginto near shore waters for spawning during winter-spring and on the schools of immature codfeeding on spawning capelin along the northern coasts in April to June. A certain developmenttowards offshore fishing took place at end of the 19th century when cod were caught on theSvalbard banks and driftnetting of herring was initiated off northern Iceland. However, thequantities caught in these “offshore” fisheries were small compared to the near-shore catches inthe traditional fisheries for both species. Estimates of annual yields of cod and herring prior to1900 were given by Øiestad (1994). For both species large fluctuations were experienced. Thedominant feature is the 5-10 fold increases between 1820 and 1880 as compared to yield levels ofprevious centuries. For fish species other than cod and herring reliable estimates of yield prior tothe 20th Century are not available.

Landing statistics for herring, capelin, polar cod, Greenland halibut, northern shrimp and Arcticcod in the 20th Century are shown in Figure 12.2.2.Change)) Total fish landings from the areaincreased from about 0.5 million t at the beginning of the century to about 3 million t in the 1970s.This increase was caused mainly by a series of major technological improvements of fishingvessels and gear, including electronic instruments for fish finding and positioning, which tookplace during the 20th century and dramatically enhanched the effectiveness of the fishing fleet.Furthermore, there was as a growing market demand for fish products.

Page 8: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–8 DO NOT CIRCULATE OR CITE

Fisheries of Arctic speciesWhen herring became scarce in the late 1960s the purse seine fleet targeted capelin and catchesincreased rapidly in the 1970s. Management measures such as minimum allowable catch size andclosing of areas where undersized fish occurred, as well as limited fishing seasons, wereintroduced in the early 1970s, first by Norway and later jointly by Norway and Russia. Totalallowable catches (TACs) have been enforced since 1978. Landings have fluctuated widely. In2002 the total catch of capelin was 628 000 t (see Fig. 12.2.2). During the 1980s the importance ofcapelin and juvenile herring as food sources for cod and other predators was fully realized (seeNakken, 1994 for references). As a consequence, there was increased effort in research on speciesinteractions and since 1990 the cod stock’s need for capelin as food has been taken into account inthe scientific advice on management measures.

Russia and Norway started regular fisheries with bottom and pelagic trawls for polar cod in thelate 1960s. The catches increased to approximately 350 000 t in 1971. The Norwegian fleet wasactive until 1973, when the fishermen lost interest because of declining catches. Since thenlandings have been exclusively Russian. Catches in 2001 were about 40 000 t.

Until the early 1960s, the fishery of Greenland halibut (see Fig. 12.2.2) was mainly pursued bycoastal long liners off the coast of northern Norway. Annual landings were about 3 000 t. Aninternational trawl fishery developed in the areas between 72°N and 79°N and catches increased toabout 80 000 t in the early 1970s. Landings decreased throughout the 1970s, the spawning stockbiomass declined from more than 200 000 t in 1970 to about 40 000 t in the early 1990s and hassince remained at this low level. Since 1992 only vessels less than 28 m in length using long linesor gillnets have been permitted to carry out a directed fishery. The rest of the fishing fleet has beenrestricted by by-catch rules. The total catch in 2002 was 13 000 t.

Prior to 1970, trawling for northern shrimp took place in the fjords of northern Norway andcatches were low. During the 1970s offshore grounds became exploited. Catches increased until1984 when 128 thousand t were landed. Since then, catch levels have fluctuated (see Fig. 12.2.2).Fisheries have been regulated by by-catch rules and closed areas since the mid 1980s. Areas areclosed to fishing when the catch rates of young cod, haddock and Greenland halibut exceed acertain limit. In later years, young redfish has also been included in the by-catch quota. Areas arealso closed when the proportion of minimum-size shrimp (15 mm carapace length) is too high. Inthe Russian EEZ an annual TAC is also enforced. Estimated cod consumption of shrimp has since1992 been approximately 10 times higher than the landings, which were about 58 000 t. in 2001.

Other important fisheries in the regionThe two most important fisheries in the region are those for cod and herring. Until the 1950s,herring fisheries remained largely seasonal and near shore. The bulk of the landings came fromNorwegian vessels. In the 1950s Russian fishers developed a gillnet fishery in offshore waters inthe Norwegian Sea, and in the early 1960s purse seiners started using echo sounding equipment tolocate herring. These technological developments resulted in a large increase of the total catchesuntil 1966 (2 million t). Thereafter catches decreased rapidly and the stock collapsed. Althoughindividual scientists expressed concern about the stock, effective management measures wereneither advised nor implemented until after the stock had collapsed completely. Minor catches in

Page 9: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–9 DO NOT CIRCULATE OR CITE

the early 1970s (7-20 000 t) removed most of the remaining spawning stock as well as juvenilesand it was not until 1975 that the fishing pressure was brought to a level which permitted the stockto start recovering. For 25 years the stock was very small and remained in Norwegian coastalwaters throughout the year. Norway introduced management measures including minimumallowable landing size and and annual TACs. Furthermore, a complete ban on fishing herring wasenforced for some years. During the 1990s the stock recovered, started to make feeding migrationsinto the Norwegian Sea and catch quotas and landings increased. In 2002 the total landings were830 000 t.

Prior to 1920, the bulk of the cod catch was from two large seasonal and coastal fisheries: thefishery for immature cod feeding on spawning capelin along the northern coast of Norway andRussia and the fishery for spawning cod (“skrei“) further south off Northern Norway (the Lofotenfishery). In the 1920s and 1930s an international bottom trawl fishery targeting cod as well asother species (haddock, redfish) developed in offshore areas in the Barents Sea and off Svalbard.Annual catches increased from about 400 000 t in 1930 to 700-800 000 t at the end of the decade.Landings also remained high after World War II until the end of the 1970s when catches declinedsharply due to reduced stock size and the introduction of EEZs. Management advice was given bythe International Council for the Exploration of the Sea (ICES) from the early 1960s. Increases oftrawl mesh sizes were recommended in 1961 and in 1965 a variety of further conservationmeasures were recommended in order to increase yield per recruit and limit the overall fishingmortality. From 1969 onwards, ICES has expressed concern about the future size of the spawningstock, considering that at low levels of spawning stock biomass there would be an increased riskof poor recruitment to the stock. The first TAC for cod was set in 1975, but was far too high.Although minimum mesh size regulations had been in force for some years at that stage, it seemsfair to conclude that no effective management measures were in operation for demersal fish in thearea prior to the establishment of the EEZs in 1977.

The estimated average fishing mortality for the 5 year period 1997-2001 is a record high (0.90)and about twice the fishing mortality corresponding to the precautionary approach (0.42). Since1998 the spawning stock biomass has been well below the recommended precautionary level of500 000 t. Recruitment to the stock has been low in most of the recent years. Landings have variedconsiderably over time and in 2002 stood at 430 000 t (Fig. 12.2.2).

Insert Figure 12.2.2 here. Landings from the most important commercial fish stocks in theNortheast Atlantic

Marine mammalsThree species of marine mammals are commercially exploited in the Northeast Atlantic byNorwegian and Russian fishers, i.e. minke whales, hooded seals (Cystophora cristata) and harpseals. In addition, coastal seals (grey (Halichoerus grypus) and harbour (Phoca vitulina) seals areexploited along the Norwegian coast by local hunters. Offshore exploitation of marine mammalsin the area began in the sixteenth century. Basque and later Dutch and British vessels hunted rightwhales (Balaena mysticetus) and seals. Processing plants were established at shore stations as farnorth as northwestern Spitzbergen (Arlov 1996). Russian and Norwegian hunters have caughtwalrus (Odobenus rosmarus), polar bear (Thalarctos maritimus) and seals at the Svalbardarchipelago since the 16th Century. By the first decades of the 19th Century the stocks of right

Page 10: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–10 DO NOT CIRCULATE OR CITE

whales had all but disappeared, and the walrus was so depleted that the hunt became unprofitable.A new era of offshore exploitation began around 1860-1870 when the use of smaller ice-goingvessels (“sealers”) permitted Norwegian hunters to penetrate into the drift ice. At about the sametime the invention of the grenade harpoon made hunting of great whales profitable. Catches ofgreat whales increased between 1870 and 1900, but leveled off and decreased rapidly during thefirst decade of the 20th Century.

Minke whales have been hunted in landlocked bays (“whaling bays”) along the coast of Norwaysince olden times. Offshore hunting, using small motorized vessels, developed prior to WorldWar II, essentially as an extension of fishing activities. Catches increased until the 1950s, themean annual take in that decade being about 2 300 animals. Since 1960 catches have decreaseddue to reductions of annual TACs. Between 1987 and 1992 no commercial hunting was allowed.In recent years annual catches have been 400-600 animals and the quota for 2002 is 674 minkewhales. The stock in the area is estimated at 112 000 animals (Michalsen 2003).

Two stocks of harp seal , in the West Ice (Greenland Sea) and in the East Ice (White Sea - BarentsSea), and one stock of hooded seal (Cystophora cristata) in the West-Ice are subject to offshoresealing; since about 1880 mainly by Norwegian and Russian hunters. The total annual catch fromthese stocks increased from about 120 000 animals around 1900 to an average of about 350 000per year in the 1920s. Since then catches have declined, mainly because of catch regulations(TACs). In recent years the loss of markets has, however, been the main limiting factor. In the1990s catches of harp seal in the West-Ice were 8-10 000 animals each year and 8-9 000 forhooded seal, while catches of harp seal in the East-Ice ranged from 14 000 to 42 000 per annum.Russian catches, which constitute about 82% of the total, are taken in the East-Ice, while theNorwegian catches (about 18%) are taken in both the West- and East-Ice.

Hooded seals are found in the north Atlantic between Novaja Semlja, Svalbard, Jan Mayen,Greenland and Labrador. All the Norwegian catch of hooded seal takes place in the West-Ice(Greenland Sea). Russia has not caught hooded seals since 1995. The total catch in 2001 was 3820 animals. All seal stocks are assessed every second year by a joint ICES/NAFO workinggroup, which provides ICES with sufficient information to give advice on stock status and catchpotential. All three stocks are well within safe biological limits, and harvesting rates aresustainable.

12.2.4 Past climate variations and their impact on commercial stocksThe relationship between the physical effects of climate change and effects on the ecosystem iscomplex. It is not possible to isolate, let alone quantify, the effects of climate change on biologicalresources. The following discussion is therefore of a tentative and qualitative nature.

A number of climate-related events have been observed in the Northeast Atlantic fisheries (seechapter 8.3.3). During the warming of the Nordic Seas between 1900 and 1940, substantialnorthward dislocations of geographical boundaries took place for a range of marine species fromplankton to commercial fish, as well as for terrestrial mammals and birds (Dickson, 1992).Recruitment of both cod and herring is positively related to inflows of Atlantic waters to the areaand thus to temperature changes. Both stocks increased significantly between 1920 and 1940 whenwater temperatures increased (Hylen, 2002; Toresen and Østvedt, 2000). The increase in stock size

Page 11: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–11 DO NOT CIRCULATE OR CITE

was probably an effect of enhanced recruitment, because catches increased in the same period. Asimilar development may have occurred between 1800 and 1870 (Øiestad, 1994). Øiestad (1994)also provided evidence that the abundance of cod was low during the cold years 1650-1750.

Since World War II both cod and herring have been subject to overfishing. This resulted in acollapse of the herring stock in the 1960s, with serious consequences for other inhabitants of theecosystem as well as man. For cod the most likely result of overfishing is a far lower averageannual yield during the past two decades (1980-2000) than the stock has potential to produce.Recruitment of cod depends heavily on parent stock size in addition to environmental factors(Ottersen and Sundby, 1995; Pope et.al., 2001). For several decades a heavy fishing pressure hasprevented maintenance of the cod spawning stock at a level, which optimizes recruitment levelsthe long run. Therefore, the management of these stocks is the key-issue when it comes toassessing the effects of potential climate variations on fish-stocks (Eide & Heen 2002).

Textbox 12.2.1 The fall and rise of the Norwegian spring spawning herring

In the beginning of the 1950s the spawning stock of Norwegian spring-spawning herring wasestimated at 14 million t - one of the largest fish stocks in the world. Most of the adult stockmigrated between Norwegian and Icelandic coastal waters to spawn in winter and feed in summer,respectively. The herring-fishery was important for several countries, especially Norway, Iceland,Russia and the Faroes. However, after 15 years of over-exploitation and a decreasing spawningstock, this herring stock collapsed in the late 1960s.

Deteriorating climatic conditions north of Iceland and in the western Norwegian Sea are crucial inexplaining changes of feeding areas and migration routes of these herring in the late 1960s. Highfishing intensity was, however, the major factor behind the actual stock collapse. The breakdownhad large social and economic consequences for those depending on this fishery. Nevertheless, theindustry managed to redirect its effort to other pelagic species - primarily capelin.

In the following decades, what was left of these herring kept close to the Norwegian coast. Thestock was strictly regulated and fishing was prohibited for several years. These regulations,probably in combination with favorable climatic conditions, contributed to a considerable increasein stock size from the mid-1980s, making it possible to start fishing again. By the late 1980s thespawning stock had reached a level of 3-4 million t, in the main due to an above averagerecruitment by the 1983 year class.

By 1995 the spawning stock had reached 5 million t. As a consequence, the stock extended itsfeeding grounds by resuming its old migration pattern westward into the Norwegian Sea. Ittherefore became available for fishing beyond areas under Norwegian jurisdiction. The unilateralNorwegian management regime was no longer adequate to regulate fishing of the stock.Meanwhile, there was no arrangement in place to oversee the international management of thefishery. Negotiations between Norway, Russia, Iceland and the Faroes failed, and the total catchquota recommended by ICES was exceeded in the following year.

Page 12: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–12 DO NOT CIRCULATE OR CITE

High economic values were at stake for all actors. Fishermen and fisheries managers in allinvolved countries and in the EU were highly engaged in the conflict. A first agreement wasreached between Norway, Russia, Iceland and the Faroes in May 1996. In December 1996, the EUwas also included in the arrangement, which gives the five parties rights to set and distributeTACs of Norwegian spring spawning herring, based on an ICES advice. The formal responsibilityto manage the share of the stock in international waters is vested with the Northeast AtlanticFisheries Organization (NEAFC), where all the mentioned parties are members. Negotiations areheld every year, but the percentage allocation key has not changed since the 1996 agreement.However, changes in the migration pattern may upset the present arrangement.

This example shows that not only negative, but also positive changes in stock abundance maycreate management problems. If the participants had not reached an agreement, there would havebeen devastating consequences for the exploitation and development of the Norwegian springspawning herring stock, almost certainly resulting in large economical losses. Thus, this examplealso shows the importance of political efforts to solve these kinds of conflicts. Such politicaldialogues may become even more important if climate change results in changes in migrationpatterns of fish stocks in general.

12.2.5 Possible impacts of climate variations on fish stocksGlobal models show an increase of surface temperature in the area of 3-5°C by 2070. Regionalmodeling has however demonstrated that for surface temperatures in this area, there will be “…acooling of between 0 and –1 °C. ” (Furevik et al 2002) during the first two decades. By 2050 thearea will have become warmer according to this model, and by 2070 surface temperatures willhave increased by 1-2 °C.

Research during the past two-three decades has revealed that production of cod increases withincreasing temperature for stocks inhabiting areas of mean annual temperature below 6-7ºC, whilecod stocks in warmer waters suffer reduced recruitment when the temperature increases (Sundby2000). The mean annual ambient temperature for northeast Arctic cod is 2-4 ºC (depending onage group) and the stock has experienced greatly improved recruitment during periods of hightemperatures in the past (Sundby 2000). A rise in mean annual temperature in the Barents Sea overthe next 7-8 decades may therefore favor cod recruitment and production, and extend itsdistribution area (spawning and feeding) towards north and east. A similar statement may be madefor herring (see chapter 8). This statement is based on an assumption that the production anddistribution of animals at lower trophic levels (for copepods in particular - the food for larvae)remain unchanged. The prediction is also based on the assumption that harvest rates are kept atlevels that maintain spawning stock biomass above the level where recruitment is adverselyaffected.

Experience indicates that it is likely that a rise in temperature as predicted for the area, will causelarge displacements towards north and east of the distribution areas of resident marine organisms,including fish, shrimps and mammals. Their boundaries will be extended as waters get warmerand ice melts. “Warm water” pelagic species, such as blue whiting and mackerel will occur in thearea in higher concentrations and more regularly than in the past. Eventually these species mayinhabit the southwestern parts of the present “Arctic area” on a permanent basis.

Page 13: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–13 DO NOT CIRCULATE OR CITE

The effects of a temperature rise on the production by the stocks of fish and mammals, presentlyinhabiting the area, are more uncertain. This depends on how a temperature increase isaccompanied by changes in oceanic circulation patterns and plankton transports and production. Inthe past, recruitment to several fish stocks in the area, cod and herring in particular, has shown apositive correlation with increasing temperature. The increase in recruitment has however beendue to higher survival of larvae and fry, which in turn stems from increased availability of food.Food is transported into the area with inflows of Atlantic water masses, which also have causedthe ocean temperature to increase. Hence, high recruitment in fish is associated with hightemperature but not caused by the high temperature (Sundby 2000).

Provided that the fluctuations in Atlantic inflows to the area are maintained along with a generalwarming of the North Atlantic waters, it is to be expected that annual average recruitment ofherring and cod will be about the long-term average during the first 2-3 decades of the 21st

Century. This prediction is also based on the assumption that harvest rates are kept at levels thatmaintain spawning stocks well above the level at which recruitment is impaired. How productionwill change further into future is impossible to guess, since the predicted temperatures, particularlyin some of the global models, are so high that species composition and thus the interactions in theecosystem of this area may change completely.

12.2.6 Economic and social importance of the fisheries in the Northeast Atlantic4

The fishery sector is of considerable economic significance in Norway, being among the country’smain export earners. In 2001, the export of fish products accounted for 14 per cent of the totalexports from mainland Norway.5 The fisheries constituted 1.5 per cent of the Norwegian GNP(1999), excluding petroleum. In northwest Russia, the fisheries are of less economic importance inthe broader picture. A substantial share of the catches taken in Russian fisheries in the north islanded abroad.

Most northern coastal communities are heavily dependent on the fisheries in economic terms, aswell as culturally and historically attached to fisheries. As early as year 1000 an extensive trade indried cod developed in Northern Norway, through the Hanseatic trade (Solhaug 1983). The coastalfishery and trade made up the economic foundation for the communities along the northern coast.In the last two decades, aquaculture has become increasingly important, accounting for asignificant part of the economic value produced by the fisheries sector (Ervik et al. 2003).

The total fishery in the Arctic part of the Northeast Atlantic yields about 2.1 million tonnes. Thetotal annual value of these fisheries is in the order of 2 billion $US.6 The resources found in theArctic areas are significant also to fishery communities elsewhere. A substantial part of the

4 The figures in this section are based on a statistics from “Fisken og Havet” (various years) and The Directorate of Fisheries, and includeslandings from catches taken in the ICES statistical areas I and IIa and IIb.5 Total fish exports were 31 bn NOK in 2001, while total exports from mainland Norway stood at 216 bn NOK. Informationfrom tables 259 in Statistical Yearbook of Norway (www.ssb.no) and information from the Norwegian Seafood ExportsCouncil (www.seafood.no).6 This figure is arrived at by adding the fishery of cod, haddock, capelin, Shrimp, Norwegian Spring Spawning Herring,Saithe, Greenland halibut and redfish in the ICES subarea I, IIa and IIb, and using a rough average of the first hand values setby the Norwegian Raw Fish Sales Association and the Norwegian Herring Sales Association.

Page 14: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–14 DO NOT CIRCULATE OR CITE

catches in the North are taken by fishermen from outside the region - from southern Norway andelsewhere in Europe.

Fish stocks and fisheriesMost of the Norwegian fish harvest is taken in the Norwegian EEZ. Altogether, the waters underNorwegian jurisdiction cover some 2 million square kilometers – more than six times the area ofmainland Norway. The Arctic fisheries occur in three main areas: The Barents Sea/Svalbard, theNorth Norwegian coast, and by Jan Mayen.

In the Norwegian fisheries, Arctic cod is by far the most important in economic terms. The landedvalue was approximately 350 million $US in 2000 (Fig. 12.2.3). The landed value of herring hasalso increased considerably during the last decade, to about 160 million $US in 2000. The thirdmost valuable species is shrimp, of which the landed value was approximately 100 million $US in2000. Other important fisheries in the north include those for capelin, Greenland halibut, king crab(Paralipodes camchaticus), haddock and saithe. These fisheries are important to the activity in theprocessing plants along the coast, and hence to the viability of coastal communities.

Insert Figure 12.2.3 here. Nominal values of landings of Arctic cod, Norwegian spring spawningherring, capelin, Greenland halibut and northern shrimp 1991-2000

For the Northwest Russian fishing fleet, Arctic cod is the most important fish stock. Catches aretaken in Russian as well as Norwegian waters. Since the early 1990s, most of the cod caught byRussian fishermen in the Barents Sea has been landed abroad, primarily in Norway. Only smallquantities of mainly pelagic fish have been landed in Russia from the Barents Sea in recent years.The share of the total catch from the northeast Atlantic has however increased. The NorthwestRussian fishing fleet, previously engaged mainly in distant water fishing, now works in theimmediate northern vicinity. While only 234 000 t were taken in the Northeast Atlantic in 1990,catches from the area have been above 500 000 t in all years since then.

The economic value of the commercial exploitation of marine mammals in Norway and Russia isof minor direct significance nationally and regionally. But since these mammals are majorconsumers of commercial fish species, the harvest of marine mammals is seen as an importantcontribution to maintain a balance in the ecosystem. The fishery of mammals has also longtraditions. Archeological excavations and early historical records clearly show that whaling hasbeen conducted since ancient times (Kalland 1995), and whales were exploited before year 1000(Haug et al., 2000). In the 17th century English and Dutch whalers killed 250 Greenland rightwhales (Bowhead; Balaena mysitcetus) as an annual average in the North. These whales wereprocessed at shore stations established along the west-coast of Spitsbergen (Arlov 1996,Hacquebord 2001).

The fishing fleet and fishersThe fishing fleet in northern Norway consists of approximately 1250 vessels operating on a year-round basis (Fiskeridirektoratet 2002b). More than half of these are small vessels of 13 meters orless. The fleet has been considerably reduced over the past thirty years. Small vessels fishing withconventional gear like nets, lines and jigs dominate. A large share of the fishery therefore takesplace close to the shore and in the fjords. Larger coastal vessels are ocean going. Trawlers and

Page 15: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–15 DO NOT CIRCULATE OR CITE

purse-seiners dominate the offshore fisheries. The vessels are required to carry a license to fish,and need a fish quota to be admitted to a particular fishery. There are virtually no open accessfisheries in Norwegian waters. Most coastal communities have a number of vessels attached tothem. The number of people participating in the Norwegian fishery is shown by areas in Figure12.2.4.

Insert Figure 12.2.4 here

Northwest Russian fisheries include a variety of fishery-related activities and participants. Theyare based in Murmansk and Arkhangelsk Oblasts, and the Republic of Karelia (Hønneland &Nilssen 2000, Nilssen & Hønneland 2001). Most of the activity is located in the city of Murmansk,where most vessel owners, fish processing plants and management authorities in the region havetheir premises. The association of fishing companies in ”the northern basin” of the Soviet Union,“Sevryba” (North Fish), was founded in 1965 and given the status of General Directorate of theSoviet Ministry of Fisheries in Northwestern Russia. “Sevryba” was made into a private joint-stock company in 1992. The majority of the approximately 450 fishing vessels located innorthwestern Russia are controlled by a handful of fishing companies (in the following referred toas the “traditional” companies). The rest are distributed between kolkhozy (fishing collectives) andprivate fishing companies (in the following referred to as the “new” companies). The total numberof vessels has been stable over the last decade. Old vessels have only to a limited extent beentaken out of service, and relatively few new vessels have been purchased (Hønneland 2004).

The “traditional” fishing companies are a legacy from the Soviet period. This fleet consists mainlyof medium-size (50-70 meters) and large (over 70 meters) vessels, and counts 250-300 ships.Before the dissolution of the Soviet Union, their main activity was the exploitation of pelagicspecies in distant waters and fisheries in the Northern Atlantic Ocean. These companies nowdeploy a fleet of mid-sized factory trawlers for fishing and processing of codfish. The collectivefleet is significantly smaller in number, totaling some 80-100 vessels. Nearly all are of medium-size (50-70 meters). The fishing collectives are more diversified than other companies. Like thetraditional companies, the collectives also aim at upgrading their fleet. The “new” companies(including the so-called coastal fishing fleet) have the smallest fleet, both in number and vesselsize, limiting the range of the vessels and therefore the markets for sale of the fish. The fleetconsists of some 100 vessels, including approximately 30 coastal fishing vessels of less than 50meters.

The Russian perception of “coastal fishing” differs from that in the neighboring countries. While aNorwegian “coastal” fishing vessel normally has a small crew and goes to port for delivery ofcatches each day, a northwest Russian “coastal” fishing vessel has a crew of more than a dozenand stays at sea for weeks before landing the catch. The reasons for this are twofold. The fishingindustry that was developed during the Soviet period was based on large-scale fishing andprocessing. Traditions, skills, and infrastructure for small-scale coastal fisheries are therefore non-existent in the main fishing regions of the Russian Federation. In addition, fish stocks fordeveloping a viable coastal fishery are not available. Furthermore, the financial situation of thefishing companies is an obstacle to the development of coastal fisheries (Hønneland 2004).

The land side of the fishing industry

Page 16: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–16 DO NOT CIRCULATE OR CITE

More than 90% of the fish landed in Norway – by Norwegian, Russian and other countries’vessels - is exported. Changes in the international market for fish and fish products may thereforehave substantial effects on the processing plants as well as the rest of the industry. In 2001 about70% of the Russian cod quota was landed in Norway. This percentage has since decreased, withthe increase of landings in other countries and transshipments in the open ocean. The fishingindustry, especially the fillet-producing plants, has experienced low profitability and an increasingnumber of bankruptcies in later years (Bendiksen & Isaksen 2000). Increased competition for rawmaterial and the high costs of production in Norway go a long way to explain the problems. Inaddition, the advantage of the Norwegian industry has been its location near the resources. Newfreezing- and defrosting- technologies, and infrastructural developments that make frozen productsmore valuable (Dreyer, 2000) reduce the advantage of being close to the resource.

There are approximately 170 fish processingplants in North Norway.7 The size of the plants variessubstantially. Most of them are engaged in producing traditional products of whitefish, forexample dried cod, salted fish and stockfish. In Finmark, a relatively large share of the plantsconcentrates on filet-production, while the shrimp industry is more important in the Troms county(NORUT 2002). In Nordland county, both fillet and traditional production is important.

Many fish processing plants are heavily dependent on landings by Russian vessels. In 2001 about120 000 were landed by Russian vessels in Norway. The fishing industry, especially the fillet-producing plants, has experienced low profitability and an increasing number of bankruptcies inlater years (Bendiksen & Isaksen 2000). Increased competition for raw material and the costs ofproduction in Norway explain these problems. The advantage of the Norwegian industry has beenits location near the resources. New freezing- and defrosting- technologies, and infrastructuraldevelopments that make frozen products more valuable (Dreyer, 2000) reduce the advantage ofbeing close to the resource.

Before the dissolution of the Soviet Union, Murmansk had the largest fish-processing plant of theentire Union. Since fishing in distant waters has been reduced and catches from northern waterslanded abroad, activities at the fish-processing plants in Murmansk have been drastically reduced.The production of consumer products fell from 83 300 t in 1990, to 10 100 t in 1998 (Nilssen &Hønneland 2001). Processing of fish outside the city of Murmansk is insignificant

AquacultureOver the past two decades Atlantic salmon and trout-based aquaculture has developed in Norway,making the country the world’s biggest farmed salmon producer. The total production in 2000 was485 000 t worth 1.6 billion $US. Of this some 145 000 t of salmon and trout were produced innorthern Norway in 2000, at a first hand value of approximately 470 million $US. This makessalmon the single most important species in terms of economic value, both in the North Norway aswell as the Norwegian fishing industry as a whole.

In 2000 there were 854 licenses for salmon and trout in Norway, of which some 30% is located inthe three northern counties (Fiskeridirektoratet 2001). The number of plants and sites in the northare expected to grow considerably in the years to come (Stokka, 2000). This development will

7 2002 figures, pers. Comm.. Roger Richardsen, Fiskeriforskning.

Page 17: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–17 DO NOT CIRCULATE OR CITE

involve not only salmon, but also other fish species such as Atlantic halibut (Hippoglossushippoglossus) and cod. Over time, aquaculture is expected to become more important to the NorthNorwegian economy than the combined marine fisheries.

An important aspect of the aquaculture industry is that it is dependent on a huge supply of pelagicfish species. Fishmeal and oils are important components of the diet of many species of farmedfish, including salmon and trout. The quantity of supplies needed is so high that the industry at aglobal level is sensitive to rapid fluctuations in important pelagic stocks. ENSO events in thePacific have already affected the industry through impacts on anchovy (Engraulis spp.) stocks.From 1997 to 1998 the global marine fishery was reduced by nearly 8 million t, mainly caused byEl Nino-events (FAO 2000). Reduced supply on the international market led to increased prices offish meal in this period. The IPCC states in its 2001 report that unless alternative sources ofprotein are found, aquaculture could in the future be limited by the supply of fishmeal and oils.

Aquaculture is in its infancy in Nortwest Russia, and the total production is negligible. It ishowever likely to increase in the future.

Employment in the fisheries sector and the fisheries communitiesDemographic pressure towards urbanization, which is expected to continue (IPCC 2001:38), maybe said to be one of the major driving forces behind this development. Other factors, such as lackof employment opportunities and inferior public services, may be seen both as a cause of theproblem as well as a consequence. There is also a trend of fishing boats being sold out of thecommunities. This situation indicates that the small fishery dependent societies are undercontinuous pressure. These societies are subject to a "double exposure" (O’Brien and Leichenko,2000), where climate change occurs simultaneously with economic marginalization. TheNorwegian government has for a long period run programs aimed at strengthening the viability offishery-dependent societies in the north. In later years these efforts have been directed towardsmarket orientation, flexibility and a more robust industrial structure (St. melding 51 1997-1998),rather than subsidies to the industry. Some regional development programs are aimed atdiversification of the economic activity in remote areas by supporting, among other things,female-run enterprises (Lotherington & Ellingsen 2002)

Among the Russian federation subjects in the northwest, the Murmansk Oblast is most importantfrom the point of view of fisheries. This region is one of the most urbanized in Russia, with some92 per cent of the population living in cities and towns. Most of the Northwest Russian fishingfleet is concentrated in the city of Murmansk. Some companies are located in the three otherRussian federation subjects of the area: Arkhangelsk (Arkhangelsk Oblast), Petrozavodsk (theRepublic of Karelia) and Narjan-Mar (Nenets Autonomous Okrug).

The fishing industry is important for several major cities in Northwestern Russia, but these cannotbe characterized as ‘fishing communities’ in the sense this concept is understood in the West.Their viability is not dependent on fisheries. Also, the significance of the fishing industry has beenseverely reduced in the post-Soviet period as the catches of Russian vessels are mainly deliveredto the West. The redirection of landings to the home market has been one of the main ambitions ofRussian fishery authorities at both the federal and regional levels since the early 1990s. The factthat this has not been achieved points to the relative impotency of these bodies. At the federal

Page 18: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–18 DO NOT CIRCULATE OR CITE

level, the State Committee for Fisheries has twice lost its status as an independent body ofgovernance (subsumed into the Ministry of Agriculture in 1992-93 and 1997-98) and seen itstraditional all-embracing influence over fisheries management significantly reduced. Notably, theMinistry of Trade and Economic Development in 2000 succeeded in introducing a system forquota auctions, against the will of the State Committee for Fisheries. Regional authoritiesincreased their influence during the 1990s. This development is now reversed as a result of theprocess of re-centralization that set in around 2000, commensurate with wider developments inRussia since President Putin came to power. Hence, while regional authorities in northwesternRussia have a declared aim of developing coastal fisheries, actual development in this sphere canonly be characterized as miniscule.

In 2000 there were 854 licenses for the culture of salmon and trout in Norway, of which some30% are located in the three northern counties (Fiskeridirektoratet 2001). The number of plantsand sites in the north are expected to grow considerably in the years to come (Stokka, 2000). Thisdevelopment will involve not only salmon, but also other fish species such as halibut and cod.Over time, aquaculture is expected to become more important to the North Norwegian economythan the combined marine fisheries.

Markets8

Norway is one of the worlds biggest fish exporters - more than 90% of the landings are exported.9There are two aspects to this: The income generated by fish exports is substantial – around 4billion $US in 2001. As the production in aquaculture will increase, and the production ofpetroleum decrease, exports of fish products can be expected to become more important in thefuture. The Ministry of Fisheries envisions that aquaculture will be a mainstay of the Norwegianeconomy in the years to come, and that the sales value in northern Norway will be nearly fivetimes higher in 2020 than today. The second aspect is that Norway is a major supplier to manymarkets. The Norwegian imports are important to for example the EU market for seafood, whichis therefore to some extent vulnerable to fluctuations in Norwegian fisheries.

The single most important species in terms of export value is salmon, which fetched 1.8 billion$US in 2000. The second most important category are whitefish, the exports of which (consistingof mainly cod, haddock and saithe) is in the range of 1.2 billion $US annually. Pelagic species,where herring is the most important one, represent an export value of 920 million $US in 2001.The number four species in terms of export value is northern shrimp.

As to Russian exports of fish products from the North, landings of Russian-caught cod to Norwayas mentioned have increased since 1990. During the years 1995-97, landings were about 250 000-300 000 t per year. Since then, there has been a reduction in Russian landings of cod as well asother fish in Norway. Transshipments of fish at sea and landings in other countries are increasingwhile landings in Norway are decreasing. The catches landed in Russia mostly go to the Russianconsumer market. Imports of fish to Russia from Norway is rapidly increasing.

The management regime

8 All figures from the Norwegian Seafood Export Council (http://www.seafood.no/).

9 In 2001 Norway was the world’s number two fish-exporter, after Thailand.

Page 19: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–19 DO NOT CIRCULATE OR CITE

In addition to her EEZ, Norway also manages the resources in the fishery zone around Jan Mayenand in the Fishery Conservation zone around Svalbard. The Norwegian EEZ borders the EU-zonein the south, the Faroe Islands in the southwest and Russia to the east. A large area beyond theEEZ boundary in the Norwegian Sea and a smaller area in the Barents Sea are internationalwaters. Most of the economically important stocks move between the zones of two or more states.

Cooperation between the owner countries in the management of these stocks is essential to ensuretheir sustainable use. A series of agreements has been negotiated among the countries in theNortheast Atlantic that establish bilaterial and multilateral arrangements for cooperation onfisheries management. The most extensive management regime on Arctic stocks in the NortheastAtlantic is the one between Norway and Russia. A joint fisheries commission meets annually toagree on total allowable catches and allocation for the major fisheries in the Barents Sea: cod,haddock and capelin.10 The total quotas set are shared between the two countries - the allocationkey is 50-50 for cod and haddock, and 60-40 for capelin. A fixed additional quantity is traded tothird countries. There are also agreements on mutual access to each other's EEZs and exchange ofquotas through this arrangement (Hoel 1994). An important aspect of the cooperation with Russiais that a substantial part of the Russian harvest in the Barents Sea is taken in the Norwegian zoneand, as mentioned earlier, and landed in Norway. The cooperation entails joint efforts also infisheries research and in enforcement of fisheries regulations.

Despite disagreement between Norway and Russia on the delimitation of the boundary betweentheir EEZ and the shelf in the Barents Sea, the cooperation on resource management between thetwo countries may generally be characterized as well functioning (Hønneland 1993). However,agreed TACs by Norway and Russia have, in some years, exceeded the ones recommended byfisheries scientists. In addition, the actual catches have sometimes been larger than the agreedones. Since the late 1990’s, a precautionary approach has been gradually implemented in themanagement of the most important fisheries. However, retrospective analyses have shown thatICES estimates of stock sizes have often been too high, thereby incorrectly estimating the effect ofa proposed regulatory measure on the stock. This has had the unfortunate effect that stock sizesfor a given year are adjusted downward in subsequent assessments, rendering adoptedmanagement strategies ineffective (Nakken 1998, Korsbrekke et.al. 2001). However, the JointCommission has decided that from 2004 onwards multi-annual quotas based on a precautionaryapproach will be applied. A new management strategy adopted in 2003 shall ensure that TAC(total allowable catch) levels for any 3-year period shall be in line with the precautionary referencevalues provided by the ICES.

A number of other agreements are in effect in the area, notably a 5-party agreement among thecoastal states in the Northeast Atlantic to manage Atlanto-Scandian herring (Ramstad 2001). Heretotal quotas for the following year’s herring fishery are set, and divided among the parties. Aseparate quota is set for the area on the high seas in the Norwegian Sea. The high seas quota, mostof which is given to the same coastal states, is formally managed by the Northeast AtlanticFisheries Commission (NEAFC), who is mandated to manage the fishing on the high seas in theNortheast Atlantic. Norway has also an extensive cooperation with the EU on the management onshared stocks in the North Sea, as well as exchange of fish quotas which entails access for EU

10 Since 2001 a total quota has also been set for the King crab fishery.

Page 20: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–20 DO NOT CIRCULATE OR CITE

vessels to North Norwegian waters. The EU is given a major share of the third country quota ofcod in the North.

Management measures for marine mammals, harvested in the area, are decided by the InternationalWhaling Commission (IWC), the North Atlantic Marine Mammals Commission (NAMMCO), andthe Joint Norwegian-Russian Fisheries Commission. The IWC has not been able to adopt a RevisedManagement Scheme and does therefore not set quotas. Since 1993 Norway has set unilateralquotas for the take of minke whales, on the basis of the work of the IWC Scientific Committee(Hoel 1998). NAMMCO adopts management measures for cetaceans and seals in the NorthernNortheast Atlantic area (Hoel 1993).

A precondition for sound management of living marine resources is that sufficient knowledgeabout the resources is available. In Norway, the Institute of Marine Research (IMR) is the maingovernmental research institution, while the Northern Institute of Marine Research (PINRO) hasthe same role on the Russian side. The International Council for the Exploration of the Sea (ICES)is the international institution for formulation of scientific advice to the fisheries authorities in theNorth Atlantic countries. Its work is generally based on inputs from the research institutions in themember countries. The ICES advice is now based on a precautionary approach, which seeks tointroduce a greater sensitivity to risk and uncertainty into management. One of the challenges forfisheries-management in the future is a better understanding of species interactions (multi-speciesmanagement), more reliable data from scientific surveys, as well as a better understanding of theimpact of physical factors - like changing climatic conditions - on stocks. A major challenge is thedevelopment and implementation of an ecosystem-based approach to the management of livingmarine resources, where also the effects of climate change is taken into consideration whenestablishing management measures.

The management measures essentially fall in three categories:• Input regulations in the form of licensing schemes restricting access to a fishery.• Output regulations, consisting of the fish quotas given to various groups of fishermen

which limit the amount of fish they are entitled to in any given season.• Technical measures specifying for example the type of fishing gear that must be used in a

particular fishery.

The objectives of fisheries management in Norway are related to conservation, efficiency andregional considerations (Report to Parliament 1998). Conservation of resources is seen as aprecondition for the development of an efficient industry and maintenance of viable fishingcommunities. An important objective of the fisheries policy is to improve the economic efficiencyof the industry. An important issue is therefore to reduce the capacity of the fishing fleet, which ismuch larger than needed to take the quotas available and therefore makes the costs of fishing toohigh. Attempts to remove excess capacity include subsidies for scrapping vessels, regulatorymechanisms, and vessel construction regulations. A quota arrangement allowing for merging twovessels’ quotas while removing one of the vessels from the fishery gives vessel owners an incentiveto remove excess fishing capacity, and can contribute to a more efficient fleet. Another aspect ofthis is, however, that coastal communities can see their local fleet reduced or even disappearing,threatening the viability of that community.

Page 21: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–21 DO NOT CIRCULATE OR CITE

The enforcement of the fisheries regulations in Norway is carried out both at sea and when the fishis landed. At sea, the Coast Guard is responsible for inspecting fishing vessels and checking theircatch against vessel log-books. Foreign vessels fishing in Norwegian waters are also inspected.The activity of the Coast Guard is vital for the functioning of the management regime as a whole.A recent development is that ocean going vessels can be required to install and use a satellitebased vessel-monitoring system enabling the authorities to continually monitor their activities.The Directorate of Fisheries also inspects the activities on the fishing grounds, as well as on thelanding sites. When fish is landed, the sales organization buying the fish reports the landedquantity to the Fisheries Directorate, which is responsible for maintaining the fisheries statistics.

Fisheries regulations were in the Soviet northern fishery basis issued by the Sevryba association.As this organization lost its status in fisheries regulation in the mid-1990s, the regulatory taskswere partly taken over by the enforcement body Murmanrybvod, partly by the fisheriesdepartments of regional authorities in each federal subject in the area, and since 2000 to anincreasing extent been the remit of federal authorities. During the 1990s, the Russian share of theBarents Sea quotas were first divided among the four federal subjects of the region by the so-called Scientific Catch Council (formerly headed by Sevryba, since 2001 by the federal StateCommittee for Fisheries). Within each federal subject, a Fisheries Council (led by regionalauthorities) distributed quota shares among individual shipowners. The influence of both theScientific Catch Council and the regional Fisheries Councils was reduced after the introduction ofquota auctions in 2000/2001. Since then, an increasing share of the quotas has been sold onauctions, administered by the federal Ministry of Trade and Economic Development. In November2003, the Russian Government decided to abolish the auctions and instead introduce a resourcerent (a fee on quota shares). The quotas will from 2004 be distributed by an inter-ministerialcommission at the federal level, so the regional authorities will further lose the influence of inter-regional quota allocation (Hønneland 2004).

Apart from quotas, the Russians have fishery regulations quite similar to those found in theNorwegian system: regulations pertaining to fishing gear, size of the fish and composition ofindividual catches. In addition, the Russians have a more fine-meshed system than the Norwegiansfor closing and opening of fishing grounds. Individual inspectors from the enforcement bodyMurmanrybvod or researchers from the scientific institute PINRO can close a ‘rectangle’ (a squarenautical mile) on site for a period of three days. After three days, the ‘rectangle’ is re-opened ifscientists make no objections, i.e. if the proportion of undersized fish in catches does not continueto exceed legal limits.

Traditionally, the civilian fishery inspection service Murmanrybvod, subordinate to the RussianState Committee for Fisheries, has been in charge of the enforcement of Russian fisheryregulations in the Barents Sea. In 1998, responsibility for fisheries enforcement at sea in theRussian Federation was transferred to the Federal Border Service. In the northern fishery basin,the Murmansk State Inspection of the Arctic Regional Command of the Federal Border Servicewas established to take care of fisheries enforcement. It is only responsible for physicalinspections at sea, which has been transferred to the Border Guard. Murmanrybvod is still incharge of keeping track of how much of the quotas has been caught by individual ship owners atany one time. It has also retained its responsibility for the closing of fishing grounds in areas withexcessive intermingling of under-sized fish, a very important regulatory measure in both the

Page 22: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–22 DO NOT CIRCULATE OR CITE

Russian and Norwegian part of the Barents Sea. Finally, Murmanrybvod is still responsible forenforcement in international convention areas. In practice, Murmanrybvod places its inspectors onboard Northwest Russian fishing vessels that fish in the NEAFC or NAFO areas.

The reorganization of the Russian enforcement system is generally believed to have led to areduction in the system’s effectiveness, at least in a short-time perspective. For one thing, officersin the Murmansk State Inspection of the Federal Border Service generally lack experience infisheries management and enforcement. This has partly been compensated for by transfer of someof Murmanrybvod’s inspectors. More apparent is the lack of material resources to maintainpresence at sea. Quite contrary to the intentions of the reorganization of the enforcement system,the presence at sea by monitoring vessels declined since the Border Guard took over this duty in1998. Precise data for presence at sea and inspection frequency are not available, but Jørgensen(1999, pp. 88-90) estimates that the Border Guard performed around 160 inspections at sea in1998, which represents a significant reduction compared to an estimated 700-1 000 annualinspections at sea by Murmanrybvod prior to the reorganization. For several months on end during1998, not a single enforcement vessel was present on the fishing grounds in the Russian part of theBarents Sea. Officials of the Border Service explain this with lack of funds to purchase fuel.Critics question the genuineness of the Border Service’s will to play a role in fisheriesmanagement. The result of the reorganization has, in any event, so far led to a tangible reductionin the effectiveness of Russian enforcement in the Barents Sea.

12.2.7 Economic and social impacts of climate change on fisheries in the Northeast AtlanticThe economic importance of fisheries to Northern Norway is substantial, cod being the mostsignificant species. Problems related to profitability in the fishing-industry have been evident for along time, and have contributed to depopulation-problems in remote, fishery-dependent areas.Aquaculture is, however, a growing industry and is expected to be important to the future viabilityof local communities in Northern Norway. In northwest Russia, the fishing industry is based inbig cities, Murmansk in particular, and is therefore not as significant to local communities as thecase is in Norway.

A study by Furevik et al. (2002) developing regional ocean surface temperature scenarios for thenortheast Atlantic, concludes that for the 2020 scenario, no substantial change is likely in thephysical parameters. The authors conclude that as far as the ocean surface temperature goes, aslight cooling is likely. As to the longer-term scenarios, warming is indicated. For the near-termfuture, climate change is therefore not likely to have major impact on the fisheries in the region.The uncertainties surrounding these scenarios are however considerable. These uncertainties areamplified when the physical effects on biota are addressed, and amplified again when effects ofclimate change on society are to be studied. In addition, social change is driven by a vast numberof factors, climate change being only one of these. What follows is therefore tentative and shouldbe read more as discussions of likely patterns of change, not as predictions of futuredevelopments.

The issue of effects of climate change is closely related to the vulnerability of industries andcommunities, and their capability to adapt to change and mitigate the effects of change (see alsochapter 16). Vulnerability "... is defined as the extent to which a natural or social system issusceptible to sustaining damage from climate change" (Mc Carthy et al., IPCC 89:2001). It

Page 23: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–23 DO NOT CIRCULATE OR CITE

depends on the ability and capacity of society at the international, national and regional level tocope with changes and remedy its negative effects. Also, climate change may produce positivechanges.

The fisheries sector is one in which the industry has always had to adapt to and cope withenvironmental change. As demonstrated above, the abundance of various species of fish andmarine mammals has varied throughout history, often dramatically and also within short timespans. Adapting to changing circumstances is therefore second nature to the fishing industry aswell as to the communities that depend upon it. An important issue in this context is thus whetherclimate change brings about changes at scales and rates that are unknown, and whether adaptationcan be achieved within the existing institutional structures.

Resource managementResource management is the key factor in deciding the biological and economical sustainability ofthe fisheries. The fishing opportunities are decided by the management regime. There are virtuallyno remaining fisheries where the economic result is left to the decision by industry itself. Thedesign and operation of both the domestic and international management regimes are thereforecrucial to the sustainability and economic efficiency of the fisheries, and hence the economicviability of the communities that depend upon them. The development and implementation of aprecautionary approach, as well as the emergence of an ecosystem-based management, mayenhance the resilience of the stocks and therefore make the industry and communities more robustto future external shocks. As noted above, the main arrangements for managing living marineresources in the Northeast Atlantic are being modified in this direction, with the implementation ofa precautionary approach and the development of an ecosystem-based approach to management.

A major challenge for the management regime is that of adjusting to the possible changes inmigration patterns of stocks, resulting from climate change. This finding is in conformity with thatof the IPCC (2001) and Everett et al. 1996. Changes in migration patterns of fish stocks havepreviously upset established arrangements for resource management, and can trigger conflictsbetween countries. One example is that of Arctic cod: in the early 1990’s, the stock extended itsrange northwards in the Barents Sea, into the high seas in the area (the so-called “loophole”).Vessels from a number of countries without fishing rights in the cod fishery took the opportunityto initiate an unregulated fishery in the area, thereby undermining the Norway-Russia managementregime. This triggered a conflict between Norway and Russia on the one hand, and Iceland on theother. The conflict was later resolved through a trilateral agreement (Stokke 2001). Anotherexample is that of the Norwegian Spring Spawning Herring (cf. textbox 12.2.1): Following morethan two decades of effort at rebuilding the stock on the part of Norwegian authorities, it started tomigrate from the Norwegian EEZ and into international waters for parts of the year in the mid-1990s. Thereby the stock became accessible to vessels from other countries, and, in the absence ofan effective management regime for the stock at the high seas, the efforts at rebuilding the stockcould prove futile. As mentioned above, a regime securing a management scheme for the stockeventually came into place, but this took several years to negotiate. The upshot of this is thatchanges in migration patterns, which are likely to be triggered by changes in water temperatures,tend to give rise to unregulated fishing and conflicts among countries. The outcome of suchconflicts may be conflicting management strategies, new distribution formulas, or even newmanagement regimes.

Page 24: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–24 DO NOT CIRCULATE OR CITE

Another factor to consider is that negative events tend to be a liability to the management regime.The so-called cod-crisis in the late 1980’s, for example, led to several modifications of the existingregime. The management regime is likely to be held responsible for social and economicalconsequences of climate change. This may in turn affect the legitimacy and authority of theregime, and its effectiveness in regulating the industry. An important aspect in that regard is theway decisions about resource management and allocation of resources are made. A regime thatinvolves those interests that are affected by decisions in the decision-making processes, tends toproduce regulations that are considered as more legitimate than do regimes that do not involvestakeholders (Mikalsen and Jentoft 2003).

Fisheries management models of today are mainly based on general assumptions of constantenvironmental factors. When the environment changes, the current methods applied in fisheriesmanagement are not adapted for that. A study by Eide and Heen (2002) investigates the economicoutput from the fisheries under different environmental scenarios and under different managementsituations for the cod and capelin fisheries in the Barents Sea. Using the ECONMULT fleet model(Eide and Flaaten, 1998) and a regional impact model for the North-Norwegian economy (Heenand Aanesen, 1993), they conclude that even a narrow range of management regimes has a varietyof possible economic outcomes. Even though global warming may have significant potential effecton catches, profitability, employment, and income, changes in the management regimes seem tohave an even larger impact. This conclusion sets the discussion of effects of global climate changein perspective. It implies that a large number of factors influence the economic activities and theiroutput and, furthermore, that the operation of the management regime seem to be the mostsignificant of these.

The crucial factor for resource management in the face of climate change is therefore thedevelopment of robust and precautionary approaches and institutions for managing the resources.The decisive factor for the health of fish stocks, and therefore the fate of the fishing industry andits dependent communities, appear to be the resource management regime.

The fishing fleetThe ability to adapt to changes in migration patterns or stock size of commercially exploitedspecies will vary between different vessel groups in the fishing fleet. The ocean-going fleet iscapable of adjusting to changes in migration patterns, as it has a wide operating range. Smallcoastal vessels are more limited in that regard. Thus, northern communities with a strongdependency on small coastal vessels are likely to become more affected if migration patterns andavailability of important fish stocks change significantly. Obviously, if fish stocks move closer tothe coast it is an advantage to the coastal fleet, while it is a disadvantage for this fleet if the stocksmove more seawards. Such a development may be confounded by changing weather patterns withsevere weather events becoming more prevalent. All vessel groups will be affected if changesimply stocks crossing jurisdictional borders. That may imply a change in distribution of resourcesamong countries.

Increased production and larger stocks of cod and herring are possible outcomes of climate changein Northeast Atlantic. A question arises of what fleet groups are most capable of making the bestof such positive changes of the resource. Such changes may result in different availability to the

Page 25: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–25 DO NOT CIRCULATE OR CITE

resources between groups of fisheries (e.g. coastal versus ocean-going vessels), affecting thedomestic allocation of resources. It may also lead to a greater political pressure to change theallocation of resources between the main groups of resource users.

Changes in stock abundance and migration patterns are not new to the industry. The availability offish stocks and their accessibility to the coastal fleet has changed throughout recorded history, andthe industry as well as the management regime is used to adapt to changing circumstances. Thekey question is whether climate change would amplify such variations and aggravate their effectsbeyond the scale, which the industry and the regulating authorities are familiar with.

Changes in oceanic conditions may also affect the migrating ranges of marine mammals, andhence marine mammal – fisheries interactions. Such interactions could be mammals predating onfish, thus increasing competition with fishermen, or marine mammals interacting directly with thefishery, e.g. by interference with fishing gear. Marine mammals are also vectors of parasites thatmay affect fish and fisheries.

AquacultureHigher temperature has generally positive effects on aquaculture in terms of growth of fish: IPCC2001 notes that warming and consequent lengthening of the growing season “…could havebeneficial effects with respect to growth rates and feed conversion efficiency..” (pt 6.3.5). Warmerwaters may also have negative effects sincethe presence of lice and diseases may be related towater temperature. In recent years high water temperatures in late summer have caused highmortality in farms rearing halibut and cod, the production of which is still at a pre-commericalstage. Salmon is also affected by high temperatures and farms may expect higher mortalities dueto this. A rise in sea temperatures may therefore favor a northward movement of production, tosites where the waters are unlikely to reach peak temperatures above levels affecting fishnegatively.

Increase of severe weather-events can be a cause of escapes from fish pens and consequent loss ofproduction. Escapes are also a potential problem in terms of spreading of disease. However,technological developments may compensate for this.

As pointed out above, the aquaculture industry is dependent on capture fish for salmon feed.Climate change may cause lack of and/or variability in the market for such products, but this isalso an area where research may lead to the development of other feed sources.

The processing industry, communities, and marketsThe fish processing industry in the north faces challenges in the structural changes both in thefirst-hand market (from fisher to buyer) and in the export-market. Increased internationalcompetition for scarce resources has left the processing side of the industry increasinglyvulnerable to globalization pressures. At the same time many of the communities, depending onfisheries for their existence, experience economic marginalization and depopulation relatedproblems. The vulnerability of the fishing industry and fishing communities can therefore beconsidered as relatively high at the outset, rendering them particularly susceptible to any negativeinfluences stemming from climate change. Such impacts may however be minor compared to thatof other drivers of change. Furthermore, the fish processing industry is very varied. The size of

Page 26: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–26 DO NOT CIRCULATE OR CITE

fish processing plants is one aspect of this, their versatility and ability to vary production andadapt to changing circumstances is another. The ability of the particular type of industry to adaptto various earlier “crises”, be it in terms of demand or supply failures, could be an indicator oftheir future "coping-capacity" towards effects stemming from climate change. Another issue isthat climate-induced changes elsewhere in the world may affect the situation for the NorthNorwegian fishing industry and fishing communities. Experiences from e.g. the fisheries crisis inCanada in the 1990’s indicate that such situations tend to intensify competition for fish to go intothe production. To the industry in Norway, with high labor costs, such a scenario is negative.

12.2.8. Ability to cope with changeMany factors contribute to a community's "coping-capacity" in relation to depopulation and thestructural changes in the fisheries sector described above (Baerenholdt & Aarsaether 2001). Thefuture of these settlements may depend on their ability to adapt to increased competition,efficiency, deregulation and liberalization of the markets as much as the accessibility of fishingresources for their local production systems (Lindkvist, 2000).

While the management regime can be seen as an instrument to ease negative effects of climatechange, it is however also important to consider public measures beyond the fisheries managementregime that affect the conditions of the fishing industry more broadly, as for example regionalpolicies and development of alternative means of employment. Measures for buildinginfrastructure as roads or develop harbor facilities is one example of this. Government support forfisheries in the form of direct subsidies is now in practice all but prohibited by internationalagreements. But in Norway in particular there is a strong tradition for supporting regionaldevelopment in a more broad sense, and programs to this end may enhance the resilience ofnorthern communities.

In addition to adaptation to possible changes in the resource stemming from climate change, thefishing communities will also need to adapt to possible other climate-related changes in theirliving-area (e.g. weather-events) and their effects on terrestrial biota and infrastructure. These mayhave indirect influence on the fishery-sector, related economic activities, or on other aspects oflife, valued by the people in the respective communities.

Page 27: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–27 DO NOT CIRCULATE OR CITE

12.2.9. References

Arlov, T.B. 1996. Svalbards historie. Aschehoug, OsloBerg, R. 1995. Norge på egen hånd. Norsk utenrikspolitisk historie bd 2. Universitetsforlaget,

Oslo.Bergstad, O. A., Jørgensen, T. And Dragesund, O., 1987. Life history of the gadoid resources of

the Barents Sea. Fisheries Research., 5: 119-161.Bendiksen, B.I & Isaksen J.R., 2000: Fiskerinæringen i Finnmark - Analyse av verdiskapning og

råstoffomsetning. Rapport nr. 12/2000, Tromsø: Fiskeriforskning.Bogstad, B., Hiis Hauge, K. and Ulltang, Ø., 1997. MULTSPEC-A multispecies model for fish

and marine mammals in the Barents Sea. J.Northwest.Atl.Fish.Sci. 22: 317-341.Bogstad, B., 2002. Økosystemet I Barentshavet. Fisken og Havet, særnummer 1 – 2002. ISSN

08020620.Bogstad, B., Haug, T. And Mehl, S., 2000. Who eats whom in the Barents Sea? NAMMCO

Sci.Publ. 2: 98-119.Baerenholdt, J.O & Aarsaether, N., 2001. Transforming the local. Coping Strategies and Regional

Policies. Nord 2001: 25Dickson, R., 1992. Hydrobiological Variability in the ICES area, 1980-1990. Introduction. ICES

mar.Sci.Symp., 195: 1-10, 1992.Dreyer, B., 2000: Globalisering av råvaremarkedet - strategiske utfordringer for lokal

fiskeindustri. Økonomisk Fiskeriforskning Rapport nr. 10:2, Fiskeriforskning, 2000.Eide, A. og O. Flaaten, 1998. Bioeconomic Multispecies Models of the Barents Sea Fisheries. In

Rødseth, Tor (ed.): Models for Multispecies Management, Physica-Verlag, Heidelberg,New York, ISBN 3-7908-1001-0, p. 141-172.

Eide, A. og K. Heen, 2002. Economic impacts of global warming. A study of the fishing industryin North Norway. Fisheries Research 56/3, p. 261-274.

Ervik, A., Kiessling, A., Skilbrei, O. and van der Meeren, T. (eds) 2003: Havbruksrapport. Fiskenog havet, særnummer 3.

Everett, J.T., A. Kronvin, D. Luch-Belda, E. Okemwa. H.A. Regier, and J.P. Troadec 1996:Fisheries. In Climate Change 1996: Impacts, Adaptations, and Mitigation of Climate Change:Scientific-technical Analyses. Contribution of Working Group II to the Second AssessmentReport of the Intergovernmental Panel of Climate Change, pp. 511-537

FAO,2000: The s ta te of world f isher ies and Aquacul ture 2000.(http://www.fao.org/documents), Food and Agriculture Organization of the United Nations

Fiskeridirektoratet 2001: Nøkkeltall fra norsk havbruksnæring. Rapport, Fiskeridirektoratet,Bergen.

Fiskeridirketoratet 2002a: Beskrivelse av havbruksnæringen i området Lofoten til den norsk-russiske grense. Fiskeridirektoratet, Bergen.

Fiskeridirektoratet 2002b: Fiskeriaktiviteten i området Lofoten – Barentshavet. Fiskeridirektoratet,Bergen

Furevik, T., Drange, H. and Sorteberg, A. 2002: Anticipated changes in the Nordic seas marineclimate: Scenarios for 2020, 2050, and 2080). Havforskningsinstituttet, Bergen

Gjøsæter, H., 1995: Pelagic fish and the ecological impact of the modern fishing industry in theBarents Sea. Arctic, Vol.48 no.3: 267-278.

Page 28: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–28 DO NOT CIRCULATE OR CITE

Hamre, J., 2003. Capelin and herring as key species for the yield of north-east Arctic cod. Resultsfrom multispecies model runs. SCI.MAR., 67 (Suppl.1): 315-323.

Hacquebord, L. 2001. Three centuries of Whaling and Walrus Hunting in Svalbard and its Impacton the Arctic Ecosystem. Environment and History 7, 169-185.Hartvigsen, R., A. Olsen, PK Hansen, Ø Bjerk, E Bolle and S Jensen: Beskrivelse av

havbruksnæringen I området Lofoten til den norsk-russiske grense. Fiskeridirektoratet,Bergen.

Heen, K. and M. Aanesen, 1993. Kryssløpsmodell for nordnorsk økonomi 1987.Rapport NORUTsamf.forskn., Tromsoe. (In Norwegian).

Hylen, A., 2002. Variation in abundance of Arcto-Norwegian cod during the 20th century. ICESmar.Sci.Symp., 215: 000-000.

Hoel, A.H., 1993: Regionalisation of international whale management: the case of the NorteastAtlantic Marine Mammals Commission Arctic Vol XX, No. pp Summer 1993Hoel, A.H., 1994: The Barents Sea: Fisheries Resources for Europe and Russia. In Schram Stokke,

O., & Tunander, O. (ed.), 1994: The Barents Region. Cooperation in Arctic Europe.International Peace Research Institute, Oslo and the Fridtjof Nansen Institute, 1994.

Hoel, A.H., 1998: Political uncertainty in international fisheries management. Fisheries ResearchVol 37, No. 763, pp. 239-250.Hønneland, G. & Nilssen, F. (2000), ‘Comanagement in Northwest Russian fisheries’, Society and

Natural Resources 13, 635-648.Hønneland, G. 2004: Russian Fisheries Management: The Precautionary Approach in Theory and

Practice. Leiden and Boston: Martinus Nijhoff Publishers/Brill Academic PublishersIPCC 2001: Climate Change 2001 – Impacts, Adaptation, and Vulnerability. Contribution of

Working Group II to the Third Assessment Report of the IPCC. Cambridge University Press,Cambridge.

Iversen, S.A. (red), 2002. Havets ressurser 2002. Fisken og Havet, særnr.1-2002Jørgensen, A.K. (1999), Norsk og russisk fiskerikontroll i Barentshavet – en sammenligning

med hensyn til effektivitet (Norwegian and Russian Fisheries Enforcement in the Barents Sea– a Comparison with Regard to Effectiveness), MA Thesis, Department of Political Science,University of Oslo, Oslo.

Kola Science Centre (2001): Obshchiy vylov predmpriyatiyami Severnogo basseyna 1990-2000(Total catch of the enterprises of the Northern basin 1999-2000), Apatity: Kola ScienceCentre.

Korsbrekke, K., Mehl, S., Nakken, O. and Pennington, M., 2001. A survey-based assessment ofthe Northeast Arctic Cod stock. –ICES Journal of Marine Science, 58: 763-769.

KPMG and SINTEF 2003: Betydningen av fiskeri- og havbruksnæringen for Norge – enringvirkningsanalyse. Report, KPMG, Trondheim

Lotherington, Ann-Therese & Ellingsen, May Britt 2002: Små penger og Store Forventninger.Evaluering av Nettverkskreditt. NORUT, rapport nr.8.

Lindkvist, K.B., 2000: Dependent and Independent Fishing Communities in Norway. In Symes(ed.), 2000: Fisheries Dependent Regions. Fishing News books, Blackwell Science.

McCarthy J., Canziani, O., Leary, N.A., Dokken, D.J., White, K.S., Climate Change 2001:Impacts, Adaptation, and Vulnerability. Contribution of Working Group II to the ThirdAssessment Report of the Intergovernmental Panel of Climate Change. published for theIntergovernmental Panel of Climate Change. Cambridge University Press

Michalsen, K. (ed.) 2003: Fisken og havet, særnummer 1. Instiutte of Marine Research, Bergen.

Page 29: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–29 DO NOT CIRCULATE OR CITE

Mikalsen, K. and Jentoft, S. 2003: Limits to particpation? On the history, structure, and reform ofNorwegian fisheries management. Marine Policy Vol. 27, No. 5, pp. 397-407

Ministry of Fisheries 2002: Fiskeri og havbruk 20002. Statistical Report, Ministry of Fisheries,Oslo.

Nakken, O., 1998. Past, present and future exploitation and management of marine resources inthe Barents Sea and adjacent areas. Fisheries Research, 37: 23-35.

Nakken, O., 1994. Causes of trends and fluctuations in the Arcto-Norwegian cod stock. ICESmar.Sci.Symp., 198: 212-228.

Nilssen, K.T. 1995. Seasonal distribution, condition, and feeding habits of Barents Sea harp seals.In: Blix, AS, L. Walloe, and O. Ulltang (eds) 1995): Whales, Seals, Fish and Man.Proceedings of the International Symposium on the Biology of Marine Mammals in theNortheast Atlantic, Tromsø, Norway, 29 November – 1 December 1994. Developments inMarine Biology 4, Elsevier, Amsterdam

Nilssen, F. & Hønneland, G. (2001), ‘Institutional change and the problems of restructuring theRussian fishing industry’, Post-Communist Economies 13, 313-330.

Nordøy, E., PE Mortensson, AR Lager, LP Folkow and AS Blix: Foord consumption of theNorteast Atlantic stock of harp seals. In: Blix, AS, L. Walloe, and O. Ulltang (eds) 1995):Whales, Seals, Fish and Man. Proceedings of the International Symposium on the Biology ofMarine Mammals in the Northeast Atlantic, Tromsø, Norway, 29 November – 1 December1994. Developments in Marine Biology 4, Elsevier, Amsterdam

Norut Samfunnforskning, 2002: Konjunkturbarometer for Nord-Norge, NORUTO’Brien, K. and Leichenko, R. 2000. “Double Exposure: Assessing the Impacts of Climate

Change Within the Context of Globalization.” Global Environmental Change 10: 221 - 232.Ottersen, G. and Sundby, S., 1995. Effects of temperature, wind and spawning stock biomass on

recruitment of Arcto-Norwegian cod. Fish. Oceanogr. 4: 278-292.Pope, J., Large, P. and Jakobsen, T., 2001. Revisiting the influences of parent stock, temperature

and predation on the recruitment of the Northeast Arctic cod stock, 1930-1990. – ICES Journalof Marine Science, 58: 967-972.

Ramstad, S., 2001: Etableringen av et Internasjonalt Forvaltningsregime for Norsk VårgytendeSild. En kamp om ressurser mellom Norge, Russland, Island, Færøyene og EU. Hovedoppgavei Statsvitenskap, Universitetet i Tromsø.

Report to Parliament 1998: Perspectives on the Development of the Norwegian Fishing Industry.(In Norwegian). (Stm nr 51 (1997-98). Ministry of Fisheries, Oslo

Rydningen A., & Johnsen J.P., 2002: Marine samfunn - samfunnsmessige forutsetninger for øktverdiskaping i marin sektor. NORUT Samfunnsforsking AS

Sergeant, D.E., 1991: Harp seals, man and ice. Can Spec Publ Fish Aquat Sci;114:1-53Solhaug, T., 1983: De norske fiskeriers historie 1815-1880. Bergen, Universitetsforlaget.Stokka, A. Sparboe, L.O., Sætermo, I.A.F., Jonassen T., Vik L.H., Giæver, A., Dreyer, H.,

Lilletun, A., Fieler, R., White, P., Stefanussen, M., 2000: Potensialet for havbruk som envesentlig basisnæring i Nord-Norge. Rapportnr. STF38 A00611, Akvaplan-Niva & Sintef.

Stokke, O.S. 2001 (ed): Governing High Seas Fisheries: The Interplay of Global and RegionalRegimes. Oxford University Press, Oxford

Sundby, S., 2000. Recruitment of Atlantic cod stocks in relation to temperature and advection ofcopepod populations. Sarsia 85: 277-289.

Symes, D. 2000: Fisheries Dependent Regions. Fishing News books, Blackwell Science.

Page 30: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–30 DO NOT CIRCULATE OR CITE

Toresen, R. and Østvedt, O.J., 2000. Variations in abundance of Norwegian spring-spawningherring throughout the 20th century and the influence of climatic fluctuations. Fish andFisheries, 2000, 1: 231-256.

Øiestad, V., 1994. Historic changes in cod stocks and cod fisheries: Northeast Arctic cod. ICESmar.Sci.Symp., 198: 17-30.

Norwegian Central Statistical Bureau: http//www.ssb.noThe Directorate of Fisheries: http//www.fiskeridir.noStortingsmelding nr.51 (1997-98): perspektiver på utvikling av norsk fiskerinæring.

Page 31: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–31 DO NOT CIRCULATE OR CITE

12.3 Iceland-Greenland

12.3.1 IntroductionThis section deals with the marine ecosystems of Iceland and Greenland. Although there are largedifferences, both physical and biological, between these two ecosystems they are linked as will beexplained presently. Furthermore, the export of seafood produce represents a major proportion ofincome for exported goods for both countries.

An overview of the geographic positions and place names most frequently referred to in the text isgiven in Figure 12.3.1.

Insert Figure 12.3.1 here. The Iceland/Greenland ecosystems and adjacent areas. The blue lineindicates the position of the Polar Front

The waters surrounding Iceland and Greenland are different in many ways. Icelandic waters arewarmer due to greater Atlantic influence and generally ice free under normal circumstances.Exceptions are infrequent and usually short periods in late winter and spring when drift ice maycome close inshore and even become landlocked off the north and east coasts. In extreme cases,drift ice is known to have surrounded the island during spells of cold climate, e.g. in the winter of1918. On the other hand, Greenlandic waters are colder, ice conditions much more severe andports on the coastline commonly closed for long periods due to winter ice and icebergs.

The basic reason for treating these apparently dissimilar ecosystems together is the obvious linkbetween the all-important stocks of Atlantic cod (Gadus morhua) - at Iceland and Greenland. Thislink is manifested by documented drift of larval and 0-group cod (in its first year of life) fromIceland to Greenland with the western branch of the warm Irminger Current (Jensen 1926).Spawning migrations in the reverse direction were proven by numerous tagging experiments (e.g.Tåning 1934, 1937; Hansen et al. 1935; Jónsson 1996). There are, however, large variations in thenumbers of cod and other fish, which drift across from Iceland to Greenland and by no means allof these fish return to Iceland as adults.

The history of fishing the waters around Iceland and Greenland goes back hundreds of years but ismainly centered round Atlantic cod, which was the preferred species in northern waters in oldentimes. Icelandic waters are usually of a cold/temperate rather than pure Arctic nature and thereforerelatively rich in species. Consequently, with the diversification of fishing gear and vessel types inthe late 19th Century and the beginning of the 20th, numerous other fish species, both demersal andpelagic, began to appear in catches from Icelandic waters. On the other hand, the Greenlandicmarine environment is much colder and commercially exploitable species are therefore fewer.Present-day catches only comprise 9 demersal, 2 pelagic and 3 species of invertebrates. At thetime of writing, there is practically no catch of cod at Greenland.

Whale products figure in Icelandic export records from 1948 until the whaling ban (zero quotas)was implemented in 1986, but their value never was a significant part of exported seafood. InIceland there is, however, a long history of hunting porpoises, seals and seabirds and gatheringseabird eggs for domestic use. Although such hunting and gathering has gradually decreased withthe passage of time, it is still pursued as tradition in various coastal communities. In Greenland, on

Page 32: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–32 DO NOT CIRCULATE OR CITE

the other hand, several species of marine mammals - at least 5 different whale species, 5 species ofseals plus walrus (Odobenus rosmarus)) and 6 species of seabirds figure in Greenlandic catchstatistics. Contrary to Iceland, catches of marine mammals and seabirds still play an importantrole, both culturally and socially as well as in local economies, in Greenland.

12.3.2 Ecosystem essentialsThe marine ecosystem around Iceland is located just south of the oceanic Polar Front in thenorthern North Atlantic (see Fig. 12.3.1). The area south and west of Iceland is dominated bywarm and saline Atlantic water of the North Atlantic Current, the most important componentbeing its westernmost branch, the Irminger Current. The Irminger Current bifurcates off thenorthern west coast of Iceland. The larger part flows west across the northern Irminger Seatowards Greenland. The other and smaller branch is advected eastward onto the North Icelandicshelf where the Atlantic water mixes with colder waters of the East Icelandic Current, an offshootfrom the cold East Greenland Current. On the shelf north and east of Iceland the degree of mixingincreases in the direction of flow and the influence of Atlantic water is therefore lowest on the E-Icelandic shelf (see Chapter 8). Hydrobiological conditions are quite stable in the domain ofAtlantic water south and west of Iceland, while there may be large seasonal as well as inter-annualvariations of hydrography and bioproduction in the mixed waters on the N- and E-Icelandic shelf,depending on the intensity of the flow of Atlantic water and the proximity of the Polar Front.Large variations in the flow of Atlantic water onto the shelf area north of Iceland on longertimescales have also been demonstrated (Malmberg 1984, 1986; Malmberg and Kristmannsson1992; ; Malmberg et al. 1999; Vilhjálmsson 1997).The East Greenland Current carries polar water south over the continental shelf off the east coastof Greenland and, after rounding Cape Farewell (approx. 60°N; 43°W), continues north along thewest coast. Off the east coast, the temperature of these cold waters may be ameliorated by Atlanticwaters of the Irminger Current, especially near the shelf break and on the outer parts of the shelf.Off W-Greenland, the surface layer is dominated by cold polar water, while relatively warm mixedwater of Atlantic origin is found at depths between 150 and 800 m north to about 64°N. Mixingand diffusion of heat between these two layers, as well as changes of the relative strength of flowof these two main water components are fundamental in determining physical marine climaticconditions as well as primary and secondary production off W-Greenland (e.g. Buch and Hansen1988; Buch et al. 1994; 2002).

The warm Irminger Current is also highly important for both the Icelandic and Greenlandecosystems as a transport mechanism for juvenile stages of various species of fish (Fig.12.3.2).Thus, its eastern branch plays a dominant role in transporting fish fry and larvae from the southernspawning grounds to nursing areas on the shelf off NW-, N- and E-Iceland, while the westernbranch may in turn carry large numbers of larval and 0-group fish across the northern Irminger Seato E-Greenland and from there to nursery areas in southern W-Greenland waters. For illustrationof the main ocean currents in the Iceland/Greenland area see Figures 8.2.1.2 and 8.2.1.5 in Chapter8.

Insert Figure 12.3.2 here. The main water masses and larval drift paths in the Iceland-EastGreenland-Jan Mayen ares

Page 33: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–33 DO NOT CIRCULATE OR CITE

The Icelandic marine ecosystem contains plentiful stocks of zooplankton such as calanoidcopepods and krill, which are eaten by adult herring (Clupea harengus) and capelin (Mallotusvillosus), adolescents of numerous other fish species as well as by baleen whales. The larvae andjuveniles of both pelagic and demersal fish also feed on eggs and juvenile stages of thezooplankton. Benthic animals are also important in the diet of many fish species, especiallyhaddock (Melanogrammus aeglefinus), wolffish (Anarchias lupus), various flatfish and cod.

Owing to the influence of warm Atlantic water, the fauna of Icelandic waters is comparatively richin species and contains over 25 commercially exploited stocks of fish and marine invertebrates. Incontrast, the Greenlandic commercial fish and invertebrate fauna counts fewer species and ischaracterized by coldwater ones such as Greenland halibut (Hippoglossoides Reinhardtius),northern shrimp (Pandalus borealis), capelin and snow crab (Chionoecetes opilio). Redfish(Sebastes spp.) are also found, but mainly in Atlantic waters outside the cold waters of the EastGreenland continental shelf and cod can be plentiful at W-Greenland in warm periods.

At Iceland, most fish species spawn in the warm Atlantic water off the south and southwest coasts.Fish larvae and 0-group (1st year of life) drift west and then north from the spawning grounds tonursery areas on the shelf off NW-, N- and E-Iceland, where they grow in a mixture of Atlanticand Arctic water (e.g. Schmidt 1909). Larval and 0-group cod, capelin, as well as species likehaddock, wolffish, tusk (Brosme brosme) and ling (Molva molva) may also be carried by thewestern branch of the Irminger Current across to E-Greenland and onward to W-Greenland (e.g.Jensen 1926; Tåning 1937; Jensen 1939; see also Fig. 12.3.2). The drift of larval and 0-group codto Greenland was especially extensive during the 1920s and 1940s.

Capelin is the largest fish stock in the Icelandic marine ecosystem. Unlike other commercialstocks, adult capelin undertake extensive feeding migrations north into the cold waters of theDenmark Strait and Iceland Sea during summer. These capelin return to the outer reaches of theNorth Iceland shelf in October/November from where the migration to the spawning groundssouth and west of Iceland begins in late December/early January (Fig. 12.3.3). Spawning isusually over by the end of March. Capelin are especially important in the diet of small andmedium sized cod. Most juvenile capelin aged 0, 1, and 2 years reside on or near the shelf offnorthern Iceland and on the E-Greenland plateau west of the Denmark Strait (see Fig. 12.3.3).These parts of the stock are therefore accessible to fish, marine mammals, and seabirds throughoutthe year. On the other hand, the summer feeding migrations of maturing capelin into the colderwaters of the Denmark Strait and the Iceland Sea, place the larger part of the adult stock out ofreach of most fish, except perhaps Greenland halibut, for about 5–6 months. However, thesecapelin are then available to whales, Greenlandic seals and seabirds. During the feedingmigrations, adult capelin grow in weight by a factor of 3-4 and their fat content increases from afew percentage points to 15-20%. When the adult capelin return to the north Icelandic shelf inautumn they are preyed upon intensely by a number of predator species, apart from cod, until theend of spawning in the near-shore waters south and west of Iceland. Thus, adult capelin act as ahuge source of energy transfer from Arctic northern regions to important commercial fish stocks inIcelandic waters proper (Vilhjálmsson 1994, 2002).

Insert Figure 12.3.3 here. Distribution and migrations of capelin in the Iceland-Greenland-JanMayen area

Page 34: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–34 DO NOT CIRCULATE OR CITE

Off W-Greenland, northern shrimp and Greenland halibut spawn at the shelf edge off the westcoast. The same is true for the northern shrimp stock, which is found in the general area of theDohrn Bank, about mid-way between E-Greenland and NW-Iceland. Greenlandic waters alsocontain capelin populations that spawn at the heads of numerous fjords on the west and eastcoasts. These capelin populations appear to be self sustaining and local, feeding at the mouths oftheir respective fjord systems and over the shallower parts of the shelf area outside these fjords(Friis-Rødel and Kanneworff 2002). During the warm period from the early 1930s until late 1960sthere was also an extensive spawning of cod off SE-, SW-and W-Greenland (e.g. Buch et al.1994).

In the pelagic ecosystem off Greenland the population dynamics of calanoid copepods and to someextent krill are considered to play a key role in the food web as a direct link to fish stocks, baleenwhales (Mysticeti) and some important seabirds, such as little auk (Alle alle) and Brünnitch'sguillemot (Uria lomvia). Polar cod (Boreogadus saida), capelin, sand eel (Ammodytes spp.) andsquid ((Illex illecebrosus) are probably the most important pelagic/semi-pelagic macrofauna asforage for fish such as Greenland halibut and cod, mammals and seabirds. Benthic animals arealso important. In particular, northern shrimp is registered as a major food item for Atlantic codand many other species of fish and marine mammals (e.g. Anon. 2002a).

12.3.3 Fish stocks and fisheriesAtlantic codHistorically, demersal fisheries at Iceland and Greenland fall into two categories, i.e. land basedfisheries conducted by local inhabitants, and those of distant water foreign fleets. For centuries themain target species was cod. Until the late 19th century, the local fisheries were primarilyconducted with open rowboats, while the distant water fishing fleets consisted obviously of muchlarger, decked ocean going sailing vessels. Until the last decade of the 19th Century, almost allfishing for demersal species, whether from small open rowboats or larger ocean going sailingvessels, was carried out by hand lines.

Jónsson (1994) estimated that the combined landings by Icelandic, Dutch and French fishingvessels were in the order of 35 000 t annually in the period 1766-1777. One hundred years later,the combined French and Icelandic catches averaged about 55 000 t. Compared to the subsequentdevelopment of fishing effort and knowledge of stock sizes and exploitation rates, it is obviousthat even large fleets of several hundred sailing vessels and open rowboats, fishing with primitivehand lines, can not have had any serious effect on the abundant cod stock and other demersalspecies at Iceland.

This changed dramatically with the introduction of the steam- and combustion engines to thefishing fleet, as well as the adoption of active fishing gear at the turn of the 19th century. Thus, bythe beginning of the 20th century the otter trawl had been adopted among the foreign fleet, whilethe smaller motor powered Icelandic boats started using gill nets, long lines and Danish seines.Landings from the Icelandic area no longer consisted almost exclusively of cod, but species likehaddock, halibut (Hippoglossus hippoglossiedes), plaice (Pleuronectes platessa) and redfish(Sebastes marinus) also became common items of the catch. It has been estimated that thedemersal catch at Iceland increased from about 50 000 t in the 1880s to about 160 000 t in 1905,

Page 35: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–35 DO NOT CIRCULATE OR CITE

reaching 250 000 t just before World War I. Although cod was still the most important species, theproportion of other demersal species had increased to about 30% (see Fig. 12.3.4).

With the increasing effort and efficiency of the international distant water and local fishing fleets,cod catches in Icelandic waters increased to peak at 520 000 t in 1933, while the catch of otherdemersal species increased to about 200 000 t (Fig. 12.3.4). Catches declined during the late1930s, while the exploitation rate increased until the fishing effort fell drastically due to WorldWar II. Nevertheless, the exploitation rate of cod remained at a moderate level due to recruitmentfrom the superabundant 1922 and 1924 year classes (Schopka 1994).

Although cod has been fished intermittently off W-Greenland for centuries, the success of the codfishery at Greenland has been variable. In spite of patchy data from the 17th and 18th centuries,there is little doubt that cod abundance at W-Greenland fluctuated a great deal (see e.g. Buch et al.1994). Information from the 19th century suggests that cod were fairly plentiful in Greenlandicwaters until about 1850. After that there seem to have been extremely few cod on the banks and ininshore waters off Greenland until in the late 1910s-early 1920s, when an increase in theoccurrence of cod in inshore areas was noted (Jensen 1926, 1939; Hansen 1949). Cod were alsoregistered in offshore regions off W-Greenland in the late 1920s, where fisheries by foreignvessels expanded quickly and catches increased from about 5 000 t in 1926 to 100 000 t in 1930.From then until the end of World War II, this fishery yielded annual catches between about 60 000and 115 000 t (Fig. 12.3.5).

After World War II, catches of demersal fish from Icelandic grounds increased again. Landingspeaked at about 860 000 t in 1954, with cod accounting for about 550 000 t (Fig. 12.3.4). Becauseof the very strong 1945 cod year class and good recruitment to other demersal stocks, theexploitation rate of cod and other demersals remained at a low level, although almost 50% higherthan during the late 1920s and early 1930s. From 1955 the exploitation rate of all demersal stocksat Iceland, but especially that of cod, increased rapidly and with few exceptions remained far toohigh during the last 45 years of the 20th Century. Until 1976, this was due to a combined effort ofIcelandic and foreign distant water fleets. However, since the extension of the Icelandic EEZ to200 miles in 1977, the high rate of fishing has been due to the enhanced efficiency of Iceland’sfishing fleet.

At Greenland, the total cod catch had reached about 200 000 t by 1950 and then fluctuated around300 000 t during 1952-1961. After that the cod catch skyrocketed and landings varied betweenabout 380 and 480 000 t during the seven year period 1962-1968. By 1970 the catch had fallen to140 000 t and was, with large variations, within a range of 10 000-150 000 t until the early 1990s(Fig. 12.3.5). Since 1993 practically no Atlantic cod has been caught in Greenlandic waters.Before the introduction of the 200 miles EEZ around Greenland in 1978 the cod fishery wasmostly conducted by foreign fleets, but from then on the Greenlandic fleet has dominated thefishery (Fig. 12.3.5).

Page 36: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–36 DO NOT CIRCULATE OR CITE

Textbox 12.3.2The Iceland/Greenland cod and climate variability

Although the abundance of the Icelandic cod stock prior to 1920 is not known, it wasunquestionably large (e.g. Schmidt 1909). Furthermore, the climatic warming of the 1920s and1930s appears to have greatly increased reproductive success of Icelandic cod, as well asbioproduction in general in the mixed waters north and east of Iceland compared to previousdecades. In addition, the huge amounts of larval and 0-group cod that drifted west across thenorthern Irminger Sea in 1922 and 1924 and grew off W-Greenland, returned to Iceland in largenumbers to spawn (Schopka 1994; Vilhjálmsson 1997; Fig. T.12.3.1). Tagging experimentsindicate that the majority of these fish then remained within the Icelandic marine ecosystem(Tåning 1934, 1937; Hansen et al. 1935; Hansen 1949; Jónsson 1996; Jakobsson 2002). Thus, thedistribution area and biomass of cod in the Icelandic marine ecosystem can be enormouslyenlarged through larval drift during warm periods.

Figure T.12.3.1. Spawning stock biomass (SSB) and recruitment at age 3 of Icelandic cod 1920-2002. The yellow part of the columns denotes immigrants from Greenland.

Greenland, also experienced a climatic warming in the 1920s (cf. Fig. 12.3.10), which resulted infar greater changes in the distribution and abundance of cod there than at Iceland. In the last yearsof the 19th Century and until the 1920s, cod occurred only in very scattered numbers in inshorewaters near Cape Farewell, the southernmost promontory of Greenland (Jensen 1926. 1949;Tåning 1948). In the 1920s, cod gradually appeared over wider areas and in increasing numbers.This is reflected in the rapid rise in the catch of cod at W-Greenland in the late 1920s (see Fig.12.3.4). Furthermore, in the 1920s and 1930s, cod extended their distribution northward along the

Page 37: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–37 DO NOT CIRCULATE OR CITE

west coast of Greenland by some 600-800 nautical miles (Tåning 1948; Fig. T.12.3.2). At E-Greenland, cod is said to have appeared in small schools in the Ammassalik area about 1920 andaround 1930 to have occurred everywhere along the east coast of Greenland south fromAmmassalik (Schmidt 1931; Fig. T.12.3.2).

Figure T.12.3.2. Changes in distribution, larval drift and migrations of cod at Iceland andGreenland in the 1920s and 1930s. Dark red: traditional spawning grounds at Iceland; light red:additional spawning areas at Iceland. The colored areas describe the increase of cod distribution atGreenland in the 1930s. Blue arrow: larval and 0-group drift; Green arrows: feeding migrations;red arrow: spawning migrations. (Adapted from Vilhjálmsson 1997).

The drift of 0-group cod from Iceland to Greenland continued intermittently from the 1930s untilthe mid-1960s, although on a smaller scale compared to the superabundant year classes of 1922,1924 as well as that of 1945 (Schopka 1994; Fig. T.12.3.1). However, after about 1930, cod alsoreproduced very successfully in W-Greenlandic waters (Jensen 1939; Hansen 1949; Hansen et al.1935; Tåning 1937; Buch et al. 1994; Fig. T.12.3.3). Wile the establishment of a self-sustainingcod stock at Greenland was most likely initiated first and foremost by fish that drifted as juvenilesof the 1922 and 1924 year classes from Iceland, it is possible that the small inshore componentalso played a part (Buch et al. 1994).

The sudden and unexpected cooling of the Nordic Seas in the mid-1960s was also recorded off W-Greenland (Fig. 12.3.11). Following this cooling, the drift of 0-group cod from Iceland toGreenland ceased completely for a number of years and has since then only been registered twice,i.e. in 1973 and 1984 (Vilhjálmsson and Fri_geirsson 1976; Vilhjálmsson and Magnúson 1984;Schopka 1994; see also Fig. T.12.3.1). Another event, associated with the cooling from the mid-1960s onwards, was the large reduction in the total cod catch from W-Greenland waters fromabout 240 000-470 000 t annually in the period 1950-1969 to 75 000 t in 1973 (cf. Fig. 12.3.5).

Page 38: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–38 DO NOT CIRCULATE OR CITE

From 1973 to 1993 the average annual catch of cod off W-Greenland was about 55 000 t. Peaks incatches in this period have been associated with year classes which drifted as 0-group fromIceland to Greenland. Presently, there are few cod at E- and W-Greenland and no local recruitmentto the cod stock (Buch et al. 2002; Fig. T.12.3.3).

Figure T.12.3.3. Spawning stock biomass (SSB) and recruitment at age 3 (R3) for cod atGreenland 1920-2000. (From Buch et al. 1994; 2002).

As described earlier, there were practically no cod in Greenlandic waters prior to the onset ofwarming in the 1920s and 1930s. In spite of Danish fishing experiments in the early 20th century,the few cod found and caught were inshore. This did not change much until the late 1920s, whichcorresponds to the time necessary for the very large cod year classes from 1922 and 1924, whichundoubtedly were of Icelandic origin, to reach marketable size (Fig, T.12.3.4).

Page 39: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–39 DO NOT CIRCULATE OR CITE

Figure T.12.3.4. Temperature deviations and the catch of cod at Greenland 1910-1940.

However, it is equally clear that by the early 1930s W-Greenland waters had become warmenough for successful spawning of cod. Subsequently, some of the fish of the 1922 and 1924 yearclasses took advantage of this, spawned off W-Greenland and, along with the small inshore codpopulation, were instrumental in giving rise to a local self-sustaining component. The Greenlandcod component indeed became very large and sustained annual catches in the range of 300-470000 t for almost two decades in the 1950s and 1960s (e.g. Buch et al. 1994; cf. Fig. 12.3.5).

Although fishing mortalities of cod at Greenland increased in the 1950s and 1960s, the spawningstock remained above 500 000 t until 1970 and produced large year classes until 1964 (Fig.T.12.5.3). Like at Iceland, there was a severe cooling of the Greenlandic marine environment inthe latter half of the 1960s and since then the only year classes at Greenland of any commercialsignificance are those of 1973 and 1984, both of which drifted to Greenland as 0-group fromIceland. In spite of somewhat warmer climatic conditions of Greenlandic waters after the coolingof the late 1960s, no cod year classes of Greenlandic origin have appeared. Therefore, all theavailable evidence indicates that cod cannot reproduce efficiently at Greenland except underhydrographic conditions, which usually are classified as warmer than ‘normal’.

During the second half of the 20th century, the fishable part of the Icelandic cod stock (age 4+)declined from almost 2.5 million t in the early 1950s to about 550 000 t in year 2000 (seeFig.12.5.6). The spawning stock decreased from about 1 260 to 235 000 t during the same period.The initial very large stock size was not only due to the low fishing pressure during World War IIand in the following few years, but also to the recruitment of the superabundant year class of 1945.A very large part of this year class drifted across to Greenland as 0-group and grew in Greenlandicwaters. Later, about 500 million individuals of the 1945 year class migrated back to Iceland forspawning and appear not to have left again (Schopka 1994; see Fig. T.12.3.1). In spite of the coldperiod 1965-1971 and warmer but more variable conditions since then, recruitment remained at a

Page 40: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–40 DO NOT CIRCULATE OR CITE

normal level until 1985, with occasional boosts by immigrants from Greenland, although on amuch smaller scale than in 1922, 1924 and 1945 (Schopka 1994).

For most of the 20th century for which data are available, recruitment variability of cod that grewwithin the Icelandic ecosystem has been low or about 1:4 for 1920-2002 (Fig. T.12.3.2). It seemstherefore that while the Icelandic ecosystem cannot support juvenile cod year classes muchbeyond sizes corresponding to 300 million recruits at age 3, it has easily accommodated even verylarge numbers of adult cod migrating back from Greenland to their natal spawning grounds.

It is instructive to look at some simple calculations of S/R relationships of Icelandic cod in the 20th

century.

Table T.12.3.1. Average recruitment of cod at age 3 (R3) from Icelandic nursery areas, spawningstock biomass (SSB) and total stock biomass (TSB) in different periods during 1920-2000.

It appears that even the very cold period from 1965-1971, and the variable conditions since then,did not have much detrimental effect on recruitment to the cod stock by fish that grew up locally.However, since 1985, a small and young spawning stock, resulting in lower quality eggs, shorterspawning time, smaller spawning grounds and possibly different drift routes, seems to be the mostplausible explanation for the large and protracted decline of recruitment (Marteinsdottir andSteinarsson 1998; Marteinsdottir and Begg 2002). The most likely explanation of the large yearclasses of 1983 and 1984, which derived from small spawning stocks, is that old fish of theabundant year classes of 1970 and 1973 were still present in the spawning stock in sufficientnumbers to enhance recruitment.

Greenland halibutAn Icelandic Greenland halibut fishery began in the early 1950s (Fig. 12.3.4). Initially, long linewas the main fishing gear but this method was abandoned because killer whales removed morethan one half of the catch from the hooks. Since the early 1970s this fishery has been conductedusing otter trawls.

At Greenland a fishery for Greenland halibut began in a very modest way around 1915 and had by1970 only reached an annual catch of about 4 000 t, most of which taken by Greenland. From1970 to 1980 other countries participated in the Greenland halibut fishery, which peaked in 1976at about 26 000 t. By 1980 the catch had fallen to about 7 000 t. During the 1990s the catchincreased rapidly to about 23 000 t in 1992 and a record high of about 35 000 t in 1999. After1980, foreign vessels have not played a significant role in the Greenland halibut fishery off W-

Period R3 Iceland Average SSB Average TSB Average 1920-1964 205 1188 1939 Avegage 1965-1971 200 1193 1559 Average 1972-1985 207 355 1068 Average 1920-1985 205 922 1722 Average 1986-2002 137 276 748

Page 41: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–41 DO NOT CIRCULATE OR CITE

Greenland. The catch history of Greenland halibut in W-Greenland waters is shown in Figure12.3.5.

Northern shrimpA small inshore fishery for northern shrimp began in Icelandic waters in the mid-1950s. Initially,this was a fjordic fishery of high value to local communities. An offshore shrimp fishery, whichbegan in the mid-1970s on the outer shelf off the western north coast, soon expanded to moreeastern areas. Annual landings from this fishery increased to 25-35 000 t in the late 1980s and to45-75 thousand t in the 1990s. Recently, catches have declined drastically, both in offshore andcoastal areas (Fig. 12.3.4).

The catch of northern shrimp off W-Greenland has increased steadily since it’s beginning in 1960.At the outset, this species was fished only by the Greenlandic fleet, but from 1972 large vesselsfrom other countries joined this fishery. This lead to a large increase in the total catch of northernshrimp, which peaked at about 20 000 t in 1976. Between 1976 and the early 1980s, the catch byother countries decreased and has been insignificant since. On the other hand, the Greenlandiccatch increased steadily, from a total catch in 1960 of about 1 800 t to 132 000 t in 2002 as shownin Figure 12.3.5.

HerringAt Iceland, commercial fishing for herring started in the 1860s when Norwegian fishermeninitiated a land-based fishery on the north and east coasts using traditional Scandinavian beachseines. This fishery proved very unstable and was abandoned in the late 1880s. Drift netting wasintroduced at the turn of the 19th Century and purse seining in the early 20th Century (1904). Thelatter proved very successful off the north coast, where the herring schools used to surfaceregularly, while drift nets had to be used off the south and west coasts where the herring rarelysurfaced. The north coast herring fishery increased gradually during the 1920s and 1930s and hadreached 150 000-200 000 t by the beginning of the 1940s (Fig. 12.3.4). During this period, thefishery was limited mainly by lack of processing facilities. Around 1945 the herring behaviorpattern changed and as a result purse seining for surfacing schools north of Iceland becameineffective and catches declined. The reasons for this change of behavior have never beenidentified.

Horizontally ranging sonar, synthetic net fibers and hydraulic power blocks for hauling the largeseine nets, were introduced to the herring fishery during the late 1950s and early 1960s (Jakobsson1964; see also section 12.2). These technical innovations, as well as better knowledge of themigration routes of the great Atlanto-Scandian herring complex (Norwegian spring spawningherring and much smaller stocks of Icelandic and Faroese spring spawners), lead to aninternational herring boom in which Icelandic, Norwegian, Russian (USSR) and Faroesefishermen were the main participants (for Icelandic catches see Fig. 12.3.4). This extraordinaryherring fishery ended with a collapse of the Atlanto-Scandian herring complex during the late1960s due to overexploitation of both adults and juveniles (cf. section 12.2. Textbox). Catches ofAtlanto-Scandian herring (now called Norwegian spring spawning herring since the Icelandic andFaroese components have not recovered) in the Icelandic area have been negligible since the late1960s and Iceland’s share of the TAC of this herring stock since the mid-1990s has mainly beentaken outside of Icelandic waters. There is no fishery for herring at Greenland

Page 42: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–42 DO NOT CIRCULATE OR CITE

It took the Norwegian spring spawning stock about two and a half decades to recover in spite ofsevere catch restrictions (see section 12.2 Textbox). Both the Icelandic spring and summerspawning herring suffered the same fate. Retrospective analysis of historical data shows that therewas no more than 10-20 000 t left of the Icelandic summer spawning herring stock in the late1960s/early 1970s (Jakobsson 1980). A fishing ban was introduced and since 1975 the fishery hasbeen regulated, both by area closures and minimum landing size, as well as a catch rulecorresponding to a TAC of roughly 20% of the estimated adult stock abundance in any given year.The stock recovered gradually, is at a historical high at present and the annual yield during the lasttwo decades of the 20th century was on average about 100 000 t.

CapelinAn Icelandic capelin fishery began in the mid-1960s and within a few years replaced the rapidlydwindling herring fishery as also witnessed for the Barents Sea (Vilhjálmsson 1994, 2002;Vilhjálmsson and Carscadden 2002). The capelin fishery is conducted by the same high-technology fleet as used for catching herring. During the first 8-10 years, the capelin fishery onlypursued capelin spawning runs in near-shore waters off the southwest and south coasts of Icelandin February and March and annual yields increased to 275 000 t. In 1972 the fishery was extendedto deep waters east of Iceland in January, resulting in an increase of the annual catch by about 200000 t. In 1976, an oceanic summer fishery began north of Iceland and in the Denmark Strait. In1978 the summer fishery became international as it extended north and northeast into to the EEZsof Greenland and Jan Mayen (Norway). Within two years the total seasonal (July-March) capelincatch increased to more than one million t. Total annual international landings of capelin from thisstock during 1964-2002 are shown in Figure 12.3.4.

Historically, capelin have been caught at Greenland for domestic use and animal fodder. A smallcommercial fishery for roe bearing females began at W-Greenland in 1964 with a catch of 4 000 t,which is also the largest catch on record. There were relatively large fluctuations of the capelincatch from 1964 to 1975, but since then the catch of capelin has been insignificant. This fishery isconducted by Greenlanders.

Blue whitingThe most recent addition to Icelandic fisheries is that of the semi-pelagic blue whiting(Micromesitius poutassou). This is a straddling species commonly encountered in that part of theIcelandic ecosystem dominated by Atlantic water, i.e. off the west, south and southern east coast.A small blue whiting fishery began in the early 1970s, increased to about 35 000 t in 1978 andthen dwindled to 105 t in 1984. There was a renewed interest in this fishery in the mid-1990s andin 1997-2002 the blue whiting catch increased from 10 000 t to 285 000 t (see Fig. 12.3.4).

Fisheries off East GreenlandEast-Greenland waters have been fished commercially only since after World War II. The mainreason is rough bottom as well ad the speed and irregularity of ocean currents, especially near theedge of the continental shelf. These conditions render it difficult to fish E-Greenland waters exceptwith large powerful vessels and robust fishing gear. The main species, which have been fishedcommercially off E-Greenland, are Greenland halibut, northern shrimp, cod and redfish. With the

Page 43: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–43 DO NOT CIRCULATE OR CITE

exception of northern shrimp in the last two decades, the fisheries off E-Greenland have almostexclusively been conducted by foreign fleets (Fig. 12.3.6).

Marine mammals and seabirdsThe Icelandic marine ecosystem contains a number of species of large and small whales, most ofwhich are migratory. Commercial whaling has been conducted intermittently in Iceland for almosta century. Initially, large Norwegian whaling stations were operated from the mid-1880s untilWorld War I, first on the Vestfirdir peninsula (NW-Iceland) and later on the east coast. By about1912, stocks had become depleted to the extent that whaling was no longer profitable and in 1916the Icelandic Parliament passed an act prohibiting all whaling. In the following decades whalestocks gradually recovered and from 1948, until zero quotas on whaling were set in 1986, a smallIcelandic company operated with four boats from a station on the west coast, just north ofReykjavík. The main target species were fin (Balaenopter physalus), sei (Balaenopter borealis)and sperm (Physeter catodon) whales and the average yearly catches were 234, 68 and 76 animalsrespectively. In addition, 100-200 (average 183) minke whales (Balaenoptera acutorostrata)where taken annually by small-time operators in the years 1974-1985. Although nevereconomically important on a national scale, whaling was highly profitable for those engaged in theindustry. Icelandic whale catches by species are shown in Figure 12.3.7).

The numbers of seals in Icelandic waters are comparatively small. The populations of the twomain species, common seal (Phoca vitulina) and grey seal (Halichoerus grypus), are estimated at15 000 and 6 000 animals, respectively (Anon. 2003). The abundance of the former is stable whilethe numbers of grey seals have decreased. Sealing has never reached industrial proportions inIceland, the total number of skins varying between 1 000 and 7 000 annually during the last fourdecades of the 20th century.

Although foreign fleets have pursued large-scale whaling in Greenlandic waters in past centuries,native Greenlanders have hunted whales only for domestic use. The harvest of the main species,has been modest and is unlikely to have had any effect on the stocks. Five seal species areexploited in Greenland. Two (harp (Phoca groenlandica) and ringed Pusa hispida) seals) are byfar the most important. Ringed seal catches increased from the mid-1940s until the late 1970s,then dropped until the mid-1980s, whereafter they increased again. The harp seal catchesincreased in the first half of the century, but started to drop during the 1960s and were very lowduring the 1970s. Since then, harp seal catches have increased continuously and are higher thanever at the time of writing.

Greenlandic catches of whales, seals, walrus and sea birds in the years 1993-2000 are shown inFigure 12.3.8. Sealskin prices were subsidized in Greenland when prices started to decline on theworld market and there is no reason to believe that sealskin campaigns have influenced huntingeffort for seals in Greenland. There have, however, been indirect positive effects, because theCanadian catches of both species fell dramatically (shared populations) and the harp sealpopulation increased to double its size within relatively few years. The decrease in ringed sealcatches during the early 1980s coincided with the sealskin campaign, but the underlying reasonwas probably population dynamics, triggered by climatic fluctuations (Rosing-Asvid in prep.).

Aquaculture

Page 44: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–44 DO NOT CIRCULATE OR CITE

In the late 1970s and the 1980s there was much interest in aquaculture in Iceland. A number offacilities were erected for the cultivation of salmon (Salmo salar), rainbow trout (Salmo gairdneri)and Arctic char (Savelinus alpinus) at various locations on the coast. Practically all of these failed,either for financial reasons or lack of expertise, or both. The few that survived, or were rebuilt onthe ruins of others, have until recently not produced much more than necessary for the domesticmarket.

In comparative terms, aquaculture has therefore been of little economic importance for Iceland inthe past. However, there was a renewed interest in aquaculture starting in the 1990s. While salmonand Arctic char continue as the main target species, accounting for over 95% of the total producedin 2001 (about 4000 t), it is expected that production of other species will increase at aproportionally higher rate in the coming years.

Today, Iceland is once again investing heavily in fish farming - but this time it is private capitalrather than State funding which governs the progress. The largest quantitative increase will almostcertainly be in salmon. The total production in 2001 was around 4 000 of salmon and relatedspecies. It is expected that by 2010 the production of these species will have increased to 25-30000 t. In addition there is increased interest and success in the farming of Atlantic halibut, oceanperch (Disentrachus labrax), turbot (Psetta maxima), cod, and some other marine fish and recentlythere has been a considerable increase in the production of abalone (Haliotis rufucens) and bluemussel (Mytilus edulis)

Despite the fact that fish farmers are working closely with the industry and researchers toaccelerate growth in both salmonids and whitefish species, it is not until after a few more yearsthat the industry is expected to be operating smoothly. Area conflicts with wild salmon have notbeen resolved, cod farming is still at the fry stage and char - a high price product - has a limitedmarket. Nevertheless, fish farming is being developed to something more than an extra source ofincome and because of that, major fisheries companies are investing in development projects inthis sector.

Aquaculture was attempted in Greenland in the 1980s. The experiment failed and aquaculture isnot conducted in Greenland at the present time.

12.3.4. Historic climate variations and related changes of marine ecology and commercialstocksDuring the 20th century, the main change in climate over the Nordic Seas and in the northwestNorth Atlantic was a rise in air temperature during the 1920s and 1930s with a concurrent increasein sea temperature and a decrease of drift ice. There was distinct cooling in the 1940s and early1950s followed by reversal to conditions similar to those of the 1920s and 1930s. These changesand their apparent effect on marine biota and commercial stocks in Icelandic and Greenland waterswere studied and reported on by a number of contemporary researchers (e.g. Jensen 1926, 1939;Sæmundsson 1934; Tåning 1934, 1948; Fridriksson 1948). Summaries have been given by e.g.Buch et al. (1994) and Vilhjálmsson (1997).

Figure 12.3.8 shows 3-year running averages of sea surface temperatures off the central north

Page 45: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–45 DO NOT CIRCULATE OR CITE

coast of Iceland and illustrates trends in the physical marine environment of Icelandic waters inthe 20th century. The main features are increased flow of Atlantic water onto the shelf north andeast of Iceland during 1920-1964 followed by a sudden cooling in 1965-1971 and more variableconditions since then. A strong presence of Atlantic water on the North- and East-Icelandic shelfpromotes vertical mixing and thus favors both primary and secondary production, i.e. prolongsalgal bloom and increases biomass of zooplankton. Greenland, also experienced a climaticwarming in the 1920s probably with similar effects on the lowest links on the food chain (Fig.12.3.10)

Ai Iceland, one of the most striking examples of the effects of the climatic warming during the1920s was a mass spawning of cod off the north and east coasts in addition to the usual spawningoff South and West Iceland (Sæmundsson 1934). Furthermore, there was large-scale drift of larvaland 0-group cod (in first year of life) across the northern Irminger Sea to Greenland in 1922 and1924 (Jensen 1926; Schopka 1994). At Iceland, the spawning grounds of herring also expandedand after 1928 capelin became practically absent from their usual spawning areas off S- and W-Iceland for some years, but spawned instead in fjords and bays on the north coast as well as in thevicinity of Hornafjördur on the southeast coast (Sæmundsson 1934). Finally, several fish speciessuch as bluefin tuna and mackerel, usually associated with warm waters and rare at Iceland,increased in occurrence and abundance (Fridriksson 1948).

The changes in the marine fish fauna at West-Greenland were even more spectacular than those atIceland. There was a large increase of cod abundance and catches in the 1920s and other gadids,such as saithe, haddock, tusk and ling, previously rare or absent at Greenland, also appeared there inthe 1920s and 1930s. Furthermore, herring appeared in large numbers at W-Greenland in the 1930sand began to spawn there in the period July-September, mainly south of 65°N (Jensen 1939). Theseherring spawned near beaches, similar to capelin in these waters. Like capelin, herring are bottomspawners, their eggs adhere to the substrate or even the fronds of seaweed as in this case. In 1937the northernmost distribution of adult herring reached 72°N (Jensen 1939). However, a herringfishery of commercial scale has never been pursued at Greenland.

In the early 1900s capelin were very common at W-Greenland between Cape Farewell and DiskoBay (see Fig. 12.3.1), but unknown farther north (Jensen 1939). In the 1920s and 1930s, thecenter of the W-Greenland capelin populations gradually shifted north and capelin became rare intheir former southern area of distribution. By the 1930s the main spawning had shifted north bysome 400 nautical miles to the Disko Bay region (cf. 12.3.1). At E-Greenland capelin are said tohave gradually extended their distribution to the north along the east coast to Ammassalik (Jensen1939). However, capelin are an Arctic species and have probably been common in that area forcenturies past since Ammassalik means the place of capelin (cf Fig. 12.3.1).

During the latter half of the 1960s there was a sudden and severe climatic cooling with anassociated drop of sea temperature, salinity and plankton production (Fig. 12.3.11), as well asincreased drift ice in the area north and east of Iceland. Temperatures increased again in the 1970s,but were more variable in the last three decades of the 20th century than during the warm period.The low sea temperatures were also recorded in W-Greenland waters (Fig. 12.3.9). This lowtemperature, low salinity water (the Great Salinity Anomaly) drifted round the North Atlantic and

Page 46: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–46 DO NOT CIRCULATE OR CITE

had noticeable, and in some cases serious, effects on various marine ecosystems (reviewed e.g. inJakobsson 1992).

In the Icelandic area, herring was the fish species most affected by the environmental adversitiesof the 1960s (Jakobsson 1969, 1978, 1980; Dragesund et al. 1980; Jakobsson and Østvedt 1999).This is not surprising since herring are plankton feeders and in North-Icelandic waters near theirlimit of distribution. This was both manifested in large-scale changes of migrations anddistribution (see Figure XX in Chapter 8 – Marine Systems) and a sudden and steep drop inabundance, which however was mostly brought about by overfishing (cf section 12.2. Textbox).The abundance of the Norwegian spring spawning herring stock increased dramatically in the1990s (cf. section 12.2.4) and these herring resumed some semblance of their previous feedingpattern (for an overview of these changes see also Chapter 8). Presently, the Norwegian springspawning herring still overwinter in the Lofoten area on the northwest coast of Norway. Whetherand when they revert completely to the ‘traditional’ distribution and migration pattern cannot bepredicted.

The two Icelandic herring stocks, i.e. the spring- and summer spawners, suffered the same fate.The spring spawners still show no sign of recovery, while the summer spawning stock recoveredin a few years after a fishing ban was imposed in the early 1970s. It seems that, like the W-Greenland cod, the Icelandic spring spawning herring had difficulties in self propagation in coldperiods and would most likely have collapsed in the late 1960s and early 1970s, even without afishery (Jakobsson 1980). The summer spawning herring, on the other hand, have adapted muchbetter to climatic variability of Icelandic waters. For all three stocks it can be concluded thatenvironmental adversities placed them under reproductive stress and disrupted feeding andmigration patterns. Environmental stress, coupled with a far too high fishing pressure on bothadults and juveniles, resulted in the actual collapses of these herring populations.

While the growth rate of Icelandic capelin has shown a significant positive correlation withtemperature and salinity variations in the North-Icelandic area in the past 25 years, thisrelationship probably describes feeding conditions in the Iceland Sea rather than direct effect oftemperature (Vilhjálmsson 1994. 2002; Astthorsson and Vilhjálmsson 2002). Results of attemptsto relate recruitment of the Icelandic capelin stock to physical and biological variables, such astemperature, salinity and zooplankton abundance, have been ambiguous. Nevertheless, judging bytheir stock size, the Icelandic capelin, which spawn in shallow waters off the south and west coastsof Iceland, seem to have been quite successful in recent decades and probably also in most yearsduring the latter half of the 20th century.

However, at the peak of warming in the late 1920s and in the first half of the 1930s, it was notedthat capelin had ceased to spawn on the traditional grounds off the south and west coasts ofIceland and spawned instead off the easternmost part of the south coast as well as in fjords andinlets on the southeast and north coasts (Sæmundsson 1934). Sæmundsson also noted that the codhad become unusually lean and attributed this to lower capelin abundance. Although there can beother reasons affecting growth of cod, e.g. competition due to a large stock size, Sæmundsson’sconclusion may well have been right. The change of capelin spawning areas he described, isprobably disadvantageous for this capelin stock. The reason is that suitable spawning areas wouldbe much reduced as compared to those previously and presently occupied by the stock.

Page 47: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–47 DO NOT CIRCULATE OR CITE

Furthermore, larval drift routes could be quite different and part of the larvae would likely end upin the western Norwegian Sea and be carried far afield, where their survival rate might be muchlowered.

The catch history and series of stock assessments of northern shrimp in deep waters northwest,north and east of Iceland as well as at Greenland are too short for attempting to establish linkageswith environmental variability. Being a frequent item in the diet of small and medium sized cod,stocks of northern shrimp would be expected to be larger when cod abundance is low. In generalterms though, the stock probably benefits from cooler marine climate, possibly both throughenhanced recruitment and reduced overlap of shrimp distribution with that of cod.

12.3.5 Possible impacts of global warming on fish stocksTo predict what will happen in a marine ecosystem in connection with climate warming is adifficult and daunting task, even in light of earlier experiences of such changes. In previoussections it was described how the marine climate around Iceland changed during the 20th century,i.e. from cold to a warm state in the 1920s, lasting with some deviations for about 45 years. Thenthere was a sudden cooling in 1965 lasting until 1971. After that, conditions have been warmer butvariable and temperatures not risen to the 1925-1964 levels. As described earlier, the availableevidence from the 20th century suggests that, as a general rule, plankton production and therebythe carrying capacity of the Icelandic marine ecosystem is enhanced in warm periods, while lowertemperatures have the reverse effect. Within limits, this is a reasonable proposition since thenorthern and eastern parts of the Icelandic marine ecosystem border on the oceanic Polar Front,which may be located close to the coast in cold years but is pushed quite far offshore in warmperiods when bioproduction is enhanced through nutrient renewal and associated mixingprocesses, due to increased flow of Atlantic water onto the north and east Icelandic plateau.

In the last few years it has been noted that the salinity and temperature of Atlantic water off S- andW-Iceland have increased and approached the pre-1965 levels. At the same time, there have beenpositive signals of increased flow of Atlantic water onto the mixed water areas over the shelf northand east of Iceland in spring and, in particular, also in late summer and autumn. This might wellbe the beginning of a period of increased presence of Atlantic water resulting in highertemperatures and increased vertical mixing over the N-Icelandic plateau, but the series is still tooshort for a conclusion to that effect.

There are, however, many other parameters, which can determine how an ecosystem and itsinhabitants, especially those of the upper ranges of the food chain, react to changes of temperature,salinity and the associated variations of primary and secondary production. Two of the mostimportant and obvious ones are stock sizes and fisheries, which of course are interconnected. Dueto high fishing pressure in the last three and a half decades of the 20th century, the majority of themost important commercial fish stocks in Icelandic waters are smaller than they used to be, evenat the onset of the climatic warming in the 1920s. Associated with this are changes in age and sizedistributions of spawning stocks, in that spawners are now fewer, younger and smaller than then.All of these changes can influence reproductive success through retracted spawning areas andduration of spawning, smaller eggs of lower quality and changes of larval drift routes and survivalrates (Marteinsdottir and Steinarsson 1998; Marteinsdottir and Begg 2002). In fact it is unlikelythat reactions of commercial fish stocks to a warming of the marine climate at Iceland, similar to

Page 48: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–48 DO NOT CIRCULATE OR CITE

that of the 1920s and 1930s, will be the same in scope, magnitude and speed as they were at thattime. Nevertheless, a moderate warming will in all likelihood improve survival of larvae andjuveniles of most species and thereby contribute to increased abundance of commercial stocks ingeneral. The magnitude of these changes will, however, be no less dependent on the success offuture fishing policies in enlarging stock sizes in general and spawning stock biomasses inparticular, since the carrying capacity of Icelandic waters is probably about 2-3 times greater thanthat needed by the biomass of commercial species in the area at the present time.

ScenariosIn the following we describe and discuss three possible future scenarios of warming of the marineecosystems of Iceland and Greenland and try to evaluate likely associated biological andsocio/economic changes in both of these countries:

1) No change from the ACIA climate reference period 1981-2000.Although the marine climate may dictate year class success in some instances, there is little if anyevidence that year class failure and thereby stock propagation can primarily been blamed onclimate related factors. Therefore, the development and possible yield in biomass of commercialstocks will in most cases depend on effective rational management, i.e. a management policyaiming at increasing the abundance of stocks through reduced fishing mortalities and protection ofjuveniles. This is the Icelandic present day policy. Unfortunately, it has not resulted in muchtangible success as yet, but should eventually do so and with a speed that largely depends on howwell incoming year classes of better than average size can be protected from being fished asadolescents.

Given a successful fishing policy of this kind, it seems reasonable to assume an increase inabundance of many demersal fish stocks by the end of the first three decades of the 21st century.This would considerably increase the sustainable yield from these stocks compared to the presentsituation. The same could also apply for the Icelandic summer spawning herring although thatstock is already exceeding its historical maximum abundance. The increase in yield in tonnes is,however, not directly proportional to increase in stock abundance. Thus, a doubling of the fishablebiomass of the Icelandic cod stock would probably increase its long-time sustainable yield in t byabout 20-30% as compared to the present annual catch of about 200 000 t. Furthermore, due tonatural variability in the size of recruiting year classes, increases of stock biomasses of the variousspecies will most likely occur in a stepwise fashion and the value of the catch would notnecessarily increase proportionally.

On the negative side, it is likely that the catch of northern shrimp would be reduced due toincreased predation by cod and the capelin summer/autumn fishery would have to be scaled downor curtailed altogether, in order to meet the needs of their more valuable fish predators and largewhales, if whales are left untouched. Developments of the abundance, but especially extendedmigrations of the Norwegian spring spawning herring to feed in north Icelandic waters, willdetermine the value of the yield from that stock for Iceland. For this to happen on a lasting basis,the intensity of the cold East Icelandic Current has to weaken and temperatures north of Iceland toincrease. This is not envisaged in this scenario.

Page 49: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–49 DO NOT CIRCULATE OR CITE

At Greenland, the no change scenario will have little effect on the present situation, given thatstocks are presently managed in a rational manner and continue to be so.

2) Moderate warmingMost criteria in the above no-change scenario are probably also valid for a moderate warming of1-3 °C. However, due to greater plankton production and a direct temperature effect per se, stock-rebuilding processes will likely be accelerated in most cases. Nevertheless, as for scenario 1, arational fishing policy must be maintained. Indeed, it is very likely that harvesting strategies canbe devised and implemented which would give higher returns from most of the major demersalstocks in the Icelandic area. Like in the no-change scenario, an automatic side effect of such apolicy would be a rise in the mean age and number of older fish in the spawning stock of cod,which would further enhance larval production and survival.

Drift of larval and 0-group cod across the northern Irminger Sea to E-Greenland and onwards toW-Greenland waters is expected to become more frequent and the number of individualstransported to increase as compared to the latter half of the 20th century. However, it is unlikelythat this will contribute much to cod abundance at Iceland. The reason for this is that the fishfinding and catch technologies of today are so effective that these cod can, and most likely will,easily be fished in Greenlandic waters before they would return to Iceland for spawning at the ageof 7-8 years.

An increase in temperature of 1-3 °C in the north Icelandic area is quite large in comparative termsand will i.a. be associated with a weakening of the East Icelandic Current and a considerablereduction of its domain. The degree of reduction is expected to be large enough to enable theAtlanto-Scandian herring (Norwegian spring spawners) to again take advantage of the rich supplyof Calanus finmarchicus over the north Icelandic shelf. Such a scenario would not only make iteasier and cheaper for Iceland to take its share of this stock, but also make it more valuable. Thereason is a large increase of that proportion of the catch, which can be processed for humanconsumption as compared to the current situation where a high proportion has to be reduced to thecomparatively much cheaper fishmeal and oil. It is also probable that more southern species likemackerel and tuna will enter Icelandic waters in large enough concentrations for commercialfishing in late summer and autumn.

3) Large climatic warmingAccording to the B2 scenario, model results indicate that a rise in temperature beyond 2-3 ˚C inthe Icelandic area in the 21st century is unlikely. Should that nevertheless happen, the hightemperature could lead to dramatic negative changes of the Icelandic marine ecosystem. The keyrole of capelin for the well being of many demersal stocks was described earlier, including to thelarge reduction in weight at age of Icelandic cod during the two capelin stock collapses on record.It has also been noted that capelin spawning ceased on their traditional grounds off the south andwest coasts of Iceland in the late 1920s and early 1930s and spawned instead in fjords and inletson the southeast and north coasts (Sæmundsson 1934). In such a situation the size of capelinspawning grounds would be reduced considerably. Should the rise of sea temperature go beyondthat of the 1920-1940 period, capelin spawning might be even further reduced and limited to thenorth and east coasts. This would result in drastic changes of larval drift routes and survival and,in the end, to a large reduction in, or even a complete collapse of, the Icelandic capelin stock.

Page 50: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–50 DO NOT CIRCULATE OR CITE

Because of the key role of capelin as forage fish in the Icelandic marine ecosystem there is littledoubt that such a scenario would have a far reaching negative impact on most of the commercialstocks of fish, whales and seabirds which are dominant in this ecosystem at present. However,within such a scenario one would expect that species from more temperate areas would at leastpartially replace those most affected by the lack of capelin.

To summarize: There is a high degree of uncertainty about the impact of global warming on oceantemperatures around Iceland and Greenland. There is also a considerable uncertainty about theimpact of ocean warming on commercial stocks in the region. Thus, there appears to be asignificant probability that moderately warmer waters will actually lead to a substantial expansionin some of the most valuable fish stocks. On the other hand, it is quite conceivable that excessiveocean warming will lead to large reductions, if not collapses, of ecologically crucial fish stocksand, thus, lead to sudden and quite dramatic reduction in the economic potential of the Icelandicfisheries.

12.3.6 The economic and social importance of fisheries

The fishing industry and economic fluctuations: Lessons from history

IcelandDuring the 20th century, Icelandic gross domestic production (GDP) exhibited an average annualgrowth of about 4% per year. This economic growth was apparently to a large degree generated byexpansion in the fisheries and fish processing industries. Moreover, fluctuations in aggregateeconomic output were highly correlated with variations in the fishing industry. Good catches andhigh export prices incited economic growth, whereas poor catches and adverse foreign marketconditions led to slowdowns and even depressions in the economy. All five major economicdepressions experienced during the 20th century can be directly related to changes in the fortunes ofthe fishing sector, either wholly or partially.

The first of these major depressions covers the period of the First World War, which hadcatastrophic effects on Iceland as on so many other European countries. The first two war yearswere, however, favorable for the fishing sector, as increased demand pushed up foreign prices, butin 1916 the international trade structure broke down and Iceland had to accept harsh terms of tradewith the Allies. In 1917, Iceland was forced to sell half of her trawler fleet to France. Demersal fishand herring catches were consequently seriously reduced in 1917 and 1918. The result was a sharpdip in the GDP and a generally depressed economy until 1920 (Figure 12.3.12).

[Figure 12.3.12 here]

The effects of the “Great Depression” were first felt in Iceland in the autumn of 1930, and in thefollowing two years GDP fell by 0.5% and 5% respectively as demand for maritime exportsdeclined sharply. Following a brief recovery, the economy was hit again when the Spanish CivilWar broke out in 1936 and closed Iceland’s most important market for fish products. Despite theseshocks, economic growth still averaged 3% in the 1930s, mostly because of strong rebound in thefisheries, especially the herring fisheries, in 1933 to 1939. From this it appears that it was primarily

Page 51: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–51 DO NOT CIRCULATE OR CITE

because of the strong performance of the fisheries in the 1930s that the “Great Depression” was lessfelt in Iceland than most other Western Countries.

The Second World War was a boom period for Iceland led by good catches and very favourableexport prices. But in 1947 and subsequent years, herring catches fell considerably and real exportprices subsided from the high wartime levels. The result was a prolonged economic contractionfrom 1949-52.

During the decade 1961-70, the economy exhibited a very respectable growth rate of 4.8% onaverage. This, was to a large extent based on very good herring fisheries during most of the decade.When the herring stocks collapsed toward the end of the decade the result was a severe economicdepression in 1968-69, when the GDP declined by 1.3% and 5.5% respectively. Unemploymentreached over 2% - a great shock for an economy used to excess demand for labor since the 1930s -and many households moved abroad in search of jobs. Net emigration amounted to 0.6% of the totalpopulation in 1969, and 0.8% in 1970.

High economic growth rates resumed in 1971-80 averaging 6.4%. However, just as during the1960s, this growth was to a significant extent based on over-exploitation of the most important fishstocks. Reduced fishing quotas and weak export prices reduced fishing profitability in the late1980s. And, partly as a consequence of this, the Icelandic economy remained stagnant through theyears 1988-1993, with an average annual decline in the GDP of 0.12%.

Since 1993, however, the Icelandic economy has registered steady and quite impressive annualgrowth rates. One reason for this is a recovery of some of the fish stocks. More importantly,however, are generally more favorable fish export prices and the impact of the individualtransferable quota (ITQ) system. The ITQ system has enabled the fishing industry to increase andstabilize profits and much more easily adjust to changing quotas and fish availability.

Thus, looking at the 20th century as a whole, it appears that major fluctuations in the Icelandiceconomy may to a considerable extent be attributed to changes in the fortunes of the fishingindustry both in terms of harvest quantity and output prices. This suggests that possible changes infish stocks due to global warming may have similar macro-economic impact. This, however, wouldprobably be something of an exaggeration. Most likely the macro-economic impact of any givenchange in fish availability will be smaller in the future than it has been in the past. First, theimportance of the fishing industry for the Icelandic economy has declined substantially from itsaverage during the 20th Century. Second, the ITQ fisheries management system has probably madethe fishing industry more capable of adapting effectively to changes in fish stocks than before. Aword of caution, however, is in order. If the current depressed state of some of the most importantfish stocks persists, adverse environmental changes may actually translate into larger biologicalshocks than those experienced in the past.

GreenlandThe Greenland historical experience does not offer the same overwhelming evidence of the nationaleconomic importance of the fishing industry as that in Iceland. This, however, should not be takento mean that the economic importance of the Greenland fishing industry is any less than in Iceland.In fact it is probably much greater.

Page 52: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–52 DO NOT CIRCULATE OR CITE

First, the Greenland fishing industry developed much later than the Icelandic one. Thus during thefirst half of the 20th century, Greenland fishing activity was not very significant (Figure 12.3.19)even when compared to the rest of the Greenland economy. Thus, the data period is simply shorter.Second, being based on fairly underexploited fish stocks, the Greenland fishing industry expandedrelatively smoothly until the 1980s. Thus, there are simply much fewer experiences of dramaticfluctuations in fisheries output in Greenland than in Iceland. Third, the Greenland economicstatistics are far less comprehensive than in Iceland. Thus, there are less data to draw on.

Since 1970 there have been two major cycles in the Greenland economy (Figure 12.2.13) both ofwhich are associated with changes in the fishing industry, more precisely the cod fishery.

[Figure 12.3.13 here]

The cod fishery has historically been Greenland’s most important fishery11. This fishery underwenta major expansion in the latter half of the 1970s as a result of reduction in foreign fishing followingthe extension of the Greenland fisheries jurisdiction to 200 miles and a greatly expanded Greenlandfishing effort. This led to a period of good economic growth that was reversed abruptly in 1981 witha significant contraction in the cod fishery due to a combination of overfishing and low exportprices. The subsequent period of economic depression lasted for 3 years during which the GDPcontracted by 9%. Another short-lived boom in the cod fishery from about 1985 led to acorresponding boom and burst cycle in the economy with a 5 year growth period followed by asharp depression lasting for 4 years during which time the GDP contracted by over 20%. Economicgrowth resumed in Greenland in 1995, not on the basis of cod, which has not reappeared, butshrimp fishing which expanded very rapidly during the latter half of the 1990s.

Thus, just as in Iceland, the available Greenland historical evidence points strongly to a closeconnection between GDP fluctuations and variations in the Greenland fishing industry.

11 This has now been superseded by the shrimp fishery.

Page 53: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–53 DO NOT CIRCULATE OR CITE

The economic and social role of the fisheries

IcelandThe relative importance of the fishing industry in the Icelandic economy seems to have peakedbefore the middle of the twentieth century. Since then, both the share of fish products inmerchandise exports and the fraction of the total labor force engaged in fishing have declinedsignificantly. In the year 2000, the fishing industry employed some 8% of the labor force (12.800employees), accounted for 63% of merchandise exports and generated 42% of export earnings(Figure 12.3.15). Total export value of fish products in the year 2000 was about 1 220 m. USD.

National accounts estimates of the contribution of the fishing industry to the gross domestic product(GDP) - available since 1980 - confirm this trend. Thus, in the year 1980 the direct contribution ofthe fishing industry to the GDP was over 16%. In 2000 this contribution had dropped to just over11% GDP (Figure 12.3.15) which corresponds to a value added of some 900 m. USD.

It is important to realize, however, that these aggregate statistics may not provide much guidance asto the macro-economic impact of alterations in the conditions of the fishing industry. First, theymay understate the real contribution of the fishing industry to the Icelandic economy. There are twofundamental reasons for this. First there are a number of economic activities closely linked with thefishing industry but not part of it. These activities consist of the production of inputs to the fishingindustry, the so-called backward linkages, and the various secondary uses of fish products, the so-called forward linkages (Arnason 1994). The backward linkages include activities such asshipbuilding and maintenance, fishing gear production, the production of fishing industryequipment and machinery, the fish packaging industry, fisheries research and education. Theforward linkages comprise the transport of fish products, the production of animal feed from fishproducts, the marketing of fish products, retailing of fish products, part of the restaurant industryetc.. According to Arnason 1994, these backward and forward linkages may easily add at least aquarter to the direct GDP contribution of the fishing industry.

The other reason why the national accounts may underestimate the true contribution of the fishingindustry to the GDP is the role of the fishing industry as a disproportionately strong exchangeearner. To the extent that the availability of foreign currency constrains economic output, theeconomic contribution of a disproportionately strong export earner may be greater than is apparentfrom the national accounts. While the size of this “multiplier effect” is not easy to measure, somestudies suggest it may be of a significant magnitude (Arnason 1994, Agnarsson and Arnason 2003).If that is true, the total contribution of the fishing industry to the GDP might easily be much higherthan the above direct estimates suggest, in the sense that removal of the fishing industry would,ceteris paribus, lead to this reduction in the GDP.

It is equally important to realize that there are economic reasons why a change in the conditions ofthe fishing industry due e.g. to global warming, might have a lesser economic impact thansuggested by the direct (and indirect) contribution of the fishing industry to GDP. Most economiesexhibit certain resilience to exogenous shocks. This means that the initial impact of such shocks isat least partly counteracted by the movement of labor and capital to economic activities madecomparatively more productive by the shock. Thus, a negative shock in the fishing industry wouldto a certain extent be offset by labor and capital moving from the fishing industry to alternative

Page 54: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–54 DO NOT CIRCULATE OR CITE

industries and vice versa. As a result, the long-term impact of such a shock may be much less thanthe initial impact. The extent to which this type of substitution happens depends on the availabilityof alternative industries. However, with increased labor mobility, communication technology andhuman capital this type of flexibility is probably significantly greater than in the past.

Regional importanceThe above analysis in terms of macro-economic aggregates obscures the fact that the economicimportance of the fishing industry varies greatly from one region of the country to another. Thus, inthe year 2000 when the fishing industry (harvesting and processing) employed only about 8% of thetotal workforce in the country, it provided jobs for over 35% of the working population in theWestern fjords and almost 30% of the working population in the Eastern fjords. Both regions,however, are relatively sparsely populated and account for but a small portion of the total Icelandicpopulation. Thus, in the capital region of Iceland, where most of the alternative industries, such asmanufacturing and services, are located, the fishing industry employed only about 3% of theworking population.

The local importance of the fishing industry becomes even more apparent when individualcommunities are considered. Thus, as listed in Table 12.5.2, in 1997, the fishing industry accountedfor over 40% of the local employment in 24 out of a totality of 124 municipalities in Iceland. Atypical example of a community totally dependent on fishing is Raufarhofn, a small municipality of400 inhabitants in N-East Iceland. In this community almost 70% of the adult population worked inthe fishing industry in 1997. In four other municipalities the fishing industry accounted for over60% of total employment in same year.

Table 12.5.2 The importance of the fishing sectors to communities in 1997

Labor share of thefishing sectors

Number ofcommunities

Number ofinhabitants

Percentage oftotal population

Over 40% 24 12 812 7.725-40% 16 23 063 8.610-25% 14 36 959 13.75-10% 16 26 832 10.0Less than 5% 54 161 922 60.1

By contrast, in 54 municipalities the fishing industry sectors accounted for less than 5% of totalemployment. Most of the largest municipalities in Iceland belong to this group. It is primarily, thesmaller, economically less developed, communities that are heavily dependent on fisheries. This isbrought out clearly in the statistics listed in Table 12.5.2

From this it should be clear that the effects of a significant reduction in fish availability aroundIceland, and, for that matter, the benefits of fisheries expansion would be very differently felt in thevarious regions and communities of Iceland. Broadly speaking, a significant reduction in fishavailability is liable to be economically and socially disastrous for the Western and Eastern fjordsand certain other smaller regions of Iceland, while in the more densely populated southwest such a

Page 55: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–55 DO NOT CIRCULATE OR CITE

reduction would primarily be felt as an increased influx of labor from the outlying regions and thecorresponding realignment of economic activity.

Although labor mobility is comparatively high in Iceland, it may not prove easy for some of theinhabitants of fishing villages to find jobs elsewhere in the case of fisheries contraction due toglobal warming, especially if the Icelandic economy is already depressed. Moreover, since much ofthe employment opportunities in the capital area require particular education and training,individuals transferring from the fishing industry may have to accept relatively inferior jobs, if theycan find them. At the same time, contraction of employment and emigration of people from thefisheries dependent regions and communities of Iceland will inevitably depress real estate prices inthese areas. Thus, these people may have to endure significant loss of their asset values at the sametime they have to look for new jobs.

Thus, obviously, a significant reduction in the Icelandic fishing industry would lead to noticeablesocial disruption. This disruption would be temporary, however. Given the nature of Icelandicsociety, it would probably work itself out in 5-10 years following the initial shock. During thisperiod, however, the disruption would impose a certain stress on the social and political system.

GreenlandThe fishing industry is by far Greenland’s most important production section. In the 1960s and early1970s the export of fish and fish products were between 80 and 90% of the total export fromGreenland. In 1974 there was a very large increase in the export of lead and zinc, which resulted inan increased GDP by about 50% and in the following years the share of fish and fish products intotal export value dropped to 60 - 70%. The export of lead and zinc ceased in 1990. Since then, theexport of fish and fish products has been about 90% of the total Greenland export. In year 2000, thevalue of the export of fish and fish products was about 270 million USD while the total export valuewas about 285 million USD.

Exact statistics about the direct contribution of the fishing industry to the Greenland GDP are notavailable. However, the contribution to the GNI (Gross national income) may be as high as 20%.This, however, does not tell the complete story. Greenland is part of Denmark with a “Home Rule”government. This means that Greenlanders can decide their own policies, except for foreign anddefense policy. Every year, the Home Rule government of Greenland receives economic supportfrom the Danish State. In year 2000 this supported amounted to about 350 million USD or almost25% of the GNI. Correcting for this we may have a direct contribution of the fishing industry to theGreenland GDP between 25 and 30%.

As in the case of Iceland, however, the fishing industry also has indirect contribution to theGreenland economy via forward and backward linkages as well as multiplier effects. Adding thesemay easily bring the total contribution of the fishing industry to the Greenland economy as a wholeto above 50%.

Regional importanceAs has been made clear by the above, Greenland as a whole is highly dependent upon the fishingindustry. This applies even more so to the less populated communities along the coast. About 20%

Page 56: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–56 DO NOT CIRCULATE OR CITE

of Greenland’s population lives in small villages and settlements with average population of about150 inhabitants. Many more live in small towns with less than a thousand inhabitants. In thesecommunities the economic activity consists almost exclusively of the exploitation of living marineresources, i.e., fishing and hunting. Moreover, due to the geographical isolation of many of thesecommunities, alternative employment opportunities are few if any.

Thus, it is clear that a significant drop in the fish stocks and other living marine resource wouldhave a devastating impact on these communities. Most would have to contract significantly andmany would disappear completely rendering the previous inhabitants to a substantial extenteconomically and socially dispossessed. A secondary impact would be in the form of a substantialinflux of the erstwhile inhabitants of these communities to the more urban areas of Greenlandcausing additional problems in these areas on top of those stemming directly from a contraction inthe fisheries.

The impact of a significant increase in the stocks of fish and other living marine animals would bythe reverse of this. It would strengthen the economic basis of Greenland’s smaller communities.While larger towns may be expected to benefit disproportionately from such as change, the neteffect would probably be to increase population in the smaller communities and to expand thegeographical extent of habitation in Greenland.

12.3.7 Economic and social impacts of global warming: Possible scenarios

From an economic point of view, climate change may impact fisheries in at least two differentways. First, by altering the availability of fish to the fishermen. Second, by changing the prices offish products and fisheries inputs. Although both types of impacts may be initiated by climatechange, the first is a much more direct consequence of climate change than the second.

The possible impact of climate change on fish availability may occur through changes in the size ofcommercial fish stocks, their geographical distribution and their catchability. These changes, if theyoccur, will affect the availability of fish for commercial harvesting. The direction of this impact isuncertain. It may be negative; that is to say it may reduce the maximum sustainable economic yieldfrom the fish stocks. But it may also be positive, i.e. increase the maximum sustainable economicyield from the fish stocks. Moreover, in spite of general global warming, the impact on the variousfish stocks may be very different and vary across regions. Irrespective of the direction of the impact,however, it seems highly likely that climatic change will, at least temporary, cause instability orfluctuations in harvesting possibilities as the ecosystem adjusts to new conditions. This adjustmentprocess may take a long time, even after the period of climatic change has ended.

Very much the same applies to the changes in relative prices that may continue their adjustmentprocess long after an exogenous shift like climate change has worked itself out. In fact the economicadjustments following a climatic change, being dependent on the biological/ecological adjustmentswill continue for some time after the latter are completed.

This section is restricted to speculating on the possible economic and social impacts in Iceland andGreenland of changes in fish availability. Thus, in this section we will not discuss the possible

Page 57: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–57 DO NOT CIRCULATE OR CITE

impacts of relative price changes. Note, however, that the economic and social impacts of pricechanges will be quite similar to those of changes in fish stock availability. Thus, as far as drawinginferences from the historical evidence, it doesn’t really matter whether expansions and contractionsin the fishing industry are caused by changes in prices or the availability of fish.

The available empirical evidence of the possible economic impacts of changes in fish stockavailability is of two types; (i) qualitative historical evidence and (ii) quantitative evidence. Thequalitative evidence, discussed in section 12.3.5 above, relates economic fluctuations to more orless qualitative evidence of expansions and contractions in the fishing industry. The quantitativeevidence, in the form of time series for fisheries production and production values, provides a basisfor the statistical estimation of the relationship between the production value of the Icelandic andGreenland fishing industries and their respective GDP and GNP growth.

Iceland

Quantitative evidenceReasonably reliable time series data for the output and output value of the fishing industry areavailable since 1963. These data have been used to estimate the form and parameters of arelationship between economic growth rates and the output value of the fishing industry as well asother relevant economic variables such as capital and labour (Agnarsson and Arnason 2003). Theestimated equation exhibits good statistical properties. Actual and fitted GDP growth ratesaccording to the estimated equation are illustrated in Figure 12.3.15.

[Figure 12.3.15 here]

This estimated equation can, with certain modifications, be used to predict the short and long termimpact of a change in fish stock availability due to global warming. It is important to realize,however, that to use this equation we need to predict (i) the extent and timing of global warming,(ii) the impact of global warming on fish stock availability and (iii) the impact of changed fish stockavailability on the value of fish production. Needless to say, the last item involves both the volumeand price of fish production.

Impact on GDP: ScenariosThese calculations are restricted to the impact of global warming on the availability of fish. Thismeans that the impact of other variables on the value of fish production is ignored. In this sectionwe will calculate estimates of the possible GDP impact of changed fish stock availability as a resultof climate change. These calculations are based on two key premises; (i) the impact of futureclimate change on the value of fish production in Iceland and (ii) the estimated relationship betweeneconomic growth and the value of fish production as discussed above (see Agnarsson and Arnason2003). Both of these premises, not least the impact of climate change on the value of fishproduction, are highly uncertain. Therefore, the following calculations should not be regarded aspredictions. They are merely intended to serve as indications of the likely magnitudes of the GDPimpact in Iceland stemming from certain stated premises regarding changes in fish stockavailability.

Page 58: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–58 DO NOT CIRCULATE OR CITE

Now, the available predictions [Chapter 8 in this the volume] suggest that the global warmingduring the next 50-100 years is first of all not going to have a great impact on fish stock availabilityin Icelandic waters. Second, it is more likely to benefit the most valuable fish stocks rather than thereverse. As a result, the overall effects of global warming on the Icelandic fisheries are more likelythan not to be positive. These expectations, however, are highly uncertain as already mentioned.Therefore, in this section, we will proceed on the basis of three examples or scenarios.

In the first scenario we will assume a gradual increase in fish stock availability of 20% over aperiod of 50 years. This corresponds to a 0.4% increase in the value of fish production annually. Werefer to this as the ‘optimistic scenario’. In the second scenario we will assume a gradual reductionin fish stock availability of 10% over 50 years. This corresponds to an annual reduction in the valueof fish production by some 0.2%. We refer to this as the ‘pessimistic scenario’. In the third scenariowe assume a 25% reduction in fish stock availability occurring over a relatively short period of 5years. This would correspond to a collapse in the stock size of one major or a group of importantcommercial species. Indeed, there are some biological and ecological grounds to believe that theresponse of fish stocks to climatic change may indeed be sudden and discontinuous rather thangradual. Due to the magnitude and suddenness of this reduction we refer to it as the ‘dramatic case’.

Hopefully these scenarios will serve to illustrate the likely range of the economic impacts of globalwarming around Iceland. In interpreting the outcomes of these scenarios it is important to keep inmind that they are restricted to the impact of global warming assuming all other variables affectingfish stocks and their economic contribution remain unchanged. In particular, these outcomes do notincorporate the possibly simultaneous impact of improved fisheries management or other variablesaffecting the size of the fish stocks and the value of the fisheries. Indeed, given the currentlydepressed state of many of the most valuable fish stocks in Iceland, it seems that a more sensibleharvesting policy may easily contribute at least as much to the overall economic yield of thefisheries as the most optimistic climate scenario discussed below. Needless to say, such a policywill similarly improve the outcomes from the more pessimistic climate scenarios.

The optimistic scenarioIn the optimistic case, fish stock availability is assumed to increase in equal steps by 20% over thenext 50 years. The time profile of the impact of this on GDP relative to a benchmark GDP of unityis illustrated in Figure 12.3.15.

[Figure 12.3.21 here]

Perhaps the most noteworthy result illustrated in Figure 12.3.15 is that this, quite considerableincrease in fish stock availability, only has a relatively minor impact on GDP. The maximum impactoccurs in year 50, when increased fish stock availability has fully materialized. At this point theGDP has increased (compared to the initial level) by less than 4%. The long term impact, wheneconomic adjustment processes have worked themselves out, is even less or some 2.5%.

As indicated in Figure 12.5.15, the largest annual increase in GDP is quite small. Closerexamination shows that it is actually well under 0.2%, which, incidentally, is substantially less thanthe GDP measurement error. Hence it would hardly be noticeable. In the years following the end ofthe increase in fish production, growth rates actually decline as production factors, moving to the

Page 59: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–59 DO NOT CIRCULATE OR CITE

fishing industry, reduce economic production elsewhere. Long run GDP growth rates are, of course,unchanged.

The fundamental conclusions to be drawn from these calculations seem to be that a 20% increase inthe output of the fishing industry equally spread over 50 years has a very small, hardly noticeable,impact on the short run economic growth rates in Iceland as well as the long term GDP.

The pessimistic scenarioIn the pessimistic case, fish stock availability is assumed to decrease in equal steps by 10% over thenext 50 years. The time profile of the impact of this on GDP relative to a benchmark GDP of unityis also illustrated in Figure 12.3.15.

As in the optimistic case, the most striking result under this scenario is that this quite considerabledecrease in fish stock availability has a relatively minor impact on long term GDP. The maximumimpact occurs in year 50, when the GDP has been reduced by less than 2%. The long-term impact,when economic adjustment processes have worked themselves out, is even less or just over 1%.

Similar story applies to annual GDP growth rates. The largest annual decrease in GDP is well above-0.1%. This occurs a few years after the decrease in fish stock availability begins. For most of theperiod, however, the impact on annual economic growth rates is much less. In the years followingthe end of the decrease in fish production, growth rates actually improve, as production factors,moving from the fishing industry, find productive employment elsewhere. It should be noted that allthese deviations in annual GDP growth rates are well within GDP measurement errors. Long runGDP growth rates are, of course, unchanged.

As in the optimistic case, the fundamental conclusions to be drawn from these calculations seem tobe that a 10% decrease in the output of the fishing industry equally spread over 50 years hardly hasa noticeable impact on the short run economic growth rates in Iceland as well as the long term GDP.

The dramatic scenarioThe dramatic scenario assumes a fairly substantial drop in fish stock availability and, hence, fishproduction of 25% over the next 5 years. The resulting time profile of the impact of this on GDPrelative to a benchmark GDP of unity is illustrated in the uppermost curve in Figure 12.3.15.

As illustrated in Figure 12.3.15, this sudden drop in fish stock production has a significant negativeimpact on GDP in the short run. At its lowest point, in year 8, GDP is reduced by over 9%compared to its initial level. The long term negative impact, however, when economic adjustmentprocesses have worked themselves out, is only about 3% reduction in GDP.

The decrease in annual GDP growth rates for the first seven years following the commencement ofreduction in fish stock production are quite significant or -1 to -2%. The maximum decline occurstoward the end of the reduction process in years four and five. However, already, in year nine, fouryears after, the contraction ends, the deviation in annual GDP growth rates is reversed as productionfactors released from the fishing industry find productive employment elsewhere in the economy.

Page 60: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–60 DO NOT CIRCULATE OR CITE

This adjustment process is fully worked out in year 25, according to these calculations, whengrowth rates have fully reverted to the basic underlying economic growth level

Social and political impactsIf the change in fishing industry output is gradual and the economic impact comparatively small asdescribed in the “optimistic” and “pessimistic” scenarios above, it is unlikely that the accompanyingpolitical and social changes will prove to be noticeable on a national scale. However, in the longrun, as the change accumulates, certain changes will undoubtedly take place. Whether these will belarge enough to be distinguishable from the impact of other changes is another matter, especiallywhen viewed on a national scale.

Regionally, however, the situation may be quite different. In various parts of Iceland, as alreadydiscussed, the economic and social role of the fishing industry is far above the national average. Inthese parts, the economic, social and political impact of an expansion or contraction in the fishingindustry will be much greater than for the nation as a whole and in some cases undoubtedly quitedramatic.

If, on the other hand, the change is fairly sudden, as in the “dramatic” scenario above, the short termsocial and political impact may be quite drastic. In the long run, however, as the initial impactpeters out, social and political conditions revert to the long term scenario described by the“pessimistic” scenario. Whether the economic and social adjustments will then also revert to theirinitial state is not clear.

Impacts on fish marketsReduction or increases in fish production in Iceland alone will not have a significant impact onglobal fish markets. Neither is it likely to have a large impact on the marketing of Icelandic fishproducts, provided the changes are gradual. If there is an overall decline in the global supply ofspecies of fish that Iceland currently exploits, the impact on marketing of these species is uncertain.Most likely the marketing of the species in reduced supply will become easier. Thus, prices willtend to rise counteracting the contraction in volume. However, it should be recognized that themarketing impact might actually be the opposite. For some species, a large and steady supply isrequired to maintain marketing channels. If this is threatened, these channels may close andalternative outlets have to be found.

DiscussionThe main conclusion to be drawn from the above is that the changes in fish stock availability thatnow seem most likely to be induced by global warming over the next 50-100 years are unlikely tohave a significant long term impact on the Icelandic GDP and, consequently, social and politicalconditions. Also, it appears that any impact, small as it may be, is more likely to be positive thannegative.

If, on the other hand, global warming triggers sudden rather than gradual changes in fish stockavailability, the short time impact on the Icelandic GDP and economic growth rates may be quitesignificant. The impact seems very unlikely to be dramatic (say, over 5% change in GDP betweenyears), however. In the long run the GDP impact of a sudden change in fish stocks will be

Page 61: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–61 DO NOT CIRCULATE OR CITE

indistinguishable from a more gradual change. The long terms social and political impacts maydiffer, however, although there is no particular reason to expect that.

Greenland

Quantitative evidenceReasonably reliable time series data for the export value of the Greenland fishing industry areavailable since 1966. These data have been used to estimate the form and parameters of arelationship between GDP and the real export value of fish products (Vestergaard and Arnason2003). The estimated equation exhibits reasonable statistical properties. Actual and fitted GDPgrowth rates according to the estimated equation are illustrated in Figure 12.3.16.

[Figure 12.3.16 here]

According to this estimated equation, a 1% increase in the export value of fish products will in duecourse lead to some 0.29% increase in the Greenland GDP. Subject to the same qualifications as inthe case of Iceland above, this estimated equation can be used to predict the economic impact of achange in fish stock availability due to global warming.

Impact on GDP: ScenariosThe available predictions [Chapter 8 and sections 12.3.3 and 12.3.4 in this volume] suggest that theglobal warming during the 21st Century is much more likely to benefit the most valuable fish stocksat Greenland rather than the reverse. This holds in particular for the cod stock, which couldexperience a revival from its current extremely depressed state to a level, seen during warmerclimate years in the 20th Century, where it could yield up to 300 000 t on a sustainable basis. On thenegative side, global warming and increased predation by cod could lead to a dramatic fall in thesustainable harvest of shrimp by some 50-70 000 t (see Figure 12.3.17). The value of the increasedcod harvest would, however, greatly exceed losses due to a possibly reduced harvest of shrimp. Infact, this change could easily lead to doubling or even tripling of the total production value of theGreenland fishing industry. Thus, the predicted global warming could actually have a major positiveimpact on the Greenlandic fishing industry. As before, however, we emphasize that this possibilityis highly uncertain. Therefore, as in the case of Iceland, we will proceed on the basis of threeexamples or scenarios.

[Figure 12.3.17 here]

In the first scenario which may be referred to as the pessimistic scenario, we will assume that inspite of more favorable habitat conditions, cod will not manage to re-establish itself permanently inGreenland waters. However, there will be periodic bursts of cod availability accompanied by acorresponding drop in shrimp, based on an occasional large scale larval drift from Iceland similar tothat seen in warmer periods in the past. The overall impact of this will be a slight increase in fishharvests on average with some peaks and troughs. In the second scenario we will assume a modestand gradual return of cod in Greenland which in 20 years would be capable of yielding some 100thousand t. per annum on average. As before there will be corresponding decline in the shrimpstock. Given the comparatively high likelihood of a substantial increase in the cod stock offGreenland, we refer to this as the moderate scenario. In the third scenario we assume a return of the

Page 62: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–62 DO NOT CIRCULATE OR CITE

Greenland cod stock, initially generated by Icelandic cod larval drift, to its historical levels in the1950s and 60s. The full revival, however, is assumed to take some decades and to occur in afluctuating manner. Ultimately in about 30 years time the cod stock is assumed to be capable ofproducing an average yield of some 300 thousand t annually compared to virtually nothing now.The average shrimp harvest, however, would be reduced from a current level of almost 100thousand t per annum to about 20 thousand t per annum. Nevertheless, the overall value of theGreenland fish harvest would almost double. We refer to this scenario as the optimistic scenario.

Hopefully these scenarios will serve to illustrate the likely range of the economic impacts of globalwarming around Greenland. In all three cases, the harvest predictions are based on a two speciesfisheries model developed for the Greenland fisheries (Hvinkel, 2003).

The pessimistic scenarioThe pessimistic scenario assumes an insubstantial change in overall average fish stock availability.However, due to the occasional large scale influx and survival of Icelandic cod larvae periodic burstof cod availability occurs. This gives rise to a fluctuating fish production and GDP impacts overtime. An example of one possible time profile of GDP relative to a benchmark GDP of unity isillustrated in the lowest curve in Figure 12.3.18.

[Figure 12.3.18 here]

As illustrated in Figure 12.3.18, this scenario leads to a fluctuating impact on GDP with a smallaverage increase. The average increase in the GDP level after some 50 years is about 2% comparedto what would otherwise have been the case.

The moderate scenarioIn the moderate case, fish stock availability is assumed to increase gradually by about 20% over thenext century. The time profile of the impact of this on GDP relative to a benchmark GDP of unity isillustrated in the middle curve in Figure 12.3.18.

As illustrated in Figure 12.3.24 this increase in the availability of fish leads to a moderate long termincrease in GDP of some 6% compared to what would otherwise be the case. However, since mostof this increase is predicted to occur during the first decade - in fact initial impact is predicted to begreater than the long term impact - there would be a significant addition to the GDP of some 1%annually during this initial period.

The optimistic scenarioIn the optimistic case, fish stock availability is assumed to increase gradually but in a fluctuatingway by about 100% over the next 50 years. According to the estimated relationship between thevalue of fish exports and GDP, this will lead to ultimate increase in the Greenland GDP of some28% compared the what would otherwise be the case. The time profile of the impact of this on GDPrelative to a benchmark GDP of unity is illustrated in Figure 12.3.18.

This impact spread would certainly be very noticeable. The annual addition to economic growthwould be close to 0.8% annually during the first 30 years but would dwindle thereafter and be fullyworked out after about 40 years.

Page 63: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–63 DO NOT CIRCULATE OR CITE

Social and political impactsIf the optimistic case occurs and the increased cod harvest is mainly caught and processed byGreenlanders there will be a dramatic improvement in the Greenlandunemployment/underemployment situation and in the income of the large group of self-employedsmall-boat fishermen/hunters. Everything else being the same, the current high level ofunemployment would more or less vanish and the total income of many families would increasemarkedly. Besides the positive individual and national economic effects this would inevitably alsohave a major social impact in Greenland. One interesting aspect is that cod fishing of the magnitudepredicted in the optimistic scenario might easily lead to the establishment of large scale fishprocessing factories in the more densely populated regions of Greenland.

In the other cases considered, the employment and consequently social impact would be much moremoderate, especially if whatever changes take place appears gradually over a longish period of time.If, on the other hand, the change turns out to be fairly sudden, as in the “dramatic” scenariodiscussed for Iceland above, the short term social and political impact may be quite drastic.

DiscussionThe likely impact of the possible changes in fish stock availability in Greenland waters stemmingfrom global warming cover a much wider range than in Iceland. At one extreme, it could lead up to30% increase in the GDP. At the other extreme the impact on GDP could be negligible. This widerange reflects on one hand the greater importance of the fisheries sector in Greenland compared toIceland. On the other hand, it reflects on the fact that during warm periods the very large marinehabitat around Greenland can accommodate a biomass of commercial species, which is many timesgreater than that of the present time. Since the Arctic influence is much more pronounced in theGreenland marine environment than in that around Iceland, it follows naturally that an impact by amoderate global warming, if it happens, should be expected to be much more dramatic in the seasaround Greenland.

12.3.8 The ability to cope with sudden changes

Global warming will almost certainly lead to changes in the relative sizes, biological productivityand geographical distribution of commercial fish stocks. These changes may be predominantlyadvantageous or they may be disadvantageous. They may be sudden or they may emerge gradually.

The economic and social impact of changes in fish stock availability obviously depends on thedirection, magnitude and rapidity of these changes. It is important to realize, however, that theeconomic and social impact also depend, possibly even more so, on the ability of the relevant socialstructures to adapt to altered conditions. Good social structures facilitate speedy adjustments to newconditions and thus mitigate any negative impacts. Weak or inappropriate social structures willexhibit sluggish and possibly inappropriate responses and thus quite possibly exacerbate theproblems caused by adverse environmental changes.

One of the most crucial social structures in this respect is probably the fisheries managementsystem. This will basically determine the extent to which the fisheries will be able to adapt in an

Page 64: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–64 DO NOT CIRCULATE OR CITE

optimal fashion to new conditions. Other important social structures relate to (i) the adaptability ofthe economic system - especially price flexibility, labour education and mobility and the extent ofeconomic entrepreneurship - (ii) the ability to adjust macro-economic policies, (iii) and the extentand nature of the social welfare system. These structures will influence to what extent the necessaryadjustments to new conditions will be smooth and quick or perhaps difficult and drawn out in time.

The Icelandic fisheries management system being based on permanent harvest shares in the form ofITQs is inherently forward looking and thus probably well suited to adjust optimally to changes inthe availability of fish, not the least if these changes are to some extent foreseen. Given any timepath of fish stock productivity, ITQ-holders have a strong incentive to maximize the expectedpresent value of the fishery since this will also maximize their expected wealth. As a result, there isa high likelihood that TACs and other stock size determinants will be adjusted optimally to alteredconditions. In other words, given the Icelandic fisheries management system, there is little chancethat the fishing activity will exacerbate a negative biological impact stemming from globalwarming. By the same token, the opportunities generated by a positive biological impact willprobably be close to fully exploited by the fishing sector. In Greenland, on the other hand,fishermen’s rights to harvest shares are considerably weaker than in Iceland. As a result, Greenlandseems to be somewhat less prepared than Iceland to adapt in an economically efficient manner tochanged fish stock availability than Iceland.

Turning to the other relevant social structures, it appears that in Iceland these are rather well suitedto facilitate adjustment to adverse environmental changes. First, it should be observed that Icelandicsocial structures are used to having to adapt to quite drastic fluctuations in the economy (see chapter12.3.6 above). This, presumably, means that these institutions have evolved to cope with suchfluctuations. Second labor education and mobility is quite high in Iceland. High labour mobilityrefers both to labor movements within the country as well as between Iceland and other countries.This means that negative regional impacts caused by global warming are unlikely to lead topermanently depressed unemployment areas or even persistent national unemployment. Third, thelevel of private and commercial entrepreneurship seems to be comparatively high in Iceland. Thissuggests that in the case of negative impacts, economic substitute activities are rather likely to bequickly spotted and exploited. Fourth, there is an extensive social welfare system in Iceland thatmay be expected to provide at least temporary compensation to individuals adversely affected byenvironmental change. Acknowledging that the existence of this social safety net will tend to delayadjustments to new conditions, it will at least share the burden of adverse changes amongst the totalpopulation and thus prevent individual hardships.

Thus, broadly speaking, it appears that Icelandic social structures are rather well suited to cope withsudden changes in fish stock availability caused by global warming. This means that adjustments tochanges are likely to be reasonably fast and smooth. Adverse environmental changes willnevertheless be economically and socially costly. However, they are unlikely to be exacerbated bysocial and economic responses. On the other hand, it should be noted that although the Icelandiceconomy seems well prepared to deal with changes in fish stock availability, a possible adversebiological impact may be more drastically felt than those experienced in the past. The reason is thatsome of the most valuable commercial stocks are currently close to their historical minimum.Hence, if there is an adverse environmental change, the initial reduction in harvests may have to be

Page 65: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–65 DO NOT CIRCULATE OR CITE

more dramatic than previously experienced and, at the same time, the risk of a long tem stockdepression greater.

As regards the ability of social structures to adjust to new conditions, the Greenland situation is inmany respects similar to that of Iceland. Greenlanders are used to variable environmental andeconomic conditions. Although perhaps not at the level of Iceland, general education and labortraining is also comparatively high in Greenland. However, partly due to the great size of Greenlandand the isolation of many communities, labor mobility in Greenland is considerably less than inIceland. Most importantly, however, the degree of private and commercial entrepreneurship inGreenland seems considerably inferior to that of Iceland.

Thus, while the Greenland social structure and institutions seem reasonably well placed to adjust tothe changes in fish stock availability that might be caused by global warming, they appear as yet tobe somewhat less suited in this respect than those in Iceland.

12.3.9 References

Agnarsson, S. and Arnason, R. 2003. The Role of the Fishing Industry in the Icelandic Economy.An Historical Examination. Institute of Economic Studies. Univeristy of Iceland.(Forthcoming).

Anon. 2000. Udvikling i nogle centrale økonomiske konjunkturindikatorer. Konjunkturstatistik2000:2. Statistics Greenland.

Anon. 2001. Environmental Conditions in Icelandic Waters 2000. In Icelandic –English table andfigure headings. Technical Report 83, Marine Research Institute, Reykjavík. 41 p.

Anon. 2002. State of Marine Stocks in Icelandic Waters 2001/2001. Prospect for the Quota Year2002/2003. Technical Report 88, Marine Research Institute, Reykjavík. 196 p.

Anon. 2002a. Workshop "Ecosystem West Greenland" Greenland Institute of Natural Resources,Nuuk 29 November - 03 December 2001. (Ed. Astrid Jarre). Inussuk, Arctic Research Journal1, 2002.

Arnason, R. 1994. The Icelandic fisheries: Evolution and management of a fishing industry.Fishing News Books. Oxford.

Astthorsson, O. S., and A. Gislason 1995. Long-term changes in zooplankton biomass in Icelandicwaters in spring. ICES Journal of Marine Science, 52: 657-668.

Astthorsson, O. S., and A. Gislason 1998. Short communication: environmental conditions,zooplankton and capelin in the waters north of Iceland. ICES Journal of Marine Science, 55:808–810.

Astthorsson, O. S., and H. Vilhjálmsson 2002. Iceland Shelf LME: Decadal Assessment andResource Sustainability. In (Sherman, K. and H. R. Skjoldal Eds.) Large Marine Ecosystemsof the North Atlantic. Elsevier Science B.V.: 219-243.

Astthorsson, O. S., I. Hallgrímsson and G. Jónsson 1983. Variations in zooplankton densities inIcelandic waters in spring during the years 1961 - 1982. Rit Fiskideildar 7, 2: 73-113.

Bogason, V. 1997. Fæda landsels (Phoca vitulina) vid Ísland. (The food of harbour seal (Phocavitulina) at Iceland). In Fjölstofnarannsóknir (Multispecies Research) 1992–1995, pp.

Page 66: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–66 DO NOT CIRCULATE OR CITE

319–330. Ed. by J. Jakobsson and Ó. K. Pálsson. Hafrannsóknastofnunin (Marine ResearchInstitute), Reykjavík, Technical Report, 57 (In Icelandic).

Buch, E. and H. H. Hansen 1988. Climate and cod fishery at West Greenland. In Wyatt T. and M.G. Larraneta (Eds). Long term changes in marine fish populations. A symposium held inVigo, Spain, 18-21 Nov. 1986, pp. 345-364.

Buch, E., M. H. Nielsen and S. a. Pedersen 2002. Ecosystem variability and regime shift in WestGreenland waters. NAFO, SCR Doc. 02/16, Ser. No. N4617.

Buch, E., S.A. Horsted and H. Hovgaard 1994. Fluctuations in the iccurrence of cod in Greenlandwaters and their possible causes ICES Mar.Sci. Symp. 198: 158-174.

Buch., E. 1993. The North Atlantic water component of the West Greenland Current. ICES CM1993/C: 20.

Dragesund, O., J. Hamre and Ø. Ulltang 1980. Biology and population dynamics of the Norwegianspring spawning herring. Rapp. P.-v. Reun. Cons. Int. Explor. Mer 177: 43-71.

Einarsson, H. A. 1997. Fæda ysu (Melanogrammus aeglefinus) vid Ísland (The food of haddock(Melanogrammus aeglefinus) at Iceland.) In Fjölstofnarannsóknir (Multispecies Research)1992–1995, pp. 69–77. Ed. by J. Jakobsson and Ó. K. Pálsson. Ed. by J. Jakobsson and Ó. K.Pálsson. Hafrannsóknastofnunin (Marine Research Institute), Reykjavík, Technical Report, 57(In Icelandic).

Eiríksson, H. 1999. Spatial variabilities of CPUE and mean size as possible criteria for unit stockdemarkations in analytical assessments of Nephrops at Iceland. Rit Fiskideildar 16, 1: 239-245.

Fridriksson, Á. 1948. Boreo-tended changes in the marine vertebrate fauna of Iceland during thelast 25 years. Rapp. P.-V. Reun. Cons. int. Explor. Mer 125: 30-32.

Friis-Rødel, E., and P. Kanneworff 2002. A review of capelin (Mallotus villosus) in Greenlandwaters. ICES Mar.Sci. Symp. 216: 890-896.

Gudmundsdottir, A. and H. Vilhjalmsson 2002. Predicing totsl allowable catches for Icelandiccapelin, 1978-2001. ICES Mar.Sci. Symp. 216: 1105-1115.

Gunnlaugsson, Th. and Víkingsson, G. 2001. Trends in humpback whale (Megapteranovaeangeliae) sightings rates from aerial surveys in Icelandic waters during 1986-2001.NAMMCO, SC/9/18 2001: 6 pp

Hansen J., D. Johnson, A. Lacis, S. Lebedeff, P. Lee, D. Rind and G. Russell 1983. ClimaticEffects of Atmospheric Carbon Dioxide. Science 220: 874-875.

Hansen, P. M. 1949. Studies of the biology of the cod in Greenland waters. Rapp. P.-V. Reun.Cons. int. Explor. Mer 123: 1-83.

Hansen, P. M., AD. S. Jensen and Å. V. Tåning 1935. Cod marking experiments in the waters ofGreenland 1924-1933. Meddelelser fra Kommissionen for Danmarks Fiskeri- ogHavundersøgelser. Serie: Fiskeri, 10, 1, p 3-119.

Hvingel, C. 2003. A Predictive Fisheries Model for Greenland. Discussion paper.Jakobsson 1980. Exploitation of the Icelandic spring- and summer spawning herring in relation to

fisheries management, 1947-1977. Rapp. P.-v. Reun. Cons. Int. Explor. Mer 177: 23-42.Jakobsson, J. 1964. Recent developments in the Icelandic purse-seine herring fishery. In Modern

Fishing Gear of the World 2 (H. Kristjonsson Ed.). Fishing News (books): 294-395.Jakobsson, J. 1969: On herring migration in relation to changes in sea temperature. Jökull, 19:

134-145.

Page 67: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–67 DO NOT CIRCULATE OR CITE

Jakobsson, J. 1978. The north Icelandic herring fishery and environmental conditions 1960-1968.ICES Symposium on the Biological Basis of Pelagic Fish Stock Management. Paper no. 30,101 pp.

Jakobsson, J. 1992. Recent variability in the fisheries of the North Atlantic. ICES Mar. Sci. Symp.,195: 291-315.

Jakobsson 2002. Ocean Travellers. In 100 Years of Science under ICES, E. Anderson (Ed.). ICESMar.Sci. Symp. 215.

Jakobsson, J. and Østvedt, O. J. 1999. A review of joint investigations of the distribution ofherring in the Norwegian and Iceland Seas 1950-1970. In Unnsteinn Stefánsson Festschrift.Vilhjalmsson, H., Ólafsson, J. and Astthorsson, O. S. (Eds.). Rit Fiskideildar XVI, pp. 209-238.

Jakobsson, J. and Stefánsson, G. 1998. Rational harvesting of the cod-capelin-shrimp complex inthe Icelandic marine ecosystem. Fisheries Research 37: 7-21.

Jakobsson, J. and Stefánsson, G. 1999. Management of Icelandic summer spawning herring offIceland. ICES Journal of Marine Science. 56: 827-833.

Jensen, A. S. 1926. Indberetning av S/S Dana’s praktisk videndskabelige fiskeriundersøgelserved Vestgrønland i 1925. Beretninger og Kundgørelser vedr. Styrelsen af Grønland, 2: 291-315.

Jensen, A. S. 1939. Concerning a change of climate during recent decades in the Arctic andSubarctic regions, from Greenland in the west to Eurasia in the east, and contemporarybiological and physical changes. Det. Kgl. Danske Videnskabernes Selskab. Biologiske Medd.14, 8: 1-77.

Jensen, Ad. S. 1939. Concerning a change of climate during recent decades in the Arctic andSubarctic regions, from Greenland in the west to Eurasia in the east, and contemporarybiological and physical changes. Det. Kgl. Danske Videnskabernes Selskab. BiologiskeMedd. 14, 8: 1-77.

Jónsson, J. 1994. Fisheries off Iceland, 1600-1900. ICES Mar. Sci. Symp. 198: 3-16.Jónsson, J. 1996. Tagging of cod (Gadus morhua) in Icelandic waters 1948-1986. Rit Fiskideildar

14. 1: 5-82.Jónsson, J. 1996. Tagging of cod (Gadus morhua) in Icelandic waters 1948-1986. Rit Fiskideildar

14, 1: 5-82Jónsson, S. Th. 1997. Fæda ufsa (Pollachius virens) vid Ísland. (The food of saithe (Pollachius

virens) at Iceland.) In Fjölstofnarannsóknir (Multispecies Research) 1992–1995, pp. 45–56.Ed. by J. Jakobsson and Ó. K. Pálsson. Hafrannsóknastofnunin (Marine Research Institute),Reykjavík, Technical Report, 57 (In Icelandic).

Lilliendahl, K., and Sólmundsson, J. 1997. An estimate of summer food consumption of sixseabird species in Iceland. ICES Journal of Marine Science, 54: 624–639.

Malmberg, S. A.. 1984. A note of the seventies anomaly in Icelandic waters. ICES CM 1984/Gen.Mini Symp. 8, p. 12.

Malmberg S. A., Ecological impact of hydrographic conditions in Icelandic waters. In Wyatt T.and M. G. Larraneta (Eds). Long term changes in marine fish populations. A symposium heldin Vigo, Spain, 18-21 November 1986: 95-123.

Malmberg, S. A. and S. Kristmannsson 1992. Hydrographic conditions in Icelandic waters, 1980-1989. ICES Mar. Sci. Symp. 195: 76-92.

Page 68: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–68 DO NOT CIRCULATE OR CITE

Malmberg, S. A., J. Mortensen and H. Valdemarsson 1999. Decadal scale climate andhydrobiological variability in Icelandic waters in relation to large-scale atmospheric conditionsin the North Atlantic. ICES CM 1999/L: 13. 9p.

Marteinsdottir, G. and A. Steinarsson. 1998. Maternal influence on the size and viability of Icelandcod (Gadus morhua L.) eggs and larvae. J. Fish. Biol. 52 (6) : 1241-1258.

Marteinsdottir, G. and G. Begg. 2002. Essential relatinships incorporating the influence of age,size and condition on variables required for estimation of reproductive potential in Atlanticcod Gadus morhua stocks. Mar. Ecol. Prog. Ser. 235:235-256.

Mitchell, E. 1975. Trophic relationships and competition for food in Northwest Atlantic whales.Canadian Society of Zoologists Annual Meeting: 123–133.

Muus, B.J, Salomonsen., and F. C. Vibe 1990. Grønlands fauna: fisk, fugle, pattedyr. TheGreenland Fauna: fish, birds and mammals. In Danish). 463 pp.

National Economic Institute. 1995. Sögulegt Yfirlit Hagtalna 1945-1994. Thjódhagsstofnun.Reykjavík.

National Economic Institute. 2002. Thjódarbúskapurinn: Framvindan 2001 og Horfur 2002.Thjódhagsstofnun. Reykjavík.

Pálmadóttir, E. 1989. Fransí biskví (The history of French fisheries at Iceland; In Icelandic).Almenna bókafélagid, Reykjavík. 315 pp.

Pálsson, Ó. K. 1997. Fædunám thorsks (Gadus morhua). (Food and feeding of cod (Gadusmorhua).) In Fjölstofnarannsóknir (Multispecies Research) 1992–1995, pp. 177–191. Ed. by J.Jakobsson and Ó. K. Pálsson. Hafrannsóknastofnunin (Marine Research Institute), Reykjavík,Technical Report, 57 (In Icelandic).

Perry, R. K. 1986. Bathymetry. In B. G. Hurdle (Ed.). The Nordic Seas. Springer-Verlag, NewYork, pp. 211-236. Thjódhagsstofnun Íslands

Sæmundsson, B. 1934. Probable influence of changes in temperature on the marine fauna ofIceland. Rapp. P.-v. Réun. Cons. Perm. Int. Explor. Mer 86: 3-6.

Schmidt., J. 1909. The distribution of the pelagic fry and spawning regions of the Gadoids in theNorth-N-Atlantic from Iceland to Spain. Rapp. P-v. Réun. Cons. perm. int. Explor. Mer, 10: 1-229.

Schopka, S.A. 1994. Fluctuations in the cod stock off Iceland during the twentieth century inrelation to changes in the fisheries and environment. In Jakobsson, J. et al. (Eds). Cod andClimate, ICES mar. Sci. Symp., 198: 175-193.

Sigurjónsson J., Galan, A. and Víkingsson, G. A. 2000. A note of stomach contents of minkewhales (balaenoptera acutorostrata) in Icelandic waters. NAMMCO Sci. Publ. 2: 82-90.

Sigurjónsson, J. 1995. On the life history and autecology of North Atlantic rorquals. In Whales,Seals, Fish and Man, pp. 425–441. Ed. by A. Schytte-Blix, K. Wallöe, and Ö. Ulltang. ElsevierScience.

Sigurjónsson, J., and Víkingsson, G. A. 1997. Seasonal abundance of and estimated foodconsumption by cetaceans in Icelandic and adjacent waters. Journal of Northwest AtlanticFisheries Science, 22: 271–287.

Skúladóttir, U. and G. Pétursson 1999. Defining popultions of northern shrimp (Pandalus borealis(Krøyer 1848)) in Icelandic waters using the maximum length and maturity ogive of females.Rit Fiskideildar 16, 1: 247-262.

Smed, J. 1980. Temperature of the waters off Southwest and South Greenland during theICES/ICNAF salmon tagging experiment in 1972. Rapp. P-v. Réun. Cons. perm. int. Explor.Mer, 176: 18-21.

Page 69: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–69 DO NOT CIRCULATE OR CITE

Sólmundsson, J. 1997. Fæda Grálúdu (Reinhardtius hippoglossoides) á íslenskum hafsvædum.(The food of Greenland halibut (Reinhardtius hippoglossoides) in Icelandic waters.) InFjölstofnarannsóknir (Multispecies Research) 1992–1995, pp. 101–110. Ed. by J. Jakobssonand Ó. K. Pálsson. Hafrannsóknastofnunin (Marine Research Institute), Reykjavík, TechnicalReport, 57 (In Icelandic).

Solow.R.M. 1970. Growth Theory: An Exposition. The Radcliffe lectures. Claredon Press. Oxford.Statistics Iceland. 1997. Icelandic Historical Statistics. Hagstofa Íslands. Reykjavík: Framvindan

2001 og Horfur 2002. . Thjódhagsstofnun ÍslandsStatistics Iceland. 1997. Icelandic Historical Statistics. Hagstofa Íslands. ReykjavíkStefánsson, U. and J. Ólafsson 1991. Nutrients and fertility of Icelandic waters. Rit Fiskideildar.

12, 3: 1-56.Tåning, Å.V. 1934. Survey of long distance migrations of cod in the North Western Atlantic

according to marking experiments. Rapp. P-v. Réun. Cons. perm. int. Explor. Mer, 89, (3): 5-11.

Tåning, Å.V. 1937. Some features in the migration of cod. J. Cons. int. Explor. Mer 12: 1-35.Tåning, Å.V. 1948. On changes in the Marine Fauna of the North-N-Western Atlantic Area, with

special reference to Greenland. Rapp. et Proc. Verb. ICES 125: 26-29.Thor, J. Th. 1992. Breskir togarar og Íslandsmid 1889-1916 (British trawlers and Fishing Grounds

at Iceland 1889-1916. In Icelandic). Hid íslenska bókmenntafélag, Reykjavik. 237 pp.Thórdardóttir, Th. 1977. Primary production in North Icelandic waters in relation to recent climatic

changes. In M. J. Dunbar (Ed.), Polar Oceans. Proceedings of the Polar Oceans Conferenceheld at McGill University, Montreal, 1974. Arct. Inst. Amer., Canada. pp 655-665.

Thórdardóttir, Th. 1984. Primary production north of Iceland in relation to water masses inMay–June 1970–1980. ICES CM 1984/L:20.

Vestergaaard, N. and R. Arnason. 2003. On the Relationship between Greenland’s GrossDomestic Product and her Fish Exports: Empirical Estimates. Discussion paper

Vilhjálmsson, H. 1994. The Icelandic capelin stock. Capelin, Mallotus villosus (Müller), in theIceland–Greenland–Jan Mayen area. Rit Fiskideildar, 13. 281 pp.

Vilhjálmsson, H. 1997. Climatic variations and some examples of their effects on the marineecology of Icelandic and Greenland waters, in particular during the present century. RitFiskideildar, 15(1): 7-29

Vilhjálmsson, H. 2002. Capelin (Mallotus villosus) in the Iceland–East Greenland–Jan Mayenecosystem. ICES Mar. Sci. Symp. 216: 870-883.

Vilhjálmsson, H. and J. E. Carscadden 2002. Assessment surveys for capelin in the Iceland-EastGreenland-Jan Mayen aera, 1987-2001. ICES Mar. Sci. Symp. 216: 1096-1104.

Vilhjálmsson, H., and E. Fridgeirsson 1976. A review of 0-group surveys in the Iceland-EastGreenland area in the years 1970 - 1975. ICES Coop. Res. Rep. 54: 1-34.

Vilhjálmsson, H., and J. V. Magnússon 1984. Report on the 0-group fish survey in Icelandic andEast-Greenland waters, August 1984. ICES C.M. 1984/H:66, 26 pp.

Page 70: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–70 DO NOT CIRCULATE OR CITE

12.4 Northeastern Canada (Carscadden/Lilly/Schrank)

12.4.1 Introduction

Fisheries in ACIA Sub-region IV may be subdivided into those near the coast of Greenland, thosenear the coast of Canada, and those in deep waters of Baffin Bay and Davis Strait betweenGreenland and Canada. The whole Sub-region lies within the fisheries convention area of theNorthwest Atlantic Fisheries Organization (NAFO) (Fig. 1), and the fish stocks are currentlymanaged by either the coastal state or NAFO.

Along the northeast coast of Canada the ACIA study area is extended southward to the centralGrand Bank (46º N) in order to assess climate impacts on marine ecosystems that are comparable tothose considered by ACIA in the northeast Atlantic and around Iceland. This extension far to thesouth of the other geographic areas is necessitated by the presence of the Labrador Current, whichtransports cold water southward from Davis Strait, the Canadian Archipelago and Hudson Bay. Themedian southerly extent of sea ice is on the northern Grand Bank at approximately 47º N (Anon.2001a) and bottom water temperatures on the northern Grand Bank are below 0º C for considerableperiods. The southerly extent of cold conditions is also indicated by the regular presence of Arcticcod (Boreogadus saida) along the northeast coast of Newfoundland and their occasional occurrenceon the northern Grand Bank (Lilly et al. 1994; Lilly and Simpson 2000). In order to discuss thedynamics of Atlantic cod (Gadus morhua) and capelin (Mallotus villosus), and the fisheriesexploiting them, it will be convenient to include the whole of the putative stock areas, which extendto the southern boundary of NAFO Division 3L at 46º N for the major stocks of both Atlantic cod(NAFO Div. 2J+3KL) and capelin (NAFO Subarea 2 + Div. 3KL).

Insert Figure 12.4.1 here. Geographical overview of the Northeast Canadian ecosystem. Themap also shows the NAFO statistical areas referred to in the text.

In the study area, fish has dominated the history of Newfoundland since the time of Britishcolonization. The great interest in Britain for Newfoundland after its “discovery” during the Cabotvoyage of 1497 was generated by the presence of incredibly large amounts of codfish. Exploitationof this fishery by the British would reduce that country’s dependence on Iceland for fish, adependence that was creating difficulties. The French also saw the value of Newfoundland’sfishery, so possession of the island was one of the goals of both sides during the colonial wars ofthe eighteenth century.12 As an inducement for France to enter the revolutionary war on the side ofthe American colonies, Benjamin Franklin offered a share of the Newfoundland fishery to theFrench as bounty once the war was won (Burnett 1941, 210-211). Indeed, until the late 1800s,when a cross-island railroad was built, fishing was Newfoundland’s only industry. There thenstarted a series of diversification programs, which have continued, in one form or another until thepresent day. Although in the early 1970s Newfoundland had the world’s largest hydroelectric plantin Labrador, and despite the fact that during the twenty years following Confederation with Canada

12For the historical background of Newfoundland, see, H.A. Innis, The Cod Fisheries [Toronto: University ofToronto Press, revised edition, (1954)]; R.G. Lounsbury The British Fishery in Newfoundland, 1634-1763 [New Haven:Yale University Press, (1934)]; and St. John Chadwick Newfoundland: Island into Province [Cambridge: CambridgeUniversity Press, (1967)].

Page 71: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–71 DO NOT CIRCULATE OR CITE

in 1949 there were many attempts to diversify the economy, both small scale industries (e.g.,cement production, knitting mills, a shoe factory, a chocolate factory) and numerous large-scaleindustries (e.g., the Churchill Falls hydroelectric station, a petroleum refinery, a third paper mill,iron mines in Labrador), none of these industries made a dent in the dominance of the fishery inNewfoundland (Letto 1998). What did change in Newfoundland with Confederation and followingrevisions in Canadian federal/provincial intergovernmental arrangements, was that extremely largegovernment, health and education sectors arose which, as shares of the province’s gross domesticproduct (GDP), eclipsed the fishery. By 1971, the fish and fish processing sectors accounted forless than 5% of Newfoundland GDP. By 2001, their contribution fell further, to 3.5 %.

12.4.2 Ecosystem essentials

The ecosystem off northeastern Canada has been characterized by a relatively small number ofspecies, a few of which have historically occurred in high abundance (Bundy et al. 2000;Livingston and Tjelmeland 2000; Carscadden et al. 2001). The dominant fodder fish hashistorically been capelin, with Arctic cod more prominent to the north and sand lance (Ammodytesdubius) more prominent to the south on the plateau of Grand Bank. Herring (Clupea harengus) isfound only in the bays and adjacent waters. These four species of planktivorous fish feed mainly oncalanoid copepods and larger crustaceans, the latter being predominantly hyperiid amphipods to thenorth and euphausiids to the south. The dominant piscivorous fish has been Atlantic cod, butGreenland halibut (Reinhardtius hippoglossoides) and American plaice (Hippoglossoidesplatessoides) have also been important. Snow crab (Chionoecetes opilio) and northern shrimp(Pandalus borealis) have been the dominant benthic crustaceans. The top predators are harp seals(Pagophilus groenlandicus) and hooded seals (Cystophora cristata), which migrate into the areafrom the north during late autumn and leave in the spring. Other important predators includewhales, most of which migrate into the area from the south during late spring and leave during theautumn. The most important whales are humpback (Megaptera novaeangliae), fin (Balaenopteraphysalus), minke (Balaenoptera acutorostrata), sei (Balaenoptera borealis), sperm (Physetercatodon) and pilot (Globicephala melaena). Additional immigrants from the north during thewinter include many birds, such as thick-billed murres (Uria lomvia), northern fulmars (Fulmarusglacialis) and dovekies (Alle alle). Additional immigrants from the south during summer includeshort-finned squid (Illex illecebrosus), fish such as mackerel (Scomber scombrus) and bluefin tuna(Thunnus thynnus), and birds such as greater shearwater (Puffinus gravis) and sooty shearwater (P.griseus).

The Labrador/Newfoundland ecosystem experienced major changes during the last two decades ofthe 20th century. Atlantic cod and most other demersal fish, including species that were not targetedby commercial fishing, experienced declines to very low levels by the early 1990s (Atkinson 1994;Gomes et al. 1995). In contrast, snow crab (DFO 2002a) and especially northern shrimp (DFO2002b) surged during the 1980s and 1990s and now support the most important fisheries in thearea. Harp seals increased in abundance between the early 1970s and the mid-1990s (DFO 2000c).Capelin have been found in much reduced quantities in offshore acoustic surveys since the early1990s, but indices of capelin abundance in the inshore surveys have not experienced similardeclines, leaving the status of capelin uncertain and controversial (DFO 2000b, 2001). Atlantic

Page 72: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–72 DO NOT CIRCULATE OR CITE

salmon (Salmo salar), the major anadromous fish in the area, has declined in abundance, due inpart to lower survival at sea (Narayanan et al. 1995; DFO 2003b).

The waters of eastern Newfoundland have been fished for centuries, primarily for Atlantic cod butwith an increasing emphasis on other species during the latter half of the 20th century. Thesefisheries have undoubtedly had an influence on both the absolute abundance of some species andthe abundance of species relative to one another. However, the role of the fisheries in structuringthe ecosystem is often difficult to distinguish from the role of changes in the physical environment.The area experienced cooling during the last 3 decades of the 20th century, with particularly coldperiods in the early 1970s, early to mid-1980s and early 1990s. This cooling, which was associatedwith an intensification of the positive phase of the North Atlantic Oscillation (NAO) index(Colbourne et al. 1994; Mann and Drinkwater 1994; Narayanan et al. 1995; Colbourne andAnderson 2003), is thought by some to have played an important role in the dramatic decline ofAtlantic cod and other demersal fish and the increase in crustaceans, especially northern shrimp(Pandalus borealis).

12.4.3. Fish Stocks and Fisheries

Catches are taken from official NAFO statistics (as of January 2002) or from relevant assessmentdocuments if there is a difference between the two (see, for example, Anon. 2001b,c). Anoverview of the developments of the main fisheries off East Newfoundland and Labrador since1960 is shown in Figure 12.4.2.

Insert Figure 12.4.2 here. The catch of selected species in the waters east of Newfoundland andLabrador 1960-2000

Atlantic cod (Gadus morhua)The distribution of Atlantic cod along the coast of Canada has historically been from the northernLabrador Shelf southward beyond the limit of this study, although during the 1990s there have beenfew cod off Labrador. The Atlantic cod tends to live on the continental shelf, but it has been foundat depths to at least 850 m on the upper slope off eastern Newfoundland (Baird et al. 1992).

The European fishery for Atlantic cod off eastern Newfoundland began in the late 15th century. Forthe first few centuries fishing was by hook and line, so the cod were exploited only from late springto early autumn and only in shallow water along the coast and on the plateau of Grand Bank to thesoutheast of the island. There is evidence that local inshore overexploitation was occurring even inthe 19th century (Cadigan 1999), but improvements in gear and expansion of the area fished tendedto compensate for local reductions in catch rate. Annual landings increased through the 18th and 19th

centuries to about 300 000 t during the early decades of the 20th century. The deep waters remainedrefugia until the 1950s, when larger vessels with powered gurdys were introduced to exploit cod indeep coastal waters and trawlers from Europe started to fish the deeper water on the banks.Landings increased dramatically in the 1960s as large numbers of trawlers located and exploited theoverwintering aggregations on the edge of the Labrador Shelf and the Northeast NewfoundlandShelf. At the same time, the numbers of large cod in deep water near the coast of Newfoundlandare thought to have declined quickly as the longliner fleet switched to synthetic gillnets. Catches

Page 73: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–73 DO NOT CIRCULATE OR CITE

peaked at 894 000 t in 1968, and then declined steadily to only 143 000 t in 1978. FollowingCanada’s declaration of a 200 nautical mile Exclusive Economic Zone in 1977, the stock recoveredsomewhat and catches were in the range 230-270,000 t during most of the 1980s. However, catchesfell rapidly during the early 1990s as the stock declined to very low levels. A moratorium ondirected fishing was declared in 1992 (Fig.12.4.3). A small cod-directed fishery was opened in theinshore in 1998 but closed once again in 2003.

Insert Figure 12.4.3 here. The catch of ‘northern cod’ 1960-2000

Additional details on the history of the Atlantic cod fishery of Newfoundland and Labrador,including changes in technology and temporal variability in the spatial distribution of fishing effort,may be found in Templeman (1966), Lear and Parsons (1993), Hutchings and Myers (1995), Lear(1998), Neis et al. (1999) and Hutchings and Ferguson (2000).

Greenland halibut (Reinhardtius hippoglossoides)The Greenland halibut is distributed off West Greenland from Cape Farewell northward to about78º N and thence southward off eastern Canada to beyond the limit of this study. It is a deep-waterspecies, occurring in depths from about 200 m to at least 2200 m off West Greenland (Boweringand Brodie 1995).

The history of the fishery is complicated by temporal and spatial variation in effort and catch bydifferent fleets and by alleged underreporting of landings. For details of the fisheries, refer toBowering and Brodie (1995), Bowering and Nedreaas (2000) and Anon. (2001c).

The fishery off eastern Newfoundland dates back to the mid-19th century (Bowering and Brodie1995; Bowering and Nedreaas 2000). Catches from longlines were less than 1000 t annually untilthe early 1960s, when catches began to increase substantially. With the introduction of syntheticgillnets in the mid-1960s, catches in inshore waters rose quickly but declined within a few years,after which the fishery spread onto the shelf. Landings from offshore trawlers, mainly fromEuropean countries, also increased after the mid-1960s. Catches in SA 2 + Div. 3KL fluctuatednear 25-35 000 t from the late 1960s to the early 1980s, after which there was a gradual decline toabout 15 000 t in 1986. Landings increased dramatically in 1990 with the arrival of many non-Canadian trawlers that fished deep waters within Div. 3L. Catches during the next four years werehigh (estimated at 55-75 000 t in 1991 (Anon. 2001c)), declined substantially in 1995 due to aninternational dispute, and increased again during the late 1990s under NAFO quotas thatmaintained catches well below those of the early 1990s (Fig.12.4.4).

Insert Figure 12.4.4 here. The catch of Greenland halibut 1960-2000

The fishery to the north (NAFO SA 0), which has been conducted primarily with otter trawlers inthe second half of the year (Bowering and Brodie 1995), reported an average of 2 100 t during1968-1989 (including a high of 10 000 t in 1972). Catches increased dramatically to 14 500 t in1990 with increased effort by Canada, but declined to about 4 000 t from 1994 onward. Theselandings came mainly from off southeastern Baffin Island (Div. 0B). The fishery expanded evenfurther north into Baffin Bay (Div. 0A) in the mid- to late 1990s (Treble and Bowering 2002). Thisfishery, which extended to 73o N in 2002 (M.A. Treble, Dept. of Fisheries and Oceans, Central and

Page 74: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–74 DO NOT CIRCULATE OR CITE

Arctic Region, pers. comm.), has been limited by ice cover to the months of September toNovember.

Capelin (Mallotus villosus)Prior to the initiation of a commercial offshore fishery during the early 1970s, capelin were fishedon or near the spawning beaches. Annual catches, used for local consumption, may have reached 20000 to 25 000 t (Templeman 1968). Offshore catches by foreign fleets increased rapidly, peaking in1976 at about 370 000 tons, and then declined rapidly. This offshore fishery continued at a lowlevel until 1992. Catches in the offshore fishery were taken at different times of the year in differentareas. The spring fishery was dominated by large midwater trawlers operating in Div. 3L. Duringthe autumn, the offshore fishery first occurred in Div. 2J, off the coast of Labarador, and graduallymoved south into Div. 3K as the capelin migrated towards their overwintering area. This fisherywas also dominated by large midwater trawlers, which took mostly feeding capelin that wouldspawn the following year. During the late 1970s, as the foreign fishery declined, Canadianfishermen began fishing mature capelin near the spawning beaches to supply the Japanese marketfor roe-bearing females. This fishery expanded rapidly, exhibited highest catches during the 1980sand declined during the 1990s. The catches in the inshore fishery have generally not been as high asthose from the offshore foreign fishery (Fig. 12.4.5).

Insert Figure 12.4.5.here. The catch of capelin 1960-2000

Although the fishery on Canadian capelin has been relatively small when compared to the capelinfisheries in Iceland and the Barents Sea, there has been a concern about the potential impact of acommercial fishery on capelin because of its role as a forage species. However, Carscadden et al.(2001) concluded there is no scientific evidence to support the notion that the fishery in SA2+Div.3KL has had an impact on capelin abundance.

Herring (Clupea harengus)Herring in the Newfoundland and Labrador area are at the northern extent of their geographicrange. Stocks are coastal in distribution and stock abundance is small compared to other stocks inthe Atlantic. A peak catch of 30 000 t occurred in 1979, supported by strong year classes from the1960s. Recruitment since the 1960s has been lower. Stock sizes during the late 1990s have beenless than 90 000 t and annual catches have been less than 10 000 t (DFO 2000a).

Arctic (polar) cod (Boreogadus saida)The Arctic cod is broadly distributed through the Arctic and in cold waters of adjacent seas. Itoccurs on the shelf from northern Labrador to eastern Newfoundland, with the average size ofindividuals and the size of aggregations decreasing from north to south (Lear 1979). There has beenno directed fishery for Arctic cod off eastern Canada, but a small bycatch was reported in theRomanian capelin fishery in 1979 (Maxim 1980), and it is likely that small quantities were alsotaken in other years and by other countries.

Northern shrimp (Pandalus borealis)The northern shrimp is distributed off West Greenland from Cape Farewell northward to about 74ºN and thence southward off eastern Canada to beyond the limit of this study. The depth of highestconcentration tends to vary from area to area but is generally in the range 200-600 m.

Page 75: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–75 DO NOT CIRCULATE OR CITE

A fishery with large trawlers began off northeastern Canada in the late 1970s (Orr et al. 2001a). Forthe first decade most of the catch was taken from two channels in the central and southern LabradorShelf, but in the late 1980s there was an increase in effort and landings both to the south on theNortheast Newfoundland Shelf and to the north off northern Labrador. Catches increased above 25000 t by the mid-1990s. New survey technology introduced in 1995 indicated that commercialcatches of shrimp were very small relative to survey biomass, and quotas were increasedconsiderably during the late 1990s. Total landings from the area rose to more than 90 000 t by 2000(Fig.12.4.6).

Insert Fig. 12.4.6 here. The catch of northern shrimp 1960-2000

Much of the increase in catch from 1997 onward came from a new fleet of small (< 100 feet)vessels that fished with bottom trawls primarily on the mid-shelf. During the 1990s fishing alsoexpanded to Div. 3L in the south (Orr et al. 2001b).

Snow crab (Chionoecetes opilio)The snow crab is distributed from the central Labrador Shelf at approximately 55º N southward offeastern Canada to beyond the limit of this study. The depth distribution extends fromapproximately 50 m to 1400 m, but most of the fishery occurs in 100-500 m.

The fishery off eastern Newfoundland began in the late 1960s as a small bycatch fishery, but soonexpanded into a directed fishery with crab traps (pots) along most of the inshore areas of Div. 3KL(Taylor and O’Keefe 1999). During the late 1970s and early 1980s there was an increase in effortand an expansion of fishing grounds. Catches in Div. 3KL reached almost 14 000 t in 1981, butthen declined. In the mid-1980s there was expansion of the fishery into Div. 2J and new entrantsgained access to supplement declining incomes from the groundfish fisheries. The number ofparticipants and the area fished expanded further during the 1990s, and total 2J3KL catches rosequickly, reaching almost 55 000 t in 1999. Quotas and landings were reduced for the subsequenttwo years under concerns that the resource may have declined.

Commercial catch rates in Div. 3KL increased during the late 1970s to a peak in about 1981,declined to a nadir by 1987, and then increased during the late 1980s and early 1990s to a levelcomparable to that in the early 1980s (DFO 2002a). Catch rates remained high to the end of the1990s, despite the substantial increase in fishing effort and landings (Fig. 12.4.7). This reflects inpart an increase in the area fished, but it is thought that there must also have been an increase inproductivity.

Insert Figure 12.4.7 here. The catch of snow crab 1960-2000

Marine mammalsHarp seals summer in the Canadian Arctic or Greenland but winter and breed in Canadian Atlanticwaters. There are two major breeding herds, one that breeds in the Gulf of St Lawrence and onethat breeds off southern Labrador and Northeast Newfoundland (Bundy et al. 2000). Thepopulation of harp seals increased from fewer than 2 million animals during the early 1970s tomore than 5 million animals in the mid-1990s (Healey and Stenson 2000; Stenson et al. 2002). The

Page 76: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–76 DO NOT CIRCULATE OR CITE

population increase was largely the result of a reduction in the hunt after 1982 (Stenson et al.2002). The population stabilized when the hunt was increased in the mid-1990s. Reported Canadiancatches of harp seals include harvests off the coast of Newfoundland/Labrador (the “Front”) and inthe Gulf of St. Lawrence. Seals caught in both areas belong to the same population, the NorthwestAtlantic Harp Seals. The proportion of the population that occurs in the two areas varies amongyears, as does the relative number of seals caught in each area. Catches from both areas arecombined in official statistics and thus, the catches presented here are combined “Front” and Gulfof St Lawrence catches (Fig. 12.4.8).

Hooded seals are less abundant than harp seals. Whelping occurs on pack ice off northeastNewfoundland, in Davis Strait and in the Gulf of St Lawrence. Pups migrate into Arctic waters andremain there as juveniles. Adults migrate south in the fall and return to the Arctic in April (Bundyet al. 2000). The harvest of hooded seals (“Front” and Gulf of St Lawrence combined) is shown inFigure 12.4.8.

Insert Figure 12.4.8 here. The catch of harp and hooded seals 1952-2002. The catches alsoinclude seals taken in the Gulf of St. Lawrence

There has been no commercial whaling in the area since the late 1970s. Using north Atlanticpopulation estimates, assumed growth rates and assumed proportion of the total population in theNewfoundland and Labrador area, Bundy et al. (2000) estimated the population abundances at 33000 for humpback whales, 1000 for fin whales, 5 000 for minke whales, 1000 for sperm whales,1000 for sei whales and 9 000 for pilot whales.

12.4.4 Past climate variations and their impact on fish stocks

Atlantic cod (Gadus morhua)

The severe decline of Atlantic cod in the Newfoundland-Labrador area seems to have occurredfrom north to south. On the northern and central Labrador shelf (Div. 2GH), catches of 60-90 000 twere reported in the period 1965-1969, but catches declined to less than 5 000 t in most yearsduring the 1970s and early 1980s and to less than 1000 t during the latter half of the 1980s (Fig.12..4.9). There appear to be no analyses of factors that contributed to the decline in this northernarea.

Insert Figure 12.4.9 here. The catch of ‘northern cod’ by NAFO statistical areas 1960-2000

In the area from southern Labrador to the northern Grand Bank, the Div. 2J+3KL stock (the so-called “northern cod”) collapsed during the 1970s in response to severe over-fishing. The stockrecovered somewhat during the 1980s but collapsed to much lower levels in the late 1980s andearly 1990s. There is controversy as to whether there was a rapid but progressive decline from themid-1980s onward or a precipitous decline during the early 1990s (Atkinson and Bennett 1994;Shelton and Lilly 2000). Many studies (e.g. Hutchings and Myers 1994; Myers and Cadigan 1995;Hutchings 1996; Myers et al. 1996 a,b, 1997 a,b; Haedrich et al. 1997) have concluded that thefinal stock collapse was caused entirely by fishing activity (landed catch plus discards). However,

Page 77: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–77 DO NOT CIRCULATE OR CITE

several authors point to various ways in which the decline in water temperature and increase in icecover might have contributed to the collapse, either directly by reducing productivity (Mann andDrinkwater 1994; Drinkwater 2000, 2002; Parsons and Lear 2001) or indirectly by affectingdistribution (Rose et al. 2000).

Despite the many studies that have been conducted on this cod stock, there are few uncontesteddemonstrations of the influence of climate variability on stock dynamics. There is an expectation,that recruitment might be positively influenced by warm temperatures, because the stock is at thenorthern limit of the species’ range in North America (Planque and Frédou 1999). However, therehave been conflicting reports of whether such a relationship can be detected (deYoung and Rose1993; Hutchings and Myers 1994; Taggart et al. 1994; Planque and Frédou 1999). Part of theproblem is that recruitment has also been shown to be postively influenced by the number and sizeof spawners in the population (the spawning stock biomass or SSB) (Rice and Evans 1988; Myerset al. 1993; Hutchings and Myers 1994; Morgan et al. 2000; but see Drinkwater 2002). Bothtemperature and SSB tended to decline from the 1960s to the 1990s, increasing the difficulty ofdemonstrating a temperature effect. A reported positive relationship between recruitment andsalinity (Sutcliffe et al. 1983) was subsequently supported (Myers et al. 1993) and later rejected(Hutchings and Myers 1994; Shelton and Atkinson 1994) as data for additional years becameavailable. With respect to individual growth, a negative impact of temperature has been welldocumented (Krohn et al. 1997; Shelton et al. 1999). Additional aspects of cod biology thatchanged during the early 1990s, possibly in response to changes in the physical environment,include a delay in arrival on traditional inshore fishing grounds in early summer (Davis 1992), aconcentration of distribution toward the shelf break in autumn (Lilly 1994; Taggart et al. 1994), amove to deeper water in winter (Baird et al. 1992), and an apparent southward shift in distribution(Rose et al. 1994; Kulka et al. 1995; Rose and Kulka 1999).

Of considerable interest is the possibility that an increase in natural mortality contributed to therapid disappearance of cod during the early 1990s. The sharp decline in survey abundance indicesoccurred during a period of severe cold and extensive ice cover. A considerable decline in thecondition of the cod occurred at the same time, most notably in the north (Bishop and Baird 1994;Lilly 2001). Steep declines in abundance also occurred among other groundfish during the 1980sand 1990s (Atkinson 1994; Gomes et al. 1995), and while some contend that these declines werecaused by captures during fishing for cod and other species (Hutchings 1996; Haedrich and Fischer1996; Haedrich and Barnes 1997; Haedrich et al. 1997), there is no direct evidence of largeremovals. In the case of American plaice, a species that has been studied in detail, Morgan et al.(2002) demonstrated that the declines were too large to have been caused by fishing alone. Thecontribution of increased natural mortality to the decline of cod and other demersal fish in this areaduring the last two decades of the 20th century, and particularly during the early 1990s, remainsunresolved (Lilly 2002; Rice 2002).

The Div. 2J+3KL cod stock remained at a very low level one decade after declaration of themoratorium on directed fishing (Lilly et al. 2001; DFO 2003a). Recruitment to ages 0-2 remainedvery low, possibly due in part to a very small spawning stock biomass; juveniles in the offshoreappeared to experience very high mortality, possibly due in part to predation by harp and hoodedseals; and a directed fishery during 1998-2002 targeted the inshore aggregations, resulting inincreased mortality on the larger fish. In the presence of unquantified impacts of low spawning

Page 78: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–78 DO NOT CIRCULATE OR CITE

stock biomass, high predation and fishing, it is difficult to ascertain whether some aspect of oceanclimate has had a role in impeding recovery.

Greenland halibut (Reinhardtius hippoglossoides)The status of Greenland halibut in the northwest Atlantic has been uncertain because stock structureremains unclear, the fish undertake extensive ontogenetic migrations, there appear to have beenshifts in distribution, the fisheries have undergone numerous changes in fleet composition and areasand depths fished, and individual research surveys have been able to cover only part of the species’distribution. Nevertheless, there is evidence from various surveys that the biomass of Greenlandhalibut on the western side of the Labrador Sea declined substantially during the 1980s, with thedecline in Div. 0B and 2GH being most pronounced during the first half of the decade and thedecline in Div. 2J3K to the south being most pronounced during the second half of the 1980s andthe early 1990s (Bowering and Brodie 1995). Evidence for a decline in biomass in Div. 2J3K isalso seen in declining success of the gillnet fishery during the 1980s. The history of the fishexploited during the 1990s by the new deep-water trawler fishery to the south in Div. 3L is lessclear. It is possible that at least some of these fish migrated into the area from the shelf to the north(Bowering and Brodie 1995), in which case the decline in Div. 2J3K was partly due to a southwardshift in distribution.

Reasons for the declines in biomass and shift in distribution remain unclear. Bowering and Brodie(1995) drew attention to the decline in water temperatures on the shelf that occurred during theearly 1990s, but thought it was unlikely that such a change would in itself have affected thedistribution and abundance of Greenland halibut because this species occupies relatively deepwater. It may also be noted that much of the shift in distribution must have occurred during thelatter half of the 1980s, a period during which water temperatures were low but not as low asduring the early 1990s.

Variability in the physical environment has had no observed effect on either size at age (Boweringand Nedreaas 2001) or maturity at size or age (Morgan and Bowering 1997) during the period fromthe late 1970s to the mid-1990s.

Capelin (Mallotus villosus)The relationships between capelin biology and the physical environment have been extensivelystudied in the Newfoundland and Labrador area. Of particular relevance to this study is theobservation that many aspects of capelin biology changed during the 1990s and, initially, itappeared that these changes were the result of changes in water temperatures. However, watertemperatures during the latter half of the 1990s returned to normal yet the biological changes ofcapelin did not revert to earlier patterns. In the following section, the environmental variables andtheir links to capelin biology that would appear to be relevant in the event of global climate changeare briefly described.

The mean fish length of the mature population has been smaller during the 1990s (Carscadden et al.2002). These small sizes have been attributed to smaller fish sizes at age with fewer older and moreyounger fish in the population.

Page 79: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–79 DO NOT CIRCULATE OR CITE

Spawning occurs most often on fine gravel and grain size and beach orientation have been shown toexplain 61% of the variation in egg concentration among beaches (Vilhjalmsson 1994; Nakashimaand Taggart 2002).

Water temperature is also a determinant of capelin spawning. The recorded lowest and highesttemperatures for beach spawning in Newfoundland are 3.5° and 11.9 °C, with beach spawningceasing when temperatures exceeded 12.0 °C (Nakashima and Wheeler 2002).

Capelin eggs are very cold- and salinity-tolerant, surviving down to –5 °C and in salinities rangingfrom 3.4 ‰ to 34 ‰ (Davenport and Stene 1986; Davenport 1989). The rate of egg development inthe beach gravel is directly related to average incubation temperatures, which in turn aredetermined by water temperature, maximum and minimum air temperature and hours of sunlight(Frank and Leggett 1981).

Some capelin that move close to spawning beaches eventually spawn in deeper water adjacent tobeaches. This demersal spawning can occur simultaneously with intertidal spawning whentemperatures are suitable as well as when water temperatures at the beach-water interface becometoo warm. Egg mortality among these demersal eggs has been observed to be higher. Reproductivesuccess may have been poorer during the 1990s because water temperatures encountered when thecapelin reached the spawning beaches would have increased the incidence of demersal spawning(Nakashima and Wheeler 2002).

Historically, the spawning of capelin off Newfoundland beaches during June and July was apredictable event. During the early 1990s, spawning was later, and 80% of the variation inspawning time (1978-1994) was significantly and negatively related to mean fish size and seatemperatures that capelin experienced during gonadal maturation (Carscadden et al. 1997). Capelinspawning on Newfoundland beaches has continued to be delayed through 2000 in spite of the factthat sea temperatures have returned to normal. However, mean lengths of capelin have continued tobe small throughout this period.

There are historical reports of capelin occurring outside their normal distribution range. Unusualappearances in the Bay of Fundy and on the Flemish Cap were attributed to cooler watertemperatures while occurrences in Ungava Bay coincided with warming trends (summarized inFrank et al. 1996).

During the early 1990s, capelin distribution occurred more to the south, centred on the northernGrand Banks. Originally attributed to the colder water temperatures (Frank et al. 1996), this shiftwithin the normal distribution area has continued through 2000. Outside their normal distributionarea, capelin occurred on the Flemish Cap and eastern Scotian Shelf during the early 1990s andoccasionally during earlier cold periods. Capelin continued to appear both on the Flemish Cap andon the eastern Scotian Shelf through 2000. It appears that in the case of the large-scale changes indistribution outside their normal area, capelin are gradually declining in abundance as the waterswarm. Within their normal area of distribution, capelin have not returned to their usual pattern ofseasonal distribution as water temperatures increased, suggesting that other factors are operating.

Page 80: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–80 DO NOT CIRCULATE OR CITE

For mature capelin offshore during spring, Shackell et al. (1994) concluded that temperature wasnot used as a proximate cue during migration but that seasonal temperatures moderated offshorecapelin migration patterns through the regulation of growth, maturation, food abundance anddistribution.

Capelin typically move up and disperse in the water column at night and descend and aggregate atgreater depths during the day, but during spring surveys throughout the 1990s they remained deeperin the water column and exhibited reduced vertical migration (Shackell et al. 1994; Mowbray2002). This alteration in vertical distribution was not related to several factors tested, includingtemperature and predation, but may have been linked to feeding success of capelin (Mowbray2002).

Condition (calculated as a relationship between length and weight and regarded as a measure of“well-being”) of capelin was generally higher during the 1980s than during the 1990s. Conditionwas not related to temperature (Carscadden and Frank 2002).

Recruitment of beach-spawning capelin has been shown to be partly determined by the frequencyof onshore winds during larval residence in the beach gravel (Leggett et al. 1984; Carscadden et al.2000). The assessment of capelin has been especially problematic since the early 1990s, resulting inconsiderable uncertainty in the status of the stock. However, there is no evidence to indicate thatexploitation has had a direct affect on population abundance (Carscadden et al. 2001), suggestingthat any variations in abundance have been due to the environment. It is not known whether somechanges in biology such as condition and distribution have affected abundance, however, spawningtime and increased demersal spawning may be contributing to poor survival.

In summary, exploitation has not been shown to affect any aspect of capelin biology in this area.Although there have been several changes in capelin biology beginning in the early 1990s, there isno clear indication of what external factor(s) has (have) influenced the changes. Earlier studiesconcluded that temperature was an important factor for some changes, but it now seems unlikelythat temperature is the sole factor, given that water temperatures have returned to normal. There aresuggestions that changes in food supply (zooplankton) may be having an effect on capelin biologybut the exact mechanisms have not been identified.

Herring (Clupea harengus)Recruitment is positively related to warm overwintering water temperatures and high salinities(Winters and Wheeler 1987); these conditions seldom exist in this region and as a result, large year-classes rarely occur.

Arctic (polar) cod (Boreogadus saida)The distribution of Arctic cod off eastern Newfoundland expanded to the south and east during thecold period of the early 1990s (Lilly et al. 1994; Lilly and Simpson 2000).

Northern shrimp (Pandalus borealis)The shrimp resource off northeastern Canada has increased in density and expanded in distributionsince the mid-1980s. There is no indication that the increase in catch has negatively impacted theresource (DFO 2002b).

Page 81: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–81 DO NOT CIRCULATE OR CITE

There is considerable support for the hypothesis that the increase in northern shrimp offnortheastern Canada was, at least in part, a consequence of a reduction in predation pressure byAtlantic cod and other groundfish (Lilly et al. 2000; Bundy 2001; Worm and Myers 2002).Nevertheless, there is evidence that other factors were involved. For example, Lilly et al. (2000)noted that the increase in shrimp density on the Northeastern Newfoundland Shelf might havestarted during the early 1980s, a time during which the biomass of Atlantic cod was increasingfollowing its first collapse during the 1970s. Parsons and Colbourne (2000) found that catch perunit effort in the shrimp fishery on the central Labrador Shelf was positively correlated with icecoverage 6 years earlier. They suggested that cold water or ice cover itself was beneficial to theearly life history stages of shrimp in that area.

Snow crab (Chionoecetes opilio)The increased productivity of snow crab during the 1990s may have been caused, at least in part, bythe release in predation pressure from Atlantic cod and other demersal fish (Bundy 2001).However, the relationships between Atlantic cod and snow crab have not yet been explored to theextent that they have for Atlantic cod and northern shrimp. A preliminary examination of theinfluence of oceanographic conditions upon productivity of snow crab has revealed a negativerelationship between ocean temperature and lagged catch rates (DFO, 2002a). This has beeninterpreted to indicate that cold conditions early in the life cycle are associated with the productionof strong year-classes of snow crab in this area.

Marine mammalsTrends in populations of marine mammals during recent decades appear to be influenced primarilyby the commercial harvest. As populations of harp seals have increased in abundance, somechanges in biological characteristics indicate that density-dependence may be operating (Stenson etal. 2002). Density-independent influences may also regulate harp seal populations. Harp sealswhelp on ice and mortalities may vary annually depending on ice conditions during this criticalperiod. Mortalities of newly whelped pups may also occur during winter storms.

Concerns regarding the impact of predation by seals on commercial fish species increased as sealpopulations increased. It has been estimated that 74% (about 3 million t) of the total annualconsumption by four species of seals in eastern Canada occurred off southern Labrador andNewfoundland (Hammill and Stenson 2000). Predation by harp seals has been implicated in thelack of recovery of the northern cod stock (DFO 2003a), and predation on cod by hooded seals maybe large (DFO 2003a).

AquacultureSalmonid aquaculture does not occur in the ACIA region of Newfoundland because the water is toocold in winter. The primary cultured species in the area is blue mussel (Mytilus edulis). Productionof this species has grown over the last twenty years such that, during 2002, approximately 1 700 twere raised in the whole of Newfoundland.

12.4.5 Possible future impacts of climate change

Page 82: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–82 DO NOT CIRCULATE OR CITE

Two recent papers (Frank et al. 1990; Shuter et al. 1999) discussed the possible influence of globalwarming on ecosystems and fisheries off eastern Canada. Frank et al. (1990) predicted shifts in thegeographic ranges of several groundfish stocks because of redistribution of populations andchanging recruitment patterns. Stocks at the southern limit of a species’ distribution should retractnorthward, whereas those near the northern limit should expand northward. Frank et al. (1990) didnot make predictions specifically for Labrador and eastern Newfoundland, but events during thedecade following publication of their paper were in many respects opposite to these generalpredictions. The changes off Labrador and eastern Newfoundland were unprecedented and notpredicted, and illustrate the uncertainty of predictions, even on a regional scale and in the relativelyshort term. Shuter et al. (1999) had the advantage of witnessing the dramatic changes that occurredin the physical and biotic environment during the 1990s. They concluded that greenhouse gasaccumulation will lead to a warmer, drier climate and, for fisheries of Atlantic Canada, this willresult in a “decrease in overall sustainable harvests for coastal and estuarine populations due todecreases in freshwater discharge and consequent declines in ecosystem productivity”. For fisheriesin the Arctic, they predicted “increases in sustainable harvests for most fish populations due toincreased ecosystem productivity, as shrinkage of ice cover permits greater nutrient recycling”.

As noted repeatedly in this chapter, the relative importance of fishing and environment is difficultto determine for any species or group of species, so it is perhaps not surprising that the importanceattributed to each has varied among studies. It is also perhaps not surprising, given the differencesamong species in the magnitude of fishery removals relative to stock size, that opinion tends tofavor fishing as the dominant factor for some species and environment as the dominant factor forothers. For demersal fish, there are many statements to the effect that declines were caused entirelyby overfishing, but there is evidence that changes in oceanographic properties contributed tochanges in distribution and declines in productivity. For crab and especially shrimp, it has beensuggested that increases in biomass were simply a consequence of a release in predation pressurefrom Atlantic cod and perhaps other demersal fish, but again there is evidence that changes inoceanographic factors contributed to an increase in reproductive success. For capelin, mostinformation supports the hypothesis that fishing had little impact on population dynamics, and thatenvironmental factors were the primary determinant of stock size, wellbeing (growth andcondition), distribution and timing of migrations. For Arctic cod, fishing may be dismissed as acontributor to changes in distribution and biomass.

An important constraint to predicting changes in fish stocks and the fisheries that exploit them inthe waters off Labrador and eastern Newfoundland is uncertainty about the direction and magnitudeof change in important oceanographic variables. With respect to surface air temperature, somemodel outputs project a cooling over the central North Atlantic, and it is not clear where theLabrador/Newfoundland region lies within the gradation from significant warming in the highArctic to cooling over the central North Atlantic. In addition, there is no model for downscalingGCM output to specifics of the Labrador/Newfoundland area. As an example of the importance ofregional models, many large-scale models predict an increase in air temperature over theNorwegian and Barents seas, whereas simulations with one specific regional model (Furevik et al.2002) indicate that sea surface temperature in that area may decline in the next 20 years beforeincreasing later in the century. Another concern is that natural variability in a specific region, suchas the Labrador Shelf, may be much greater than variability in the global mean (Furevik et al.2002). Thus, a warming trend in shelf waters off Labrador and Newfoundland might be

Page 83: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–83 DO NOT CIRCULATE OR CITE

accompanied by substantial annual variability, such as was witnessed during the last three decadesof the 20th century, and it is even possible that the amplitude of that variability could increase. Forbiota, extreme events associated with this variability might be at least as influential as any long-term trend. For the Labrador Shelf and Northeast Newfoundland Shelf, it is probably at least asimportant to know how the North Atlantic Oscillation will behave (especially the intensity andlocation of the Icelandic Low) as it is to know that global temperature will rise.

In the absence of region-specific information, we assume that there will be a gradual warming ofshelf waters off Labrador and Newfoundland (Chapter 8). Using the events in West Greenlandduring the first half of the 20th century (Vilhjálmsson 1997) as a spatial/temporal analogue, onemight expect better recruitment success and northward expansion of Atlantic cod and some otherdemersal fish that live mainly on the shelf. Capelin also would be expected to shift northwards. Ifzooplankton abundance is enhanced by warmer water, capelin growth may improve. It is possiblethat many existing capelin spawning beaches will disappear with the predicted rise in sea level(Shaw et al. 1998). Depending on the increase in sea level, storm events and the availability ofglacial deposits, some beaches may move while others may disappear completely. While beach-spawning capelin can adapt to spawning on suitable sediment in deeper water, survival of eggs andlarvae appears to be adversely affected (Nakashima and Wheeler 2002), suggesting that a rise in sealevel would probably result in reduced survival and recruitment for capelin. A warming of seatemperatures might retard recruitment to snow crab and northern shrimp, so these species mightexperience gradual reductions in productivity. In summary, then, we might anticipate that a gradualwarming of sea temperature might promote a change back to a cod-capelin system from the presentsystem where snow crabs and northern shrimp are the major commercial species. In addition, bothcod and capelin might become more prominent off central Labrador than they were during the1980s.

One might also contemplate scenarios of no change in temperature or even a cooling. Given thattemperatures were generally below the long-term mean during the base-line period of the ACIAstudy, both of these scenarios might tend to favour the current balance of species in the system.

The above simple scenario of a gradual change back to a cod-capelin system under warmingconditions is very uncertain, not only because we have not clearly elucidated the influence ofoceanographic variability in the past, but also because it is likely that the dynamics of some speciesare now dominated by a different suite of factors than was the case in the past. It is highly likelythat the ecosystem off northeastern Canada changed substantially as a consequence of fishingduring the first four centuries following the arrival of fishermen from Europe, changed even furtherwith the increasingly intensive fishing of the 20th century, and altered very dramatically during thelast two decades of the century. It would be difficult to predict accurately the future state of thesystem, even without the added complications of climate change. The system could remain in itscurrent state, revert to some semblance of an historic state (or at least the state of the early 1980s),or evolve toward something previously unseen.

A gradual warming of shelf waters might also promote a movement of more southerly species intothe area. For example, haddock (Melanogrammus aeglefinus) might become more abundant on thesouthern part of Grand Bank, and expand into the study area. Migrants from the south, such as

Page 84: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–84 DO NOT CIRCULATE OR CITE

short-finned squid, mackerel and bluefin tuna, might occur more regularly and in greater quantitiesthan during the 1980s and early 1990s.

Changes in sea ice, identified in Table 8.2, may have a negative effect on harp seals, the mostimportant marine mammal predator in the area. Sea ice duration is projected to shorten and it is notknown whether harp seals would be able to adjust their breeding time to accommodate this change.A decrease in sea ice extent may not seriously affect harp seals because they would probably shifttheir distribution along with the ice. However, thinner sea ice may be deleterious, resulting inincreased pup mortality. Increases in regional storm intensities (Table 8.2) may also result in higherpup mortalities if such storms occur during the critical period shortly after birth (G. Stenson, Dept.of Fisheries and Oceans, St. John’s, NL, pers. comm.). Changes in seal abundance may cascadethrough the ecosystem, since seals are important predators on many fodder fish and commerciallyimportant groundfish (Bundy et al. 2000; Hammill and Stenson 2000), and are thought by some tobe important in impeding the recovery of cod (DFO 2003a) and thus maintaining the presentbalance within the ecosystem.

In addition to uncertainty regarding the response of individual species and the ecosystem as awhole, there is uncertainty regarding the influence of changing ice cover on the fisheriesthemselves. A reduction in the extent and duration of sea ice may permit fishing further to the northand would increase the period during which ships would have access to certain fishing grounds. Inparticular, these changes in sea ice cover would affect the Greenland halibut and shrimp fisheries inBaffin Bay and Davis Strait. For example, an increased open water season and extended fishingperiod is thought to have the potential to increase harvest of Greenland halibut at the time ofspawning (late winter/spring).

A reduction in ice cover (see Table 8.2) might negatively impact Greenland halibut fisheries thatare conducted through fast ice. For example, a fishery that was developed in Cumberland Sound onBaffin Island in the late 1980s has developed into a locally important enterprise (Crawford 1992;Pike 1994). The fishery is conducted with longlines set through ice over deep (600-1125 m) water,with the season extending in some years from mid-January to June. Since the mid 1990’s, theseason has been shorter, typically from early February to May (M.A. Treble, Dept. of Fisheries andOceans, Central and Arctic Region, pers. comm.). To date, attempts at fishing during the openwater season have not proved successful. The catches have been low and the fish seem to bedispersed. It is not clear whether the fish would be present in commercial concentrations in thewinter/spring if ice were not present. Even if they were, the absence of ice would certainly affectthe conduct of the fishery.

12.4.6 The economic and social role of fisheries in Newfoundland

From an economy based primarily on the fishery, Newfoundland has, along with most of NorthAmerica, moved to a service economy. By 1971, for instance, the fishing and fish processingsectors accounted for less than 5% of Newfoundland’s GDP, whereas the service sector accountedfor more than half the GDP. Mining accounted for 11% and construction for 18% (but that includedconstruction of some of the large diversification projects). Nearly twenty years later, in 1989,shortly before the groundfish collapse, the fishery harvesting and processing sectors together

Page 85: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–85 DO NOT CIRCULATE OR CITE

accounted for slightly more than 5%, service industries had grown to 68%, mining had fallen to lessthan 6% and construction to 8%.13 By 2001, fishery harvesting and processing sectors accountedfor only 3% of GDP, the service industries remained constant at 68%, construction had slipped to4.7% and conventional mining to 3%. Oil production, a new industry in Newfoundland, alreadyaccounted for 8.4% of the provincial G.D.P., with every prospect of growing.14 Mining was alsoexpected to see a resurgence with the potential opening of a large nickel mine in Labrador.

Yet, while the fishery may not be of such great importance to the overall economy ofNewfoundland, the fishery continues to dominate completely the economy in rural areas of theprovince, and, perhaps even more important, its culture. After fifty years, there is still a dailyFisheries Broadcast on radio and when the Canadian Broadcasting Company decided to cancel theweekly television program Land and Sea, which often focuses on the fishery, public pressureforced the crown corporation to continue the program. With the fishery in deep trouble in 1989, thedominant newspaper in the province, The Evening Telegram commented in an editorial entitled“Too Many Fishermen?”, that “Newfoundland’s fishery must eventually be expanded anddiversified so that it can employ more people, not fewer...” (The Evening Telegram, June 1, 1989).

With the spectacular change in fisheries employment that accompanied the collapse of the northerncod stock, there has been a sharp reemphasis on economic diversification. The two areas that areparaded as holding the hope of the future are tourism and various information technology fields.15

Some progress has been made in both areas (fishing vessels converted to tour boats for whalewatching, and many bed and breakfast establishments, for instance). But progress in these areashas been uneven, and some government policies have been inconsistent with the objective ofpromoting these industries. For instance, despite its interest in developing tourism, theNewfoundland government decommissioned or privatized a substantial number of the parks in theprovince’s extensive parks system (Overton 2001). How many of those sold remain in operation asparks is unknown. The real problem with the emphasis on tourism and information technology isthat it is occurring at just the time when it seems to be happening throughout the world. Whyshould Newfoundland have an advantage over the rest of the world in either of these fields? Onlytime will tell how this effort of diversification works out; it is too early to judge.

The story of the Newfoundland fishery does not end with the collapse of the groundfishery, itscatastrophic consequences for many families and the serious pressures it placed on the government.Two critical sets of changes have occurred since 1992. First, the fishery management process hasevolved. Second, shellfish have replaced groundfish as the main components of fishery landings inNewfoundland.

Fishery management by means of restricting total allowable catches began in Newfoundland duringthe mid-1970s. Following a particularly dramatic race to the fish in 1981, the government imposed,

13For a more extensive discussion of these issues, see Schrank et al. 1992.

14The source of the 2001 figures is the web site of the Newfoundland Statistics Agency:www.nfstats.gov.nf.ca/statistics/GDP/GDP_Industry.asp.

15See, for instance, the emphasis on these fields in Government of Newfoundland and Labrador 2001.

Page 86: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–86 DO NOT CIRCULATE OR CITE

at the industry’s request, enterprise allocations on the offshore groundfish fleet in 1982. Theseallocations were divisible and transferable in the year in which they were assigned. Governmentemphasized that these were rights to fish as opposed to property rights, which could be permanentlysold. By this time, gear and geographic restrictions had been imposed, as had limited entry to non-groundfish inshore fisheries. With the expansion of the crab fishery, enterprise allocations havebeen assigned to this inshore sector. These allocations are not transferable. The federal governmenthas relinquished the licensing of fishermen in favor of the provincial Professional Fish Harvester’sCertification Board. This board was established as part of a professionalization program and itlicenses harvesters either as apprentices or in one of two professional classes. The federalgovernment, in turn, in 1996 established “core” fishing enterprises for the inshore fisheries. Seniorlevel professional fishermen who met certain conditions were declared the heads of core fishingenterprises. Approximately 5 500 of these were established and the government claimed that noadditional core licenses would be issued; the only way that fishermen could obtain such statuswould be to buy out the core license of an existing core fisherman. In an attempt to reduce thenumber of fishermen, the federal government has bought approximately 1 500 core licenses,claiming that these will not be reissued. The final major recent changes in the management systemare that (1) species license fees (access fees) are no longer nominal, for most important species theyare based on the anticipated gross income from the fishery; and (2) there is now an extensivesystem of public consultations before recommendations concerning total allowable catches aremade to the Minister of Fisheries and Oceans.16

The value of landings in all Newfoundland fisheries in 1991 was $282 838 000 9. This figure fellduring the first year of the moratorium to $194 745 000 and then doubled to 388 700 000 by theyear 2000. Viewed alternatively, the period of the northern cod moratorium saw an increase of one-third in the real value of Newfoundland fish and shellfish landings. But, while in 1991 43% of thevalue of landings was for cod, in 2000 46% was accounted for by crab and a further 30% byshrimp. With some cod stocks reopened on a limited basis for commercial fishing, cod in 2000accounted for 9% of the total landings, but this figure had earlier, in 1996, fallen to less than one-half of one percent.17

For environmental reasons, be it the lack of predation or favorable climatic conditions, the shellfishpopulation has surged and there has been a nearly complete conversion of Newfoundland’s fishingindustry from groundfish to shellfish. The conversion took years to occur. Labor requirements forshellfish are lower than for groundfish and, having higher unit prices, the shellfish quantities thatyield these landings figures are much smaller than for groundfish.

Since all major species are under quota, it is obvious that total allowable catches for shrimp andcrab have been increasing rapidly, along with the number of harvesting and processing licenses forshrimp and even more for crab. The number of shrimp harvesting licenses has risen from 19 in1986 and 57 in 1991 to 438 in 2000. The numbers of crab licenses in the same years were 274, 721and 3333 (Corbett 2002). There has been considerable controversy about whether the old error ofissuing too many licenses has occurred for crab. Even with the exodus from the province, there is

16For a discussion of the current fishery management system, see Schrank and Skoda 2003.

17DFO web site: www.dfo-mpo.gc.ca/communic/statistics/landings.

Page 87: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–87 DO NOT CIRCULATE OR CITE

still a 17.5% seasonally adjusted unemployment rate in the province (where the national figure is7.4%)18 and there is continued pressure to open closed plants and increase licenses for crabfishermen. While the number of crab fishing licenses has increased substantially, the increase in thenumber of crab processing plants has been modest.19

12.4.7 Past variations in the fishing industry and their economic and social impacts

The politically popular, but economically nonsensical, Evening Telegram editorial, referred toearlier, appeared in the context of a fishery that had long been in trouble. Without going too deeplyinto the past, we cite first a provincially financed report in 1967 that supported the trend away froma seasonal inshore fishery in Newfoundland towards a capital-intensive year-round offshore fishery.The report also noted that the “number of people dependent on the fishery should be reduced”(Pushie 1967, 185). This has been a recurring theme. In 1970, the federal fisheries departmentappealed to the Canadian cabinet to permit the department to establish regulations that would leadto a reduction in the number of Atlantic Canadian (meaning mainly Newfoundland) fishermen byhalf.

The authority to affect these changes was denied (Schrank 1995, 287-288). The fishery repeatedlyfaced crises, was repeatedly studied, and the conclusion was repeatedly drawn that there were toomany people dependent on the fishery. One study estimated that of the then 35 000 licensedfishermen, only 6 000 could be supported unsubsidized by the fishery at better than a poverty levelincome. The same study concluded that for every dollar of fish landed, there was a dollar ofsubsidy (Schrank et al. 1986). In 1976, with the extension of coastal states’ fishery jurisdiction to200 miles from shore, the two Canadian departments concerned (fisheries and regional economicexpansion) both published reports stating that there was sufficient extra capacity in the industry thatno significant employment benefits could be expected from the expansion of jurisdiction(Government of Canada, 1976a, 1976b). Two years later, a provincial government report said muchthe same thing, that the Canadianized fishery, when fully developed, could employ only 9 000inshore fishermen (Government of Newfoundland and Labrador, 1978). Yet in the face of popularpressure, both of the federal departments, along with the provincial government, licensed andsubsidized a tremendous expansion in the physical capacity of the industry: from 13 636 registeredfishermen in 1975 to 33 640 in 1980, from 9 517 registered inshore vessels in 1976 to 19 594 in1980, from a fish freezing capacity of 181 000 t in 1974 to 467 000 t in 1980, from net fishermen’sunemployment insurance benefits of $30 724 000 in fiscal 1976/77 to $66 060 000 in 1980/81, andfrom outstanding loans of the provincial Fisheries Loan Board (to finance inshore vessels) from$36 869 000 in fiscal year 1976/77 to $78 558 000 in 1980/81 (Schrank 1995, 91).20 By 1981 the

18These are February 2003 figures obtained on March 25, 2003 from the following web sites:a. www.nlstats.gov.nl.ca/statistics/Labour/b. www.statcan.ca/

19Government of Newfoundland and Labrador, Department of Fisheries and Aquaculture, List of LicensedProcessors, various years.

20All dollar figures in this discussion of the Newfoundland fishery are real (base year = 2000) United Statesdollars.

Page 88: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–88 DO NOT CIRCULATE OR CITE

expansion had stopped and the two federal departments agreed that no further expansion of fishprocessing facilities would be built with federal financing (LeBlanc and De Bané, 1981). But bythat time, the damaging expansion had already occurred and, in the face of the anti-inflationrecession of the early Reagan years in the United States (where the overwhelming proportion ofNewfoundland fish production was sold), the market for Newfoundland fish products shrankdramatically and most Newfoundland fish processing companies faced bankruptcy.

As a result, the industry was financially, but not structurally, reorganized and the massive industrialclosures implicit in bankruptcy were averted.21 Yet, starting in 1982, inshore groundfish catches felland after 1986 offshore catches followed. By 1992, the situation was so bad that a completemoratorium on the commercial catching of the formerly massive northern cod stock was put intoplace. Shortly thereafter, nearly all Newfoundland groundfisheries were closed and the moratoriumwas extended to non-commercial fishing (Schrank 1995, 285-286). The closure of most of theseground fisheries continues, in whole or in part, to the present day. The closure of the Newfoundlandgroundfisheries is reputed to have involved the largest mass layoff of labor in Canadian history. Insocial terms (because of the mass layoff), in biological terms (because of the decimation of the fishstock), and in governmental financial terms (because of billions of dollars spent primarily on theincome maintenance of fishermen and fish plant workers) the moratoria were disasters.

Yet the response of government, of the industry, and of the public to the moratoria is instructive ofwhat might happen in the face of the climate change, the subject that this panel has been formed toanalyze. The cause of the stock destruction in Newfoundland waters may be debated; neverthelessthe dramatic effect on the fish population is incontrovertible. Should significant changes inenvironmental conditions occur worldwide in the future, and should these changes have substantialeffects on commercial fish stocks, then the recent Newfoundland experience may provide atemplate (favorable or unfavorable) for what might be expected to happen elsewhere. Moreover, theNewfoundland experience may well suggest the desirability of alternative policies.

The decline of the cod fisheries was better understood after the publication of the reports ofAlverson (1987) and Harris (1989, 1990). That major problems were developing in thegroundfishery was no longer debatable. In response to the decline of the fishery, the federalgovernment in 1990 introduced the Atlantic Fisheries Adjustment Program (AFAP).22 Theemphasis was on the word “adjustment”. People were to be retired from the fishery, ruralcommunities were to receive money to help them diversify their economies away from the fisheryand steps were to be made to increase scientific knowledge of the problem of declining fish stocks.But only a few hundred fishermen left the industry. With the shock of the total closure of thecommercial northern cod fishery in July 1992, the federal government, anticipating that the fisherywould revive in two years, created the Northern Cod Adjustment and Recovery Program (NCARP).Once again there was to be an emphasis on people adjusting out of the fishery. This program calledfor early retirement of fishermen, buybacks of fishing licenses, training of fishermen for othertrades, and income maintenance payments to fishermen and fish plant workers. One-third of the9,000 northern cod fishermen and one-half of the 10,000 plant workers affected by the northern cod

21This was the work of the Kirby Task Force. See Kirby 1982.

22The discussion of AFAP, NCARP and TAGS is drawn from Schrank 1997.

Page 89: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–89 DO NOT CIRCULATE OR CITE

shutdown were expected to leave the fishery during the life of the program. In fact, only 1 436 tookearly retirement and 876 fishermen sold their licenses. Fishermen were simply not convinced thatthe shutdown would continue for very long, and they believed that the government would continueto support them until the fishery reopened. Under those conditions, why should relativelyuneducated, potential low end laborers leave an industry in which they were skilled and for whichthey were trained from an early age? One reason for low educational levels (until 1991, less thanhalf of Newfoundland’s adult population had completed high school) was the fishing tradition.Boys would start fishing with their fathers at a young age and would look forward to leaving schoolas soon as possible to join the family fishing enterprise. Boys and girls who had little interest in fishharvesting might anticipate a life of work in the local fish plant, jobs which, for the most part,required little formal education.

Of course, the fish did not return after two years, and in fact have still not returned a decade afterthe commencement of the moratorium. As NCARP was expiring, a new adjustment program wasput in place. The Atlantic Groundfish Strategy (TAGS) was to be a five-year program of incomemaintenance and adjustment (license buybacks and retirements) in which a 50% reduction infishing capacity was anticipated. Once again, there was very little movement of men and womenout of the fishery. Their reluctance to abandon the fishery at this time simply replicated the reasonsthey did not leave under the previous NCARP program.

As TAGS was winding down towards the end of the 1990s, the government started taking a harderline. There would be a post-TAGS program, but it would not resemble its predecessors: incomemaintenance would be severely cut and many people would be removed from the program. Withgovernment financial support gone, or going, and the fish still not returned, people started seeingthe handwriting on the wall. Finally there was an exodus from the fishery.

Between 1986 and 1991, the Newfoundland population stagnated, at least partially because of adramatic drop in family size. From the highest birth rate of Canadian provinces, by 1991Newfoundland had the lowest birth rate. In addition, there had always been modest migration outof the province. But in the five years between 1991 and 1996 (from the year after the start of AFAPto halfway through TAGS) the province’s population actually fell by 3%. With the continuingmoratorium and the change in government policy, the exodus ballooned and in the five yearsbetween 1996 and 2001 there was a further drop in the province’s population of 7%.23

While such a population drop in a province over a decade is dramatic, this figure of 10% actuallyhides the severity of the impact of the fishery collapse on Newfoundland’s rural communities.Trepassey, on the southern shore of the Avalon Peninsula, was the location of a major groundfishprocessing plant. Newfoundland groundfish operators had been under an Enterprise Allocationscheme since 1982. With the drop in fish stocks towards the end of the 1980s, enterpriseallocations were cut and, in response, a number of fish plants were closed. One of the first to go, in1990, was that in Trepassey. The result was that a town with 1 375 inhabitants in 1991 had shrunkby more than 35% to 889 in 2001. Many rural communities in Newfoundland have seen populationdeclines over the past decade of from 15% to 30% (Statistics Canada 2002b).

23 Census figures for 2001 are taken from Statistics Canada 2002a. Those for 1991 and 1996 are from StatisticsCanada 1999.

Page 90: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–90 DO NOT CIRCULATE OR CITE

12.4.8. Economic and social impacts of global warming: Possible scenarios

Global warming can be expected to cause changes in the size of fish stocks. The effects are unlikelyto be greater than the historical changes described above. When dealing with human society,reponses to impulses are not “natural” in the sense that a climatic change “causes” a humanresponse. The human response is determined by the magnitude of the stimulus together with thepolitical response of the society. In this sense, societal responses to global warming will not bequalitatively different from society’s responses to past changes. The political system will respond,and the details of that response are impossible to predict. But as we have seen, we have modelsfrom past experience. It should be noted that many of the federal and provincial interventions in thefishery since 1992 have seemed ill thought out, often unfair, and have raised controversy.

A recent case provides an illustration. The groundfish plant in Twillingate had once been owned bythe largest fish company in Newfoundland, Fishery Products International, Ltd., but had been soldto another operator in the mid-1980s (Schrank et al. 1995, 389). With the northern cod moratoriumin 1992, the plant was shut and remained so until 2002 when it opened as a shellfish plant withmore technologically sophisticated equipment than had been used for groundfish. The MarineInstitute, a branch of Memorial University, introduced for fish plant workers a course to teach themto use the new equipment, charging more than $400 per person for the course. Unemployed peoplereceiving (un)employment insurance would have the fee paid by the federal government. Otherswould have to pay the fee themselves. Most unemployed not on employment insurance cannotafford the fee. Working people, wanting either the higher paying jobs in the fish plant, or formerfish plant workers who want to get back into the industry, would have to quit their jobs to take thecourse, unless they were lucky enough to be granted time off, which is unlikely in unskilled trades.But if they quit their jobs and completed the course, there was no guarantee that they would get ajob in the Twillingate, or any other technologically advanced, fish plant (CBC 2002).Unfortunately, every aspect of the long, drawn out, adjustment process which started with thedecline of the groundfishery in the 1980s has been characterized by a deep sense of unfairness.

The Newfoundland experience illustrates that a “catastrophic” event concerning the fishery leads tosevere adjustment problems, and the adjustment may take an extended period, but it also raises newpotential for a successful industry. The questions seem to be:

• how to convince participants in the industry that there is indeed a crisis,• what adjustments need to be made,• what is the role of government, and• how to protect the new fishery from the mistakes of the failed old fishery.

Where does this leave us with respect to predicting the future socio-economic effects of long termglobal warming? This is one case where it is easier to prescribe than to predict. The Newfoundlandexperience of the past quarter century has demonstrated reactions to expanding fish populations andto shrinking fish populations. In neither case was the reaction, in terms of government action orpolitical pressure, appropriate. During the expansionary period of the late 1970s, the fisheryexpanded far too much, with excessive and ultimately largely immobile labor and capital enteringthe industry. Whereas a properly managed fishery would have restricted the expansion of factors of

Page 91: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–91 DO NOT CIRCULATE OR CITE

production, the expansion continued virtually without letup until it was stopped by a generaleconomic crisis. It was understood at the time that employment expansion was an incorrectresponse. Should fish stocks off Newfoundland increase during the next twenty or fifty years, careshould be taken to either (1) restrict by government regulation the magnitude of any expansion ofcapital and labour in the fishery; (2) ensure that such economic incentives are in place thatexcessive growth does not occur; or (3) combine the two.

Should fish stocks decrease during that period, then it should be made clear from the start theendless subsidies will not be forthcoming. License buyouts, even generous license buyouts of coreenterprises for instance, would be helpful. While these are payments are subsidies, they are limitedin scope and time and would have the desirable effect of permanently shrinking the factor base ofthe industry. Such a policy would have been much cheaper during the 1990s, and much lessstressful for the fishing families affected, than the approach of offering income maintenancepayments that was taken.

A gradual warming of shelf waters will likely lead to increased opportunity for aquaculture.Warmer temperatures and shorter periods of ice cover may enable mussel farming to be moreproductive. Warmer waters may also promote the development of Atlantic cod farming. If inshorewaters become sufficiently warm, it may be possible to farm Atlantic salmon along the east coast ofNewfoundland. Such farming is presently not possible because water temperature in winter dropsbelow the lethal temperature for salmon.

12.4.9 The ability to cope with climate change

As has been repeated often in this chapter, climate change will affect all aspects of the fishery: therange of existing species; the relative populations of different species; and the economiccircumstances of people who depend on the fishery for a living. How ready are the economic andsocial systems of Newfoundland to cope with these changes?

The past experience of Newfoundland does not make one optimistic. When the fishery wasrevitalized after the declaration of the 200-mile limit, the economic structure of the fisheryoverexpanded with largely immobile capital and labor, leading the way to subsequent disaster.When the cod fishery started to decline in the late 1980s, a half dozen years passed before many ofthe necessary adjustments occurred. Will the situation be different in response to climate change?

The answer depends largely on whether the lessons of the past quarter century have been learned,and whether the social and political systems are prepared to adjust. Both the expansion of thegroundfishery in the late 1970s and failure of the fishery to adjust to decimated stocks in the earlyand mid-1990s were largely the result of subsidization. During the 1970s the expansion was largelyfinanced by the federal and provincial governments. Despite the efforts of the federal governmentto adjust fishermen out of the industry during the 1990s, the adjustment programs turned intoincome maintenance programs, which in effect encouraged fishermen to remain in the industry inthe hope that the fish would return. It was only when the subsidies were substantially reduced after

Page 92: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–92 DO NOT CIRCULATE OR CITE

1998 that a significant number of fishermen left the industry (based on the assumption thatdepartures from the fishery are reflected in the census figures).

Subsidies, of course, are not the results of whimsical acts of governments or politics but areresponses to real social and economic concerns. As long as the government considers the survivalof small rural communities a major priority, subsidies to the fishery (the primary industry in thesecommunities) will be continued. As long as these subsidies exist, the response of people to changesin the industry will be slow. In the absence of subsidies, economic forces will require change,probably fairly rapid change if the fishery is declining. If the biological base of the fishery isexpanding, there is always the possibility that the industry will overexpand without governmenthelp. Government financial assistance would virtually ensure that overexpansion would occur.

To the extent that the adjustments induced by climate change cause human suffering, thegovernment can be expected to ameliorate the situation and ease the necessary transitions. But thereis strong precedent for transition programs being transformed into short run income maintenanceprograms. If that were to happen again, the process of adjustment would likely be as long, painfuland wasteful as in the past.

In short, it is impossible to predict how ready society is to cope with the effects of climate change.The response mechanisms are not automatic and political reactions will play a major role.

12.4.10 References

Alverson, D. L. (Chair). 1987. A Study of Trends of Cod Stocks off Newfoundland and FactorsInfluencing their Abundance and Availability to the Inshore Fishery: A Report to theHonourable Tom Siddon, Minister of Fisheries of Canada. n.p.: Task Group on NewfoundlandInshore Fisheries.

Anon. 2001a. Sea ice climatic atlas, East Coast of Canada, 1971-2000. Canadian Ice Service,Minister of Public Works, Canada.

Anon. 2001b. STATLANT 21 reported catches by stock tabulated against STACFIS estimates,1985-1999. NAFO SCS Doc. 01/5.

Anon. 2001c. Northwest Atlantic Fisheries Organization Scientific Council Report, June 2001.339p.

Atkinson, D.B. 1994. Some observations on the biomass and abundance of fish captured duringstratified-random bottom trawl surveys in NAFO Divisions 2J and 3KL, autumn 1981-1991.NAFO Sci. Coun. Studies 21: 43-66.

Atkinson, D.B., and Bennett, B. 1994. Proceedings of a northern cod workshop held in St. John’s,Newfoundland, Canada, January 27-29, 1993. Can. Tech. Rep. Fish. Aquat. Sci. 1999: 64 p.

Atkinson, D.B., Rose, G.A., Murphy, E.F., and Bishop, C.A. 1997. Distribution changes andabundance of northern cod (Gadus morhua), 1981-1993. Can. J. Fish. Aquat. Sci. 54 (Suppl.1): 132-138.

Baird, J.W., Stevens, C.R., and Murphy, E.F. 1992. A review of hydroacoustic surveys conductedduring winter for 2J3KL cod, 1987-1992. CAFSAC Res. Doc. 92/107. 14 p.

Page 93: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–93 DO NOT CIRCULATE OR CITE

Bishop, C.A., and Baird, J.W. 1994. Spatial and temporal variability in condition factors ofDivisions 2J and 3KL cod (Gadus morhua). NAFO Sci. Coun. Studies 21: 105-113.

Bowering, W.R., and Brodie, W.B. 1995. Greenland halibut (Reinhardtius hippoglossoides). Areview of the dynamics of its distribution and fisheries off Eastern Canada and Greenland. InDeep-water fisheries of the North Atlantic oceanic slope. Edited by A.G. Hopper. KluwerAcademic Publishers, Boston. pp. 113-160.

Bowering, W.R., and Nedreaas, K.H. 2000. A comparison of Greenland halibut (Reinhardtiushippoglossoides (Walbaum)) fisheries and distribution in the Northwest and NortheastAtlantic. Sarsia 85: 61-76.

Bowering, W.R., and Nedreaas, K.H. 2001. Age validation and growth of Greenland halibut(Reinhardtius hippoglossoides (Walbaum)): a comparison of populations in the Northwest andNortheast Atlantic. Sarsia 86: 53-68.

Bundy, A. 2001. Fishing on ecosystems: the interplay of fishing and predation in Newfoundland-Labrador. Can. J. Fish. Aquat. Sci. 58: 1153-1167.

Bundy, A., Lilly, G.R., and Shelton, P.A. 2000. A mass balance model of the Newfoundland-Labrador Shelf. Can. Tech. Rep. Fish. Aquat. Sci. 2310: xiv + 157 p.

Burnett, E. C. 1941. The Continental Congress. New York: Macmillan.Cadigan, S.T. 1999. Failed proposals for fisheries management and conservation in Newfoundland,

1855-1880, p. 147-169. In D. Newell and R.E. Ommer [eds.] Fishing places, fishing people:traditions and issues in Canadian small-scale fisheries. University of Toronto Press, Toronto.

Carscadden, J.E., and Frank, K.T. 2002. Temporal variability in the condition factors ofNewfoundland capelin (Mallotus villosus) during the past two decades.– ICES Journal ofMarine Science, 59: 950-958.

Carscadden, J.E., Frank, K.T., and Leggett, W.C. 2000. Evaluation of an environment-recruitmentmodel for capelin (Mallotus villosus). ICES Journal of Marine Science, 57: 412-418.

Carscadden, J.E., Frank, K.T., and Leggett, W.C. 2001. Ecosystem changes and the effects oncapelin (Mallotus villosus), a major forage species. Canadian Journal of Fisheries and AquaticSciences, 58: 73-85.

Carscadden, J.E., Nakashima, B.S., and Frank, K.T. 1997. Effects of fish length and temperature onthe timing of peak spawning in capelin (Mallotus villosus). Canadian Journal of Fisheries andAquatic Sciences, 54: 781-787.

Carscadden, J. E., Montevecchi, W.A., Davoren, G. K., and Nakashima, B. S. 2002. Trophicrelationships among capelin (Mallotus villosus) and seabirds in a changing ecosystem. ICESJournal of Marine Science, 59:1027-1033.

CBC. 2002. Item broadcast on The Morning Show in St. John’s, Newfoundland on CanadianBroadcasting Company Radio 1 in the half hour starting at 7:00 a.m. on Wednesday, April 24,2002.

Colbourne, E.B., and Anderson, J.T. 2003. Biological response in a changing ocean environment inNewfoundland waters during the latter decades of the 1900s. ICES mar. Sci. Symp. 219: 169-181.

Colbourne, E.B., Narayanan, S., and Prinsenberg, S. 1994. Climate changes and environmentalconditions in the Northwest Atlantic, 1970-1993. ICES mar. Sci. Symp. 198: 311-322.

Page 94: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–94 DO NOT CIRCULATE OR CITE

Corbett. F. 2002. Fax message from F. Corbett, Department of Fisheries and Oceans to A. St. Croix,May 27, 2002.

Crawford, R.E. 1992. Life history of the Davis Strait Greenland halibut, with reference to theCumberland Sound fishery. Can. Manuscr. Rep. Fish. Aquat. Sci. 2130.

Davenport, J. 1989. The effects of salinity and low temperature on eggs of the Icelandic capelinMallotus villosus. Journal of the Marine Biological Association of the United Kingdom, 69: 1-9.

Davenport, J., and Stene, A. 1986. Freezing resistance, temperature and salinity tolerance in eggs,larvae and adults of the capelin, Mallotus villosus, from Balsfjord. Journal of the MarineBiological Association of the United Kingdom, 66: 145-157.

Davis, M.B. 1992. Description of the inshore fishery during 1991 for 2J3KL cod as reported byinshore fisherpersons. Canadian Atlantic Fisheries Scientific Advisory Committee Res. Doc.93/37.

deYoung, B., and Rose, G.A. 1993. On recruitment and distribution of Atlantic cod (Gadusmorhua) off Newfoundland. Can J. Fish. Aquat. Sci. 50: 2729-2741.

DFO, 2000a. East and southeast Newfoundland Atlantic herring. DFO Science Stock Status ReportB2-01 (2000).

DFO, 2000b. Capelin in Subarea 2 + Div. 3KL. DFO Science Stock Status Report B2-02 (2000).DFO, 2000c. Northwest Atlantic harp seals. DFO Science Stock Status Report E1-01 (2000).DFO, 2001. Capelin in Subarea 2 + Div. 3KL - Update. DFO Science Stock Status Report B2-02

(2001).DFO, 2002a. Newfoundland and Labrador snow crab. DFO Science Stock Status Report C2-01

(2002).DFO, 2002b. Northern shrimp (Pandalus borealis) – Div. 0B to 3K. DFO Science Stock Status

Report C2-05 (2002).DFO, 2003a. Northern (2J+3KL) Cod. DFO Canadian Science Advisory Secretariat Stock Status

Report 2003/018.DFO, 2003b. Newfoundland and Labrador Atlantic salmon 2003 stock status update. DFO

Canadian Science Advisory Secretariat Stock Status Update Report 2003/048.Drinkwater, K.F. 2000. Changes in ocean climate and its general effect on fisheries: examples from

the north-west Atlantic, p. 116-136. In D. Mills [ed.] The ocean life of Atlantic salmon:environmental and biological factors influencing survival. Fishing News Books, Oxford. 228p.

Drinkwater, K.F. 2002. A review of the role of climate variability in the decline of northern cod.American Fisheries Society Symposium, 32: 113-130.

Frank, K.T. and Leggett, W.C. 1981. Prediction of egg development and mortality rates in capelin(Mallotus villosus) from meteorological, hydrographic, and biological factors. CanadianJournal of Fisheries and Aquatic Sciences, 38: 1327-1338.

Frank, K.T., Carscadden, J.E., and Simon, J.E. 1996. Recent excursions of capelin (Mallotusvillosus) to the Scotian Shelf and Flemish Cap during anomalous hydrographic conditions.Canadian Journal of Fisheries and Aquatic Sciences, 53: 1473-1486.

Page 95: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–95 DO NOT CIRCULATE OR CITE

Frank, K.T., Perry, R.I., and Drinkwater, K.F. 1990. Predicted response of Northwest Atlanticinvertebrate and fish stocks to CO2-induced climate change. Trans. Amer. Fish. Soc. 119:353-365.

Furevik, T., Drange, H., and Sorteberg, A. 2002. Anticipated changes in the Nordic Seas marineclimate: scenarios for 2020, 2050 and 2080. Institute of Marine Research, Bergen. Fisken ogHavet No. 4 – 2002.

Gomes, M.C., Haedrich, R.L., and Villagarcia, M.G. 1995. Spatial and temporal changes in thegroundfish assemblages on the north-east Newfoundland/Labrador Shelf, north-west Atlantic,1978-1991. Fish. Oceanogr. 4: 85-101.

Government of Canada. 1976a. Policy for Canada’s Commercial Fisheries. Ottawa: Fisheries andMarine Service of the Department of the Environment.

Government of Canada. 1976b. Community and Employment Implications of Restructuring theAtlantic Fisheries. Ottawa: Analysis and Liaison Branch of the Department of RegionalEconomic Expansion.

Government of Newfoundland and Labrador 1978. Setting a Course: A Regional Strategy forDevelopment of the Newfoundland Fishing Industry to 1985. St. John’s, Newfoundland:Department of Fisheries.

Government of Newfoundland.and Labrador 2001. Securing Our Future Together: Final Report onthe Renewal Strategy for Jobs and Growth.. St. John’s, Newfoundland: Government ofNewfoundland and Labrador.

Haedrich, R.L., and Barnes, S.M. 1997. Changes over time of the size structure in an exploited shelffish community. Fisheries Research 31: 229-239.

Haedrich, R.L., and Fischer, J. 1996. Stability and change of exploited fish communities in a coldocean continental shelf ecosystem. Senckenbergiana maritima 27: 237-243.

Haedrich, R.L., Fischer, J., and Chernova, N.V. 1997. Ocean temperatures and demersal fishabundance on the northeast Newfoundland continental shelf, p. 211-221. In L.E. Hawkins andS. Hutchinson, with A.C. Jenson, M. Sheader, and J.A. Williams [eds.] The responses ofmarine organisms to their environment. University of Southampton.

Hammill, M.O. and Stenson, G.B. 2000. Estimated prey consumption by Harp seals (Phocagroenlandica), Hooded seals (Cystophora cristata), Grey seals (Halichoerus grypus) andHarbour seals (Phoca vitulina) in Atlantic Canada. J. Northw. Atl. Fish. Sci. 26: 1-23.

Harris, L. (Chairman). 1989. Independent Review of the State of the Northern Cod Stock.. Ottawa:Department of Fisheries and Oceans of the Government of Canada.

Harris, L. (Chairman). 1990. Independent Review of the State of the Northern Cod Stock: FinalReport of the Northern Cod Review Panel. Ottawa: Department of Supply and Services of theGovernment of Canada.

Healey, B.P., and Stenson, G.B. 2000. Estimating pup production and population size of thenorthwest Atlantic harp seal (Phoca groenlandica). DFO Can. Stock Assess. Sec. Res. Doc.2000/081.

Hutchings, J.A. 1996. Spatial and temporal variation in the density of northern cod and a review ofhypotheses for the stock’s collapse. Can J. Fish. Aquat. Sci. 53: 943-962.

Page 96: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–96 DO NOT CIRCULATE OR CITE

Hutchings, J.A., and Ferguson, M. 2000. Temporal changes in harvesting dynamics of Canadianinshore fisheries for northern Atlantic cod, Gadus morhua. Can. J. Fish. Aquat. Sci. 57: 805-814.

Hutchings, J.A., and Myers, R.A. 1994. What can be learned from the collapse of a renewableresource? Atlantic cod, Gadus morhua, of Newfoundland and Labrador. Can. J. Fish. Aquat.Sci. 51: 2126-2146.

Hutchings, J.A., and Myers, R.A. 1995. The biological collapse of Atlantic cod off Newfoundlandand Labrador: an exploration of historical changes in exploitation, harvesting technology, andmanagement, p. 39-93. In R. Arnason and L. Felt [eds.] The north Atlantic fisheries:successes, failures, and challenges. The Institute of Island Studies, Charlottetown, PrinceEdward Island.

Kirby, M. J. L. (Chairman). 1982. Navigating Troubled Waters – A New Policy for the AtlanticFisheries. Ottawa: Department of Supply and Services of the Government of Canada.

Krohn, M., Reidy, S., and Kerr, S. 1997. Bioenergetic analysis of the effects of temperature andprey availability on growth and condition of northern cod (Gadus morhua). Can. J. Fish.Aquat. Sci. 54(Suppl. 1): 113-121.

Kulka, D.W., Wroblewski, J.S., and Narayanan, S. 1995. Recent changes in the winter distributionand movements of northern Atlantic cod (Gadus morhua Linnaeus, 1758) on theNewfoundland-Labrador Shelf. ICES J. mar. Sci. 52: 889-902.

Lear, W.H. 1979. Distribution, size, and sexual maturity of Arctic cod (Boreogadus saida) in thenorthwest Atlantic during 1959-1978. Can. Atl. Fish. Sci. Adv. Comm. Res. Doc. 79/17.

Lear, W.H. 1998. History of fisheries in the Northwest Atlantic: the 500-year prespective. J.Northw. Atl. Fish. Sci. 23: 41-73.

Lear, W.H., and Parsons, L.S. 1993. History and management of the fishery for northern cod inNAFO Divisions 2J, 3K and 3L, p. 55-89. In L.S. Parsons and W.H. Lear [eds.] Perspectiveson Canadian marine fisheries management. Can. Bull. Fish. Aquat. Sci. 226.

LeBlanc, R. and de Bané, P. 1981. “A Proposed Memorandum of Understanding between DFO andDREE Respecting Joint Action for Development of the Fisheries of Eastern Canada,” signedby Roméo LeBlanc and Pierre De Bané, March 20, 1981. National Archives of Canada,RG124, Accession 87-88/35, File #4560-F-0 Part 1.

Leggett, W.C., Frank, K.T.,and Carscadden, J.E. 1984. Meteorological and hydrographic regulationof year-class strength in capelin (Mallotus villosus). Canadian Journal of Fisheries andAquatic Sciences, 41: 1193-1201.

Letto, D. M. 1998. Chocolate Bars and Rubber Boots: the Smallwood Industrialization Plan.Paradise, Newfoundland: Blue Hill Publishers.

Lilly, G.R. 1994. Predation by Atlantic cod on capelin on the southern Labrador and NortheastNewfoundland shelves during a period of changing spatial distributions. ICES mar. Sci.Symp. 198: 600-611.

Lilly, G.R. 2001. Changes in size at age and condition of cod (Gadus morhua) off Labrador andeastern Newfoundland during 1978-2000. ICES CM 2001/V:15. 34 p.

Lilly, G.R. 2002. Ecopath modelling of the Newfoundland Shelf: observations on data availabilitywithin the Canadian Department of Fisheries and Oceans. In Information supporting past andpresent ecosystem models of northern British Columbia and the Newfoundland Shelf. Edited

Page 97: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–97 DO NOT CIRCULATE OR CITE

by T. Pitcher, M. Vasconcellos, S. Heymans, C. Brignall and N. Haggan. Fisheries Centre,University of British Columbia, Canada. Fisheries Centre Research Reports 2002 10 (1) pp.22-27.

Lilly, G.R., and Simpson, M. 2000. Distribution and biomass of capelin, Arctic cod and sand lanceon the Northeast Newfoundland Shelf and Grand Bank as deduced from bottom-trawlsurveys. DFO Can. Stock Assess. Sec. Res. Doc. 2000/091. 40 p

Lilly, G.R., Hop, H., Stansbury, D.E., and Bishop, C.A. 1994. Distribution and abundance of polarcod (Boreogadus saida) off southern Labrador and eastern Newfoundland. ICESC.M.1994/O:6. 21 p.

Lilly, G.R., Parsons, D.G., and Kulka, D.W. 2000. Was the increase in shrimp biomass on theNortheast Newfoundland Shelf a consequence of a release in predation pressure from cod? J.Northw. Atl. Fish. Sci. 27: 45-61.

Lilly, G.R., Shelton, P.A., Brattey, J., Cadigan, N.G., Healey, B.P., Murphy, E.F., and Stansbury,D.E. 2001. An assessment of the cod stock in NAFO Divisions 2J+3KL. DFO Can. StockAssess. Sec. Res. Doc. 2001/044. 148 p.

Livingston, P.A., and Tjelmeland, S. 2000. Fisheries in boreal ecosystems. ICES J. mar. Sci. 57:619-627.

Mann, K.H., and Drinkwater, K.F. 1994. Environmental influences on fish and shellfish productionin the Northwest Atlantic. Environ. Rev. 2: 16-32.

Maxim, C. 1980. Capelin (Mallotus villosus) catch, effort, and biological charaacteristics in theRomanian fishery in Division 2J, September-October 1979. Northw. Atl. Fish. Organ. SCRDoc. 80/4, Serial No. N030.

Morgan, M.J., and Bowering, W.R. 1997. Temporal and geographic variation in maturity at lengthand age of Greenland halibut (Reinhardtius hippoglossoides) from the Canadian north-westAtlantic with implications for fisheries management. ICES J. mar. Sci. 54: 875-885.

Morgan, M.J., Brodie, W.B., and Kulka, D.W. 2002. Was over-exploitation the cause of the declineof the American plaice stock off Labrador and northeast Newfoundland? Fisheries Research57: 39-49.

Morgan, M.J., Shelton, P.A., Stansbury, D.P., Brattey, J., and Lilly, G.R. 2000. An examination ofthe possible effect of spawning stock characteristics on recruitment in 4 Newfoundlandgroundfish stocks. Can. Stock Ass. Secr. Res. Doc. 2000/028.

Mowbray, F. K. 2002. Changes in the vertical distribution of capelin (Mallotus villosus) offNewfoundland. ICES Journal of Marine Science, 59:000-000.

Myers, R.A., and Cadigan, N.G. 1995. Was an increase in natural mortality responsible for thecollapse of northern cod? Can. J. Fish. Aquat. Sci. 52: 1274-1285.

Myers, R.A., Barrowman, N.J., and Hutchings, J.A. 1997a. Inshore exploitation of NewfoundlandAtlantic cod (Gadus morhua) since 1948 as estimated from mark-recapture data. Can. J. Fish.Aquat. Sci. 54(Suppl. 1): 224-235.

Myers, R.A., Barrowman, N.J., Hoenig, J.M., and Qu, Z. 1996a. The collapse of cod in EasternCanada: the evidence from tagging data. ICES J. mar. Sci. 53: 629-640.

Myers, R.A., Drinkwater, K.F., Barrowman, N.J., and Baird, J.W. 1993. Salinity and recruitment ofAtlantic cod (Gadus morhua) in the Newfoundland region. Can. J. Fish. Aquat. Sci. 50:1599-1609.

Page 98: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–98 DO NOT CIRCULATE OR CITE

Myers, R.A., Hutchings, J.A., and Barrowman, N.J. 1996b. Hypotheses for the decline of cod inthe North Atlantic. Mar. Ecol. Prog. Ser. 138: 293-308.

Myers, R.A., Hutchings, J.A., and Barrowman, N.J. 1997b. Why do fish stocks collapse? Theexample of cod in Atlantic Canada. Ecological Applications 7: 91-106.

Nakashima, B.S., and Taggart, C.T. 2002. Why are some beaches better for spawning than others?ICES Journal of Marine Science, 59:000-000.

Nakashima, B.S., and Wheeler, J.P. 2002. Capelin (Mallotus villosus) spawning behaviour inNewfoundland waters- the interaction between beach and demersal spawning. ICES Journalof Marine Science, 59: 000-000.

Narayanan, S., Carscadden, J., Dempson, J.B., O’Connell, M.F., Prinsenberg, S., Reddin, D.G., andShackell, N. 1995. Marine climate off Newfoundland and its influence on Atlantic salmon(Salmo salar) and capelin (Mallotus villosus). In Climate change and northern fishpopulations. Edited by R.J. Beamish. Can. Spec. Publ. Fish. Aquat. Sci. 121. Pp. 461-474.

Neis, B., Schneider, D.C., Felt, L., Haedrich, R.L., Fischer, J., and Hutchings, J.A. 1999. Fisheriesassessment: what can be learned from interviewing resource users? Can. J. Fish. Aquat. Sci.56: 1949-1963.

Orr, D., Parsons, D.G., Veitch, P.J., and Sulllivan, D.J. 2001a. Northern shrimp (Pandalus borealis)off Baffin island, Labrador and northeastern Newfoundland – first interim review. DFO Can.Sci. Adv. Sec. Res. Doc. 2001/043.

Orr, D.C., Parsons, D.G., Veitch, P., and Sulllivan, D. 2001b. An update of information pertainingto northern shrimp (Pandalus borealis) and groundfish in NAFO Divisions 3LNO. NAFOSCR Doc. 01/186, Serial No. N4576.

Overton, J. 2001. “Official Acts of Vandalism: Privatizing Newfoundland’s Provincial Parks in the1990s” in D. McGrath (ed.), From Red Ochre to Black Gold. St. John’s, Newfoundland:Flanker Press.

Parsons, D.G., and Colbourne, E.B. 2000. Forecasting fishery performance for northern shrimp(Pandalus borealis) on the Labrador Shelf (NAFO Divisions 2HJ). J. Northw. Atl. Fish. Sci.27: 11-20.

Parsons, L.S., and Lear, W.H. 2001. Climate variability and marine ecosystem impacts: a NorthAtlantic perspective. Progress in Oceanography 49: 167-188.

Pike, D.G. 1994. The fishery for Greenland halibut (Reinhardtius hippoglossoides) in CumberlandSound, Baffin Island, 1987-1992. Can. Tech. Rep. Fish. Aquat. Sci. 1924.

Planque, B., and Frédou, T. 1999. Temperature and the recruitment of Atlantic cod (Gadusmorhua). Can. J. Fish. Aquat. Sci. 56: 2069-2077.

Pushie, G. F. (Chairman). 1967. Report of the Royal Commission on the Economic Prospects ofNewfoundland and Labrador. St. John’s, Newfoundland: Office of the Queen’s Printer.

Rice, J. 2002. Changes to the Large Marine Ecosystem of the Newfoundland-Labrador Shelf. InLarge Marine Ecosystems of the North Atlantic. Edited by K. Sherman and H.R. Skjoldal.Elsevier Science. pp. 51-103

Rice, J.C., and Evans, G.T. 1988. Tools for embracing uncertainty in the management of the codfishery of NAFO divisions 2J+3KL. J. Cons. int. Explor. Mer 45: 73-81.

Page 99: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–99 DO NOT CIRCULATE OR CITE

Rose, G.A., and Kulka, D.W. 1999. Hyperaggregation of fish and fisheries: how catch-per-unit-effort increased as the northern cod (Gadus morhua) declined. Can. J. Fish. Aquat. Sci.56(Suppl. 1): 118-127.

Rose, G.A., Atkinson, B.A., Baird, J., Bishop, C.A., and Kulka, D.W. 1994. Changes indistribution of Atlantic cod and thermal variations in Newfoundland waters, 1980-1992.ICES mar. Sci. Symp. 198: 542-552.

Rose, G.A., deYoung, B., Kulka, D.W., Goddard, S.V., and Fletcher, G.L. 2000. Distributionshifts and overfishing the northern cod (Gadus morhua): a view from the ocean. Can. J. Fish.Aquat. Sci. 57: 644-663.

Schrank, W. E. 1995. “Extended Fisheries Jurisdiction: Origins of the Current Crisis in AtlanticCanada’s Fisheries”, Marine Policy, XIX, 285-299.

Schrank, W. E. 1997. “The Newfoundland Fishery: Past, Present and Future” in S. Burns (ed.)Subsidies and Depletion of World Fisheries: Case Studies. Washington, D.C.: World WildlifeFund.

Schrank, W. E., Roy, N. and Tsoa, E. 1986. “Employment Prospects in a Commercially ViableNewfoundland Fishery: An Application of ‘An Econometric Model of the NewfoundlandGroundfishery’,” Marine Resource Economics, III. 237-263.

Schrank, W. E., Roy, N., Ommer, R. and Skoda, B. 1992. “An Inshore Fishery: A CommerciallyViable Industry or an Employer of Last Resort,” Ocean Development and International Law,XXIII, 335-367.

Schrank , W. E.and Skoda, B. 2003. “The Cost of Marine Fishery Management in Eastern Canada:Newfoundland, 1989/1990 to 1999/2000”, 91-128 in W.E. Schrank, R. Arnason and R.Hannesson (eds). The Cost of Fisheries Management. Aldershot: Ashgate Publishers.

Schrank, W. E., Skoda, B., Parsons, P. and Roy, N. 1995. “The Cost to Government of Maintaininga Commercially Unviable Fishery: The Case of Newfoundland 1981/82 to 1990/91", OceanDevelopment and International Law, XXVI, 357-390.

Shackell, N.L., Carscadden, J.E., and Miller, D.S. 1994. Migration of pre-spawning capelin(Mallotus villosus) as related to temperature on the northern Grand Bank, Newfoundland.ICES Journal of Marine Science, 51:107-114.

Shaw, J., Taylor, R. B., Forbes, D. L., Ruz, M.-H., and Solomon, S. 1998. Sensitivity of the coastsof Canada to sea-level rise. Geological Survey of Canada, Bulletin 505, 79 p.

Shelton, P.A., and Atkinson, D.B. 1994. Failure of the Div. 2J3KL cod recruitment prediction usingsalinity. DFO Atl. Fish. Res. Doc. 94/66.14 p.

Shelton, P.A., and Lilly, G.R. 2000. Interpreting the collapse of the northern cod stock from surveyand catch data. Can. J. Fish. Aquat. Sci. 57: 2230-2239.

Shelton, P.A., Lilly, G.R., and Colbourne, E. 1999. Patterns in the annual weight increment forDiv. 2J+3KL cod and possible prediction for stock projection. J. Northw. Atl. Fish. Sci. 25:151-159.

Shuter, B.J., Minns, C.K., Regier, H.A., and Reist, J.D. 1999. Canada country study: Climateimpacts and adaptations: fishery sector. Vol. VII, National Sectoral Volume, pp 258-317.

Statistics Canada 2002b. 2001 Community Profiles. Web site:www12.statcan.ca/english/profil01/PlaceSearchForm1.cfm (March 18, 2003).

Page 100: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–100 DO NOT CIRCULATE OR CITE

Statistics Canada. 1999. Profile of Census Divisions and Subdivisions. Ottawa: Industry Canada,Catalog No. 95-182-XPB.

Statistics Canada. 2002a. Population and Dwelling Counts (On-line), 2001 Census. StatisticsCanada Catalogue No. 93F0051XIE, web site:www.statcan.ca/english/IPS/Data/93F0051XIE.htm (March 17, 2003).

Stenson, G.B., Hammill, M.O., Kingsley, M.C.S., Sjare, B., Warren, W.G., and Myers, R.A. 2002. Is there evidence of increased pup production in northwest Atlantic harp seals, Pagophilus groenlandicus? ICES J. mar. Sci. 59: 81-92.Sutcliffe, W.H., Loucks, R.H., Drinkwater, K.F., and Coote, A.R. 1983. Nutrient flux onto the Labrador Shelf from Hudson Strait and its biological consequences. Can. J. Fish. Aquat. Sci. 40: 1692-1701.Taggart, C.T., Anderson, J., Bishop, C., Colbourne, E., Hutchings, J., Lilly, G., Morgan, J.,

Murphy, E., Myers, R., Rose, G., and Shelton, P. 1994. Overview of cod stocks, biology, andenvironment in the Northwest Atlantic region of Newfoundland, with emphasis on northerncod. ICES mar. Sci. Symp. 198: 140-157.

Taylor, D.M., and O’Keefe, P.G. 1999. Assessment of the 1998 Newfoundland and Labrador snowcrab fishery. DFO Can. Stock Assess. Sec. Res. Doc. 99/143.

Templeman, W. 1966. Marine resources of Newfoundland. Fish. Res.Board Can. Bull. 154.Templeman, W. 1968. Review of some aspects of capelin biology in the Canadian area of the

northwest Atlantic. Rapports Proces-verbaux des Reunions. Conseil Permanent Internationalpour l’Exploration de la Mer, 158: 41-53.

Treble, M.A., and Bowering, R. 2002. The Greenland halibut (Reinhardtius hippoglossoides) fishery in NAFO Division 0A. NAFO SCR Doc. 02/46. Serial No. N4658. 9 pp.Vilhjálmsson, H. 1994. The Icelandic capelin stock. Capelin Mallotus villosus (Muller) in the

Iceland, Greenland, Jan Mayen area. Rit Fiskideildar, 13 (1). 281 pp.Vilhjálmsson, H. 1997. Climatic variations and some examples of their effects on the marine

ecology of Icelandic and Greenland waters, in particular during the present century. RitFiskideildar 15: 9-29.

Winters, G.H., and Wheeler, J.P. 1987. Recruitment dynamics of spring-spawning herring in theNorthwest Atlantic. Can. J. Fish. Aquat. Sci. 44 : 882-900.

Worm, B., and Myers, R.A. 2002. Top-down versus bottom-up control in oceanic food webs: ameta-analysis of cod-shrimp interactions in the North Atlantic. Ecology (in press)

Page 101: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–101 DO NOT CIRCULATE OR CITE

12.5 Bering Sea (Fluharty/Radchenko/Wilderbuer)

12.5.1 Introduction

The continental shelves of the eastern and western Bering Sea combine to produce one of theworld’s largest and most productive fishing areas. They contain some of the largest populations ofmarine mammals, birds, crabs and groundfish in the world (Hood 1981). Fully 25% of the totalglobal yield of fish came from the region in the 1970s. The central Bering Sea contains a deepbasin which cleaves the shelves on the Russian and American sides but which remains outside the200 n.m. Exclusive Economic Zones (EEZs) of the two coastal countries. Prior to extendedfishing zones, a fairly complex set of bi- and multilateral fisheries agreements was established forthe area. These range from the agreements on Northern fur seal harvests and Canada/US fisheriesfor Pacific salmon and Pacific halibut to the multilateral International North Pacific FisheriesConvention (INPFC) to develop and use scientific information for management of fisheries on thehigh seas (Miles et al. 1982a, Miles et al. 1982b).24 In the so-called “Donut Hole” a pocket of highseas area surrounded by US and Russian EEZs, scientific research and commercial fishing arecarried out under the Central Bering Sea Agreement between the two coastal states and Japan,Korea and China. The North Pacific Science Organization [PICES] and the North PacificAnadromous Fish Commission (NPAFC) have been established to facilitate fisheries andecosystem research in the North Pacific region, including the Bering Sea.

Commercial fisheries in the Bering Sea are generally large-scale trawl fisheries for groundfish ofwhich about 30% of the total catch is processed at sea and the remainder delivered to shoresideprocessing plants in Russia and the United States. Homebase for many of the Bering Sea vessels isoutside the ACIA arctic region reflecting the comparative advantage of supply and serviceavailable in lower cost regions. Small coastal communities have a strong complement ofindigenous peoples with subsistence fishing interests. They depend on coastal species, especiallysalmon, herring and halibut, however, the overlap with commercial activities is generally small.Anadromous species (marine species which spawn in fresh water) extend far inland via thecomplex river systems and are critical resources for native peoples (see Ch. 9,10,11). The chiefindigenous involvement in the marine commercial sector in the Northeast Pacific is theCommunity Development Program in the Northeast Pacific where 10% of TACs are allocated tocoastal communities and their chosen partners (Ginter 1995).25 Because the eastern Bering Sea iswithin the EEZ of the United States, harvest levels of commercially important species of fish andinvertebrates are regulated through federal laws. Management plans exist for the major targetspecies that specify target fishing mortality levels calculated to maintain the long-term femalespawning stock levels at 40% of the unfished equilibrium level for fully exploited species. In thewestern Bering Sea, i.e., in the Russian EEZ, fishery management is executed on the basis of anannual TAC established for all commercial stocks of fish, invertebrates and marine mammals.Allowable catch is calculated as a percentage of the fishable stock. Percentages for individual

24 Various post-EEZ license agreements have permitted fishing by non-Russian and US fleets. At present such fishingis precluded in the US EEZ and is considerably reduced in waters of the Russian Federation.25 This is worth some 40 million dollars.

Page 102: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–102 DO NOT CIRCULATE OR CITE

stocks and species are based on early scientific studies and do not exhibit annual change. Since1997, these harvest percentages have been revised by Government research institutes, usingapplications of new modeling and adaptive management approaches. The recommended TACsare approved by the special federal agency and issued as a governmental decree.

Annual catches of all commercial groundfish species for the period 1990 through 2001 in the U.S.eastern Bering Sea EEZ ranged from 1.3 - 1.8 million t and averaged 1.6 million ton. The walleyepollock catch averaged 1.2 million t and ranged from 0.99 – 1.45 million t. (Hiatt et al., 2002). Inthe western Bering Sea, the total groundfish catch reached 1.45 million t in 1988 of which walleyepollock contributed 1.29 million t. The annual catch for the period 1990 through 2001 averaged0.73 million t ranging from 0.45 to 1.06 million t. Walleye pollock comprised 89% of the catch,on average, over the 11-year period. Since late 1970s, information on the distribution, abundanceand catch of eastern Bering Sea groundfish species are available from 1) American resourceassessment bottom trawl surveys conducted during the summer annually since 1979, 2)summertime hydroacoustic/midwater trawl surveys conducted triennially since 1979 on the BeringSea shelf, and 3) observer sampling of commercial catches onboard fishing vessels and in fishingports since 1977.

Aquaculture is not a particularly important activity in the Eastern Bering Sea region. The state ofAlaska has generally adopted policies that prohibit aquaculture but does operate some hatcheriesthat produce salmon for release into the sea to supplement times of low escapement. Some ofthese transit the Eastern Bering Sea and may have some effect on larvae of red king crab, etc. butthis has not been demonstrated. None of the hatcheries operate in the Eastern Bering Sea region(NPAFC 2001).

12.5.2 Ecosystem essentials

The Bering Sea is a subpolar sea bounded by the Bering Strait to the north and the AleutianIslands archipelago to the south. Geographically, the Bering Sea lies between 52° and 66° northlatitude, and 162° east and 157° west longitude. The narrow (85 km-long) and shallow (< 42 mdeep) passage of the Bering Strait connects the Bering Sea to the south with the Chukchi Sea andthe Arctic Ocean on the north. The sea area covers almost 3 million km2 and is divided almostequally between a deep basin on the south-western part and a large, extensive continental shelf onthe eastern and northern parts. The eastern continental shelf is 1,200 km in length, exceeds 500km in width at its narrowest point and is the widest continental shelf outside the Arctic Ocean(Coachman 1986). The shelf is characterized as a featureless plain that deepens gradually from itsextensive shoreline to the shelf break at about 170 m depth. Very limited commercial fisheries areprosecuted in the Chukchi Sea or in the Arctic Ocean north of the Bering Straits due to knownlack of resource abundance, operating difficulties and distance from markets. Marine mammalpopulations are locally important for subsistence use.

12.5.3 Fish Stocks and Fisheries

Page 103: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–103 DO NOT CIRCULATE OR CITE

For the purpose of this report, the life history characteristics, distribution and trends of abundance,and fisheries are described for the main species which constitute or have constituted importantfisheries or which are important as forage fish for such species.

Catch records for the major groundfish species of the eastern and western Bering Sea are shownby species in Figures 12.5.1 and 12.5.2 respectively.

(Insert Figure 12.5.1 here)

(Insert Figure 12.5.2 here)

CapelinIn the Bering Sea, adult capelin (Mallotus villosus) are only found near-shore during the monthsurrounding the spawning run. During the other months of the year they are found far offshore.In the eastern Bering Sea, capelin occur in the vicinity of the Pribilof Islands and the continentalshelf break, in the western part - in the northern Anadyr Gulf and near the north-westernKamchatka coast. The seasonal migration may be associated with the advancing and retreatingpolar ice front. In the eastern Bering Sea, winter ice withdraws during the summer months. As acoldwater species, capelin may migrate in close association with the retreating ice edge resultingin the summertime capelin biomass located in the northern Bering Sea, an area not covered bysurveys with very little commercial fishing. Capelin aggregations near the northwesternKamchatka coast exhibit a stable distribution throughout the warm season. It is reported that thebiomass of capelin and smelt grow in the transitional periods when the abundance of othercommon pelagic fish (walleye pollock and herring) are at a low level in the western Bering Sea(Naumenko et al. 1990). Based on integrated expedition data, capelin biomass was estimated at200,000 t on the western Bering Sea shelf in 1986-1990.26 Their biomass may be much larger onthe expanded eastern shelf. Nevertheless, capelin are not commercially exploited in the BeringSea. In Russia, some attempts were undertaken to involve capelin and Arctic cod resources in acommercial fishery in mid-1990s. Capelin have been found to be a major component of the dietsof marine mammals feeding along the ice edge during the winter (Westpestad 1987) and of marinebirds during the springtime.

Greenland halibutIn the Bering Sea, Greenland halibut (Reinhardtius hippoglossoides) spend the first three or fouryears of life on the continental shelf after which they migrate to deep waters of the continentalslope where they live as adults (Alt et al. 1988; Shuntov 1970; Templeman 1973). Althoughresults of tagging studies indicate they undergo feeding and spawning migrations in the NorthAtlantic Ocean, it is unknown to what extent this happens in the Bering Sea. A slow growing andlong-lived species, Greenland halibut reach over 100 cm in length and 20 years of age in theBering Sea. Greenland halibut are highly desired as a commercial product and have been caughtboth in trawling operations and by longlines.27 Beginning in the 1970s the fishery for Greenlandhalibut intensified with catches of this species reaching a peak from 1972 to 1976 of between 63

26 Capelin are caught in summertime bottom trawl surveys in the eastern Bering Sea and integrated trawl surveys inthe western part.27 Catches of Greenland halibut and arrowtooth flounder were not reported separately during the 1960s. Combinedcatches of the two species ranged from 10,000 to 58,000 t. annually and averaged 33,700 t.

Page 104: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–104 DO NOT CIRCULATE OR CITE

000 and 78 000 t. annually, primarily by distant-water trawl fleets from Japan. Catches declinedafter implementation of the Magnuson Fishery Conservation and Management Act in 1977, wherethe U.S. fisheries jurisdiction was extended to 200 miles from the coast, but were still relativelyhigh in 1980-83 with an annual range of 48 000 to 57 000 t. Since 1983, however, trawl harvestdeclined steadily and averaged 8 000 t. from 1989-2000. This overall decline is due mainly tocatch restrictions placed on the fishery because of declining recruitment and market conditions. Inthe western Bering Sea, Greenland halibut were lightly exploited due to low stock abundancebefore the FCMA in the eastern Bering Sea. In 1978 the Greenland halibut fishery began on thenorthwestern continental slope, mostly by longlines. Annual harvest varied from 2 010 – 6 589 t.in 1978 - 1990 with part of the harvest resulting from bycatch in the Pacific cod longline fishery.Since the early 1990s, Greenland halibut stock abundance and catches have declined. Resourceassessment surveys conducted on the continental shelf in 1975 and 1979-2002 indicated thatintermediate size Greenland halibut (40-55 cm) were present throughout the region from 50 – 200m depth during the late 1970s and early 1980s (Alton et al. 1988). By 1985 and 1986 thegeographical distribution shrank so that Greenland halibut were only encountered in the area to thewest and south of St. Matthew Island in much reduced densities. Since that time fish of this sizerange have only been caught in small quantities on the northern part of the survey area. It isunknown whether the conditions of the late 1970s through the early 1980s were favorable forstrong recruitment of Greenland halibut and have since returned to more normal recruitmentlevels, or whether there has been a lack of recruitment to this stock since the mid 1980s. Stockassessment modeling suggests a declining population since 1985 (Ianelli at el. 2001). Greenlandhalibut are widely distributed in the western Bering Sea but are not abundant there. The mostsignificant Greenland halibut aggregations are found on the outer continental shelf and slope alongthe Korjak coast (Novikov 1974; Borets 1997). Survey results indicate that Greenland halibutabundance was higher in the northern part of the Bering Sea in the 1990s compared with the1980s. However, the total biomass and overall distribution of this flatfish decreased in the westernBering Sea as in the eastern region.

ShrimpPandalid shrimp (primarily Pandalus jordani) are well distributed along the outer third of theeastern continental shelf where they are consistently caught in resource assessment trawl catchesin small numbers. Humpy shrimp (P. goniurus) are widely distributed throughout the northernBering Sea shelf and the Anadyr Gulf, in contrast to northern shrimp (P. borealis), which are lessabundant. However, northern shrimp were the first commercially exploited shrimp in the BeringSea when aggregations were discovered on the outer shelf north of the Pribilof Islands in 1960(Ivanov 1970).28 After that the northern shrimp stock sharply declined and commercial fishingceased after 1967. At present, no commercial fishery exists for pandalid shrimp in the easternBering Sea. Humpy shrimp aggregations were discovered in the Anadyr Gulf in 1967. A large-scale Russian trawl fishery harvested humpy shrimp in the northern Bering Sea in late 1960s -1970s until they too became less abundant. Individual trawl catches of pink shrimp reached 10 t.per 15 minute haul in the Anadyr Gulf, which became the catch value record in the world shrimpfishery. Humpy shrimp biomass was estimated at 350 000 t in the Anadyr Gulf in 1975. Theannual Russian harvest of humpy shrimp fishery exceeded 11 200 t in 1978 (Ivanov 2001) after

28 This shrimp fishery was conducted by Japanese vessels and reached a maximum at 31,600 t. in 1963.

Page 105: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–105 DO NOT CIRCULATE OR CITE

which the fishery declined due to the lack of a market for small sized shrimp. Other pandalidshrimp species were also caught as bycatch during the pursuit of other target species

Arctic codArctic cod (Boreogadus saida) are caught in small amounts in resource assessment surveys at thenorthernmost survey stations on the eastern Bering Sea shelf. The southern extent of their summerdistribution is related to bottom water temperature where they have found to range from 59° N in1999 (coldest year) to 62° N in 1996 (warmest survey year on record, excluding 2003). Sincearctic cod are found at such high latitudes, little information is available on their life historycharacteristics in the eastern Bering Sea and they are not pursued as a commercial species due totheir low abundance. In the northwestern Bering Sea and the Chukchi Sea, arctic cod aredistributed at depths from 15 to 251 m (Tuponogov 2001). A local fishery on arctic cod existedthere during years of high abundance (1967-1970; see Tuponogov 2001).

CrabsSnow crab (Chionoecetes opilio) and Tanner crab (C. bairdi) are distributed throughout the easternBering Sea shelf with the exception of the shallow waters of Bristol Bay (Otto 1998). Theabundance of commercial size males of these two species was estimated at 183.5 million crabs in1988 (Stevens et al. 1993). The distribution extends beyond the study area to the north and west,and to a small extent into the Gulf of Alaska. Due to the relatively narrow shelf area of thewestern Bering Sea, snow crab abundance is notably less. In 1969 the number of commercial sizemales was estimated at 25 million crabs (Slizkin and Fedoseev 1989). An intensive directedfishery began for snow crab in the Bering Sea in the 1980s. They were initially caught incidentalto the pursuit of red king crab (Paralithodes camtschaticus) until 1964 when both Japan and Russiaincreased their effort for this species due to a bilateral agreement with the U.S. to limit king crabcatches (Davis 1982). The combined Japanese-Russian catch of snow and Tanner crab increaseduntil 1970 to 22 844 t (ADF&G 2001), after which quotas were established for these nations’fishing fleets and the catch was sharply reduced. The American pot fishery (non trawl) beganshortly thereafter and catches increased during the 1980s to a peak in 1991 at 172 588 t. Catchesrapidly declined with stock decrease but increased again in the mid 1990s as the snow crab stockcondition improved. Since 2000, the stock has again trended downward and the commercialfishery is presently operating under much reduced quotas. The Tanner crab fishery has beenclosed since 1997 in the eastern Bering Sea (NPFMC 2002 SAFE). In the western Bering Sea, thecommercial snow crab and Tanner crab fisheries were not conducted in 2000-2001. Only aninsignificant amount of catch (250 t) was allowed during research surveys, which indicated someimprovement in stock condition allowing a small-scale commercial fishery in 2002.

PollockWalleye pollock (Theragra chalcogramma), hereafter referred to as pollock) is the most abundantspecies within the Bering Sea and is widely distributed throughout the North Pacific Ocean intemperate and subarctic waters (Shuntov et al. 1993; Wolotira et al. 1993). Pollock are asemidemersal schooling fish, which become increasingly demersal with age. They are a relativelyshort lived (natural mortality estimated at 0.3) and fast growing fish, females usually becomesexually mature at age 4 and have a maximum recorded age of around 22 years. The stockstructure of Bering Sea pollock is not well defined. In the U.S. portion of the Bering Sea pollock

Page 106: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–106 DO NOT CIRCULATE OR CITE

are considered to form three stocks for management purposes.29 In the Russian EEZ, pollock areconsidered to form two stocks for management purposes.30 Pollock currently support the largestfishery in U.S. waters and comprise 75-80% of the annual catch in the eastern Bering Sea andAleutian Islands. However, from 1954 to 1963, pollock were only harvested at low levels in theEastern Bering Sea. Directed foreign fisheries first began in 1964 after which catches increasedrapidly during the late 1960s and reached a peak in 1970-75 when they ranged from 1.3 to 1.9million t annually. Following a peak catch of 1.9 million t in 1972, catches were reduced throughbilateral agreements with Japan and Russia. Since the U.S. claim to extended jurisdiction in 1977,the annual average eastern Bering Sea pollock catch has been 1.2 million t. and has ranged from0.9 million t. in 1987 to nearly 1.5 million t. (including the Bogoslof Islands area catch in 1990)while stock biomass has ranged from a low of 4-5 million to highs of 10-12 million t (NPFMC2002 SAFE). In 1980, United States vessels began fishing for pollock and by 1987 they were ableto take 99% of the quota. Since 1988, only U.S. vessels have been operating in this fishery and by1991, the current domestic observer program of this fishery was fully operational. In thesouthwestern Bering Sea, the pollock fishery slowly developed during the mid 1960s andstabilized at 200,000 – 300,000 t in the second half of 1970s to the 1980s. After 1995, there was areduction in harvest due to a decline of the pollock stocks in the western region of the Bering Sea.After that, the total pollock catch in the Russian EEZ was maintained by increasing the fisheryactivity in the Navarin region in 1996 - 1999, which ranged from 596 000 - 753 000 t. The pollockcatch subsequently declined in the northern region due to poor stock condition as well as theapplication of precautionary approaches in the pollock fishery management. The total pollockcatch in the Russian EEZ declined from 1 327 000 t. in 1988 to 393 180 t. in 2000. Vessels of“third countries” began fishing in the mid-1980s in the international zone of the Bering Sea(commonly referred to as the “Donut Hole”). The Donut Hole is entirely contained in the deepwater of the Aleutian Basin and is distinct from the customary areas of pollock fisheries, namelythe continental shelves and slopes. Japanese scientists began reporting the presence of largequantities of pollock in the Aleutian Basin in the mid-to-late 1970's, but large-scale fisheries didnot occur until the mid-1980's. Thus, the Donut Hole catch was only 181 thousand t in 1984, butgrew rapidly and by 1987 exceeded the catch within the U.S. Bering Sea EEZ. The outside ofEEZ catch peaked in 1989 at 1.45 million t and has declined sharply. By 1991, the Donut Holecatch was 80% less than the peak catch, with subsequent low catches in 1992 and 1993. Amoratorium was enforced in 1993 and only trace amounts of pollock have been harvested from theAleutian Basin by resource assessment fisheries.

In response to continuing concerns over the possible impacts groundfish fisheries may have onrebuilding populations of Steller sea lions (Eumetopias jubatus, listed as an endangered speciesafter four decades of decline), changes have been made in the regulations governing the pollock 29 These are: eastern Bering Sea which consists of pollock occurring on the eastern Bering Sea shelf from UnimakPass and to the U.S.-Russia Convention line; Aleutian Islands Region which encompasses the Aleutian Islands shelfregion from 170° W longitude to the U.S.-Russia Convention line; and Central Bering Sea – Bogoslof Island pollock,which are thought be a mixture of pollock that migrate from the U.S. and Russian shelves to the Aleutian Basinaround the time of maturity.30 a western Bering Sea stock centered in the Gulf of Olyutorski, and a northern stock located along the Navarin shelffrom 171° E longitude to the U.S.- Russia Convention line. The northern stock is believed to be a mixture of easternand western Bering Sea pollock with the former predominant (Ianelli et al. 2001). An alternative point of viewregards the northern stock as part of the large eastern Bering Sea pollock population which inhabits the entire easternand north-western Bering Sea shelf (Shuntov et al. 1993; Stepanenko 2001).

Page 107: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–107 DO NOT CIRCULATE OR CITE

fisheries in the eastern Bering Sea and Aleutian Islands. Pollock have been found to be importantprey for Steller sea lions and these changes were designed to reduce the possibility of competitiveinteraction of the fishery with Steller sea lions. For the fisheries, comparisons of seasonal fisherycatch and pollock biomass distributions in the eastern Bering Sea led to the conclusion that thefishery had disproportionately high seasonal harvest rates within critical sea lion habitat whichcould lead to reduced sea lion prey densities. Consequently, management measures were designedto redistribute the fishery both temporally and spatially according to pollock biomassdistributions.31 Three types of measures were implemented in the pollock fisheries: 1) Additionalpollock fishery exclusion zones around sea lion rookery or haulout sites, 2) Phased-in reductionsin the seasonal proportions of TAC that can be taken from critical habitat, and 3) Additionalseasonal TAC releases to disperse the fishery in time. In the Russian EEZ, 30-mile fishery closurezones were established around the sea lion and fur seal rookeries (NPFMC 200x RPA and NRC2003).

Pacific codPacific cod (Gadus macrocephalus) have a widespread distribution from southern California to theBering Sea, although the Bering Sea is the center of greatest abundance for this species. Taggingstudies have shown that they migrate seasonally over large areas. In late winter, Pacific codconverge in large spawning concentrations over relatively small areas.32 Estimates of naturalmortality range from 0.29 – 0.99, while a value of 0.37 is used in the stock assessment model.Pacific cod have been aged as old as 19 years and females are estimated to reach 50% maturity at5.7 years, corresponding to an average length of 67 cm. Pacific cod are the second largest BeringSea groundfish fishery. Beginning in 1964, the Japanese trawl fishery for walleye pollockexpanded and cod became an important bycatch species and an occasional target species duringpollock operations.33By 1977, foreign catches of Pacific cod had consistently been in the 30 000-70 000 t range for a full decade (Thompson and Dorn 2001). The foreign and joint venture sectorsdominated catches through 1988, when a U.S. domestic trawl fishery and several joint venturefisheries began operations. By 1989 the domestic sector was dominant and by 1991 the foreignand joint venture operations had been displaced entirely. Catches of Pacific cod since 1978 haveranged from 33 000 t in 1979 to 232 600 t in 1997 and average 141 900 t. Presently, the Pacificcod stock is exploited by a multiple-gear fishery, including trawl, longline, pot, and jigcomponents.34 Pacific cod were estimated to be at low abundance levels in 1978 but experiencedstrong recruitment (age 3) during the early part of the 1980s, which built the stock to high levelsduring the 1980s. The population biomass is estimated to have peaked at 2.5 million t in 1987 andhas gradually declined to about half of the peak value in 2001. In the western Bering Sea, theRussian cod fishery slowly developed and was mostly unsuccessful until the late 1960s. Severalattempts were undertaken by Japanese and local fishermen in longline and trawl fisheriesdevelopment in the 1920s - 1930s. Meanwhile, commercially significant Pacific codconcentrations were described by the scientific expeditions. In particular, dense aggregations werefound in the northwestern area near the Navarin Cape in 1950-1952 (Gordeev 1954). These results

31 The underlying assumption in this approach was that the independently derived area-wide and annual exploitationrate for pollock would not reduce local prey densities for sea lions.32 Spawning takes place over a wide depth range (40 - 290 m), near the bottom. Eggs are demersal and adhesive.33 In the early 1960s, a Japanese longline fishery harvested Bering Sea Pacific cod for the frozen fish market.34 With the exception of 1992, the trawl catch was the largest component of the fishery (in terms of catch weight) from1978-1996. Since 1997, the longline fleet catches the largest proportion of Pacific cod.

Page 108: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–108 DO NOT CIRCULATE OR CITE

led to the organization of a special cod fishery expedition in 1968 (Vinnikov 1996). Pacific codharvest ranged from 6 500 – 24 500 t there in first years, and peaked at 117 650 t in 1986.However, in the 1990s catches declined due to a restructuring of the fishery and, in recent years,due to decreases in cod abundance in the North Pacific. Pacific cod biomass was estimated at 766000 t in 1989 (Vinnikov 1996) and has declined to 172 000 t in 2000.

Yellowfin soleYellowfin sole (Limanda aspera) is distributed from British Columbia to the Chukchi Sea and intothe western Bering Sea and south along the Asian coast to about 35° N off the South Korean coast(Hart 1973). In the Bering Sea, it is the most abundant flatfish species and is the target of thelargest flatfish fishery in the United States. While also found in the Aleutian Islands and Gulf ofAlaska, the center of abundance for this stock is on the eastern Bering Sea shelf. Adults arebenthic and occupy separate winter and spring/summer spawning and feeding grounds. Theyoverwinter near the shelf-slope break at approximately 200 m and move into nearshore spawningareas as the shelf ice recedes (Nichol 1997). Spawning is protracted and variable, beginning asearly as May and continuing through August, occurring primarily in shallow water at depths lessthan 30 m (Wilderbuer et al. 1992).35 The natural mortality rate likely falls within the range of 0.12to 0.16, with a maximum recorded age of 33 years (Wilderbuer 1997). Yellowfin sole haveannually been caught with bottom trawls on the Bering Sea shelf since the fishery began in 1954.36

In the early 1980s, catches increased reaching a recent peak of over 227 000 t in 1985. During the1980s, there was also a major transition in the characteristics of the fishery in the eastern BeringSea. Prior to this time, yellowfin sole were taken exclusively by non-U.S. fisheries and thesefisheries continued to dominate through 1984. However, U.S. fisheries developed rapidly duringthe 1980s, and foreign fisheries were phased out. Since 1990, only domestic harvesting andprocessing has occurred. The 1997 catch of 181 389 t was the largest since the fishery becamecompletely domestic, but decreased to 101 201 t in 1998. The 2000 catch totaled 83 850 t and the2001 catch was 63 400 t. For many years in the 1990s the yellowfin sole fishery was constrainedby closures in order to attain the bycatch limit of Pacific halibut allowed in the yellowfin soledirected fishery. An age structured stock assessment model is used to determine stock abundancefor the annual stock assessments used to mange the resource. Over the past fifteen years, stockbiomass has declined by 1 million t from the peak biomass observed in 1985 and was estimated at1.6 million t in 2002.37

FlatfishThe flatfish harvest and resource is much smaller in the southwestern Bering Sea with itsrelatively narrow shelf. The harvest is conducted by a small-scale land-based fishery using Danish

35 Eggs, larvae and juveniles are pelagic and usually are found in shallow areas (Nichol 1994). The estimated age at50 percent maturity is 10.5 years at a length of approximately 29 cm (Nichol 1994).36 This species was overexploited by Japanese and Russian trawl fisheries in 1959-62 when catches averaged 404,000t annually. As a result of this harvest level the stock abundance declined. Similarly, catches declined to an annualaverage of 117,800 t from 1963-71 and further declined to an annual average of 50,700 t from 1972-77. The loweryield in this latter period was partially due to the discontinuation of the Russian fishery.37 Model results indicate that yellowfin sole total biomass (age 2+) was at low levels during most of the 1960s andearly 1970s (600,000-800,000 t) after a period of high exploitation. Sustained, above average recruitment from 1967-76 combined with light exploitation resulted in a biomass increase to nearly 2.5 million t by 1985. The populationbiomass has since been in a slow decline as the strong 1981 and 1983 year-classes have passed through the populationwith only the 1991 year class at levels observed during the 1970s.

Page 109: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–109 DO NOT CIRCULATE OR CITE

seines, and, to a lesser extent, trawls. In the northwestern region the flatfish fishery did not existuntil the 1990s except for some flatfish bycatch in the pollock and Pacific cod fisheries. In 1965-1984, the flatfish harvest ranged from 2 440 to 29 140 t and averaged 12 700 t in the northwesternBering Sea. It increased to 33 460 – 39 900 t in 1985-1986, leveled-off at 24000 – 29 000 t duringthe next six years, and then declined to an average annual value of 9 700 t since 1993. Yellowfinsole comprised the main part of the flatfish harvest in the southwestern Bering Sea (72,7% in thepredicted flatfish TAC for 2002) and its biomass is estimated at 78 000 t on the southwesternshelf. Among other flatfish species, Alaska plaice (Pleuronectes quadrituberculatus), rock sole(Lepidopsetta bilineata), and northern flathead sole (Hippoglossoides robustus) are the mostimportant.

Marine MammalsThe Bering Sea contains a rich and diverse assemblage of marine mammals, including northtemperate, Arctic and sub-Arctic species. Twentysix species from the orders Pinnipedia (sea lions,walrus and seals), Cetacea (whales, dolphins and porpoises) Carnivora (sea otter) and polar bearsare present for varying times during the year (Lowry and Frost 1985). Some species are residentthroughout the year38 and others migrate into the Bering Sea during summer months on feedingexcursions. Arctic species including polar bears, walruses, ringed and bearded seals and bowheadwhales are found in the Bering Sea mostly during fall and winter and are associated with thepresence of seasonal sea ice. The majority of the marine mammals species are found over thecontinental shelf and in coastal areas, although five whale species reside in the deep/oceanicwaters of the Bering Sea basin (Lowry et al. 1982).

Harvesting of marine mammals has occurred since at least 1790, the first year when northern furseal harvests were recorded (Lander 1980). The harvest peaked in the 1870s at over 100 000animals and was at levels exceeding 40 000 males annually until 1985 when the northern fur sealcommercial harvest was stopped and only subsistence hunting by native Aleuts was allowed in thePribilof Islands. In the Russian EEZ, fur seal hunting has had many changes since the mid 1980s.Since 1987, the experimental hunting of “silver” fur seals (aged 3-4 months) has been conductedon the Commander Islands (Boltnev 1996). The harvest rate was established at 60% from thebelow-the-year male abundance in 1987-1989. Actually, significantly less than 50% were killed,which has further decreased to less than 30% since 1989. The number of animals killed decreasedfrom 6 700 in 1995 to 3 000 in 1999 and 2 180 in 2000. The declining harvest is related to thedecline of the fur seal population and the negative effect of the hunting disturbance on sealreproduction. All fur seal hunting is presently restricted to Bering Island. Bachelor males agedfrom 2 to 5 years were hunted on Medny Island until the mid-1990s (2 134 animals were killed in1994) until it was closed to harvesting in 1995. Whaling spread to the Bering Sea in the mid 19th

Century when large numbers (2 500 in 1853) of bowhead whales were taken (NMFS 1999). Thisharvest continued for 50 years until the bowhead whale population became depleted. The currentsubsistence harvest totals 60-70 whales annually. Some species, such as humpback and greywhales, which are present in the Bering Sea in the summer, were historically harvested during thewinter near Hawaii and California and in waters off the Chukotski Peninsula (about 130-135whales). Kenyon (1962) reported that Steller sea lions were very abundant in the Pribilof Islandswhen they were discovered in 1786, but were overhunted soon thereafter. After some protectivemeasures were taken, population numbers grew from a few hundred in 1914 to about 6 000 in 38 e.g., harbor seal, Steller sea lion, sea otter, beluga whales and Dall’s porpoise.

Page 110: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–110 DO NOT CIRCULATE OR CITE

1960. The population has since declined to low numbers and has been the subject of extensiveresearch to find the cause of the decline.

In the U.S., stock assessment information on the 39 stocks of the 24 species of marine mammals inthe Bering Sea are used to classify each stock as either strategic, non-strategic or not available(Angliss et al. 2001). Strategic stocks are those, considered either threatened, endangered ordepleted under U. S. law. The “strategic” stocks include: northern fur seal, sperm whale,humpback, fin, the North Pacific right whale and the bowhead whale. Three Bering Sea stocksalso have further designations: northern fur seals are designated as depleted under the MarineMammal Protection Act, the western stock of the Steller sea lion is listed as endangered under theEndangered Species Act (ESA), and the bowhead whale is also listed as endangered under theESA. Nine of the 39 marine mammal stocks are estimated to be increasing, five are stable, threeare declining and the status of the others is unknown. Subsistence harvest is allowed for threespecies; northern fur seals, beluga and bowhead whales. In Russia, marine mammal populationsare classified as commercial, non-commercial or protected. Protected species include all whalesand dolphins (with the exception of grey whales, whaled by indigenous people for subsistence),sea otter and polar bears. Some commercial quota is established for beluga whales, but is nottaken. Walrus, spotted, ringed and ribbon seal are the objects of hunting in the northwesternBering Sea. However, their harvest is relatively low in recent years after the cessation of ship-based hunting operations. In 1998-2000, the harvest was less than 60% of the established TAC ondifferent seal species and averaged 32.8%.

SalmonThe Bering Sea is important habitat for many stocks comprising the five species of Pacific salmonduring the ocean phase of their life history. Here, the diverse stocks intermingle from origins inSiberia, Alaska, the Aleutian Islands, Japan, Canada and the U.S. west coast. The earliest fisheriesfor salmon were likely native subsistence fisheries in which salmon were captured returning totheir native streams to spawn. During the 20th Century there were three main fisheries for salmonin the Bering Sea: 1) the Russian and Alaskan domestic fisheries, 2) the Japanese high-seas gillnetand longline fishery and 3) the bycatch of salmon in the groundfish fisheries.

Salmon canneries first appeared on the Alaska side of the Bering Sea in the late 1890s to processfish returning to Bristol Bay. It is reported that between 1894 and 1917 the Kvichak andNushagak rivers flowing into Bristol Bay produced 10 million sockeye salmon annually (Netboy1973). Purse seines and gill nets were the primary fishing gear used in the early days of thefishery. Gill nets were hauled from the beach using horses, which were later replaced by engines,whereas the purse seine fishery started around the time of World War I with the advent of poweredfishing craft. Purse seining continues to the present as the primary gear in a highly mobile fleetfishing near-shore, which assures the targeting of specific salmon stocks. Although all five speciesof Pacific salmon are present in Bristol Bay, sockeye salmon are by far the most abundant andhave dominated the salmon catch for years. The Bristol Bay salmon catch for all species totaled 42million fish in 1993, of which 41 million were sockeye salmon, the largest catch on record.39 Onaverage, pink salmon contributed 73.8% of the Russian salmon catch in the western Bering Seafrom 1952-1993, chum salmon 24.2%, sockeye salmon 1.3%, Chinook 0.6% and coho only 0.1% 39 Fishery statistics from the Pacific salmon fishery on the western Bering Sea coast (eastern Kamchatka region) areavailable since 1906.

Page 111: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–111 DO NOT CIRCULATE OR CITE

(Chigirinsky 1994). Since 1989, the runs of pink salmon to the eastern Kamchatka coast havebeen in good condition during odd years. The historical high catch totaled 83 640 t in 1999. Theaverage pink salmon catch (38 390 t) for 1989-2001 is more than twice the level given by A.Chigirinsky (1994) for 1952-1993 (15 996 t). Similarly, chum salmon catches were also stable at11 000 - 12 000 t in 2000-2001 compared to 5 250 t for 1952 through 1993. The recent improvedstock conditions coincide with new fishery regulations, which limit the chum salmon bycatchduring the pink salmon fishery. The main sockeye salmon fishery in eastern Kamchatka resultsfrom the productive Kamchatka River, slightly south of the Bering Sea.

The Japanese highseas gillnet and longline salmon fishery expanded into the Bering Sea in 1952with three motherships and 57 catcher boats, which grew to 14 motherships and 407 catcher boatsby 1956 (Netboy 1973). The peak catch of 116 200 t occurred in 1955 and catches ranged from 71000 - 87 000 t for the 21 year period 1957 – 1977 (Harris 1989). Sockeye, chum and pink salmoncomprised 95% of the catch in this fishery, which ceased operations in 1983. The floating gillnetswere typically 50 m deep and assembled in separate sets ranging from 6-9 nautical miles in length.Mesh sizes were 13 cm and the gear was set at twilight and retrieved after 1 a.m. The bycatch ofsalmon in the commercial groundfish fisheries is of less importance than the directed fisheriesmentioned above, but still accounts for fishing mortality important to resource managers. Observersampling of the groundfish fishery indicates that Chinook salmon are more frequent in the bottomtrawl catch and the other salmon species are more frequent in the pelagic trawls (Queirolo et al.1995). In the western Bering Sea, primarily Chinook and chum salmon were present in the bottomtrawl catches during research surveys in 1974-1991 (Radchenko and Glebov 1998).

12.5.4 Past climatic variations and their impact on commercial stocks

Climate change primarily influences ocean water temperatures through the regulation of synopticprocesses and the water exchange between the Bering Sea and the Pacific Ocean. The directeffects of atmospheric forcing resulting from climate variations are very important to the physicaloceanographic dynamics of the eastern Bering Sea shelf, which has a characteristically sluggishmean flow and is separated from any direct oceanographic connection to the North Pacific Oceanby the Alaska Peninsula. Therefore, linkages between the eastern Bering Sea shelf and the climatesystem are primarily a result of the ocean-atmosphere interaction (Stabeno et al. 2001). Climatevariations in this region are directly linked to the location and intensity of the Aleutian Lowpressure center which impacts winds, surface heat fluxes and the formation of sea ice (Hollowedand Wooster 1995). The pressure index has experienced eight statistically significant shifts,alternating between cool and warm periods, during the past century, which has occurred onroughly decadal time scales (Overland et al. 1998). A well-documented shift (Trenberth 1990 andothers) from a cool to a warm period occurred in 1977-1989, which coincided with thecommencement of fishery-independent sampling programs and fishery catch monitoring of majorgroundfish species. Information from the contrast between this period and the prior andsubsequent cool periods (1960-1976 and 1989-2000) form the basis of the following discussion ofthe response of eastern Bering Sea species to these climate induced system changes (Figure12.5.3).

(Insert Figure 12.5.3 here)

Page 112: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–112 DO NOT CIRCULATE OR CITE

The influx of Pacific waters northward into the western Bering Sea results in a warming effect.The dynamics of the environmental conditions of the Bering Sea offshore zone and the relativelynarrow shelf in the west are largely determined by the periodic behavior of current patterns(Shuntov and Radchenko 1999). The direction and velocity of these currents coincide withchanges in the atmospheric circulation pattern, effects which are manifested through the change inintensity of the inflow of North Pacific Ocean water. From 1977-1989, a period of enhancedatmospheric transport, an intensification of currents into the Bering Sea resulted in enhancedfluctuations in the thermal properties of the system. During these years, the effect of horizontalwater movement and mixing on primary production was almost as important as vertical mixingdue to the renewed supply of nutrients necessary for phytoplankton blooms. According to long-term data series, the highest concentrations of spring-time nutrients in the upper mixed layer wereobserved in the Aleutian straits, over the continental slope, and in areas where the influx of NorthPacific water is present. The enhanced rate of primary production may reach rates of up to 10-13gC/m2 per day (Sapozhnikov et al. 1993), which cannot be totally utilized by the zooplankton andmicroheterotrophs (especially in the western Bering Sea shelf). The unutilized primary productionaccumulates at the upper boundary of subsurface waters, which is relatively cold formicroheterotrophs, and the organic matter gradually rises into the upper layers in divergencezones and cyclonic eddies during the warm season. Therefore, favorable conditions for planktondevelopment during spring months, both from heating and nutrient supply from Pacific waters,may cascade through higher trophic levels and play a large role in determining the total biologicalproductivity throughout the whole year (Radchenko et al. 2001).

Four physical processes primarily determine the change in ocean climate regimes in the NorthPacific (Schumacher 1999). They are: the lunar tide cycle, variations in solar radiation (Davydon1972, Van Loon and Shea 1999), changes in the North Pacific circulation that affect air-seaexchange of heat, and momentum with the Aleutian Low atmospheric pressure pattern.Undoubtedly, these processes generate a subset of basin scale factors, each of which contributes tothe formation of the Bering Sea oceanographical conditions. An example of an atmosphereactivity center in the Northern Hemisphere is the Aleutian Low (Beamish and Bouillon 1993;Latif and Barnet 1994; Wooster and Hollowed 1995; Francis et al. 1998; Luchin et al. 2000). Thewater inflow and atmospheric forcing apparently serve as links in the signal transfer chain for theBering Sea region. Their functioning reflects both the direct effect of the atmosphere on themarine environment through the temperature regime of shelf waters, and the undirectedoceanological phenomena in the offshore zone. The signal then propagates through changes in thegeneral current pattern and tidal wave parameters, which determine the intensity of the waterexchange between the shelf and open sea regions.

Changes in atmospheric climate are primarily transmitted through the Bering Sea physicalenvironment to the biota through the mechanisms of wind stress (Francis et al. 1988) and theannual variation in ice extent (Stabeno et al. 2001, Niebauer et al. 1999). These mechanismsdirectly alter the timing and abundance of primary and secondary production by changes in thesalinity, mixed-layer depth, upwelling, nutrient supply and vertical mixing in the ocean system.These changes in marine environment were found to vary on a decadal scale and resulted in higherprimary and secondary productivity during the warm period of 1977-88 relative to the earlier coolperiod (Polovina et al. 1995, Brodeur and Ware 1992, Sugimoto and Tadokoro 1997, Luchin et al.

Page 113: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–113 DO NOT CIRCULATE OR CITE

2000, Hollowed et al. 2001). During periods when low summertime storm activity persists for theBering Sea, as in 1993-98, water column stratification was found to increase. As a result, heatingof a thin surface layer above the seasonal thermocline prevented vertical nutrient transport fromthe underlying, stratified layers, which reduced the primary production and biological productivityof the Bering Sea (Shuntov et al. 1997) despite warmer surface water temperature. This isconsistent with the total heat budget of the upper layer of the Bering Sea, which was lower during1993-98 than during the warmer period of the previous decade (Figure 12.5.4.).

(Insert Figure 12.5.4 here)

In the relatively warm years of 1997-1998, significant growth occurred in euphausiid biomass inthe western Bering Sea (Radchenko et al. 2001) suggesting that warmer waters provide favorableconditions for the survival and growth some sub-Arctic zooplankton species. Crustaceans growthrate has also been found to be above average in warm conditions (Zaika 1983, Vinogradov &Shushkina 1987). This enhanced growth rate allows for a longer maturation period and spawningseason. A meta-analysis of marine copepod species indicates that growth rate is positivelycorrelated with increasing temperature and generation time decreases allowing more productivityin warmer climates (Huntley and Lopez 1992). The oceanographic conditions in the epipelagiclayer are not considered crucial for copepod reproduction in the Bering Sea, since copepod speciesreproduce in relatively stable deeper layers below 500 m. However, calanoid copepod biomasswas much higher in the eastern Bering Sea middle shelf during warm years (Smith and Vidal1986), likely due to higher growth rates. These findings suggest that climate change to a warmperiod enhances ecosystem productivity from the lower trophic levels, particularly for planktoniccrustaceans.

If warming resulting from the “greenhouse effect” evenly envelops all climatic zones of the Earth,and average seasonal temperature differences between the tropical, temperate and Arctic latitudesdo not change substantially, there is reason to expect that the scale and the periodicity of theclimatic changes, which determine wind stress on the underlying Bering Sea surface, and also thegeneral type of wind transport, will be preserved. In this case, the period of meridional typepredominance in the wind transport above the Bering Sea, beginning in the early 1990s, may beprolonged for 10-12 years and then subsequently change to a period of enhanced zonal transport.During the warm 1980s the zonal pattern of atmospheric circulation predominated (Luchin et al.1998) as was the case in the 1920s-1930s (Shuntov and Vasilkov 1982). Periods of decreasedzonal atmospheric circulation index (Girs 1974) are characterized by colder Arctic air masses overthe Bering Sea region and a decrease in air temperature. It is noteworthy that the transitional 1989-1990 years were also characterized by a decrease in the zonal atmospheric circulation patternabove the far-eastern seas (Glebova 2001).

Ice distribution and residence time are frequently regarded as integral with the thermal regime ofthe Bering Sea pelagic zone (Overland 1991, Khen 1997, Niebauer et al. 1999, Wyllie-Echeverriaand Ohtani 1999, Luchin et al. 2000). The dynamics of sea ice conditions directly depend on theintensity of the shelf water cooling in winter, wind direction and water exchange between the shelfand the open sea. Similarly, ice conditions determine the intensity and the degree of winterconvection, formation of cold near-bottom shelf waters, and the temperature distribution ofsurface and intermediate layers of the Bering Sea. The extent and timing of the sea ice also

Page 114: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–114 DO NOT CIRCULATE OR CITE

determines the area where cold bottom water temperatures will persist throughout the followingspring and summer. This area of cold water, known as the “cold pool”, varies with the annualextent and duration of the ice pack, and can influence fish distributions. Adult pollock have showna preference for warmer water and exhibit an avoidance of the cold pool (Wyllie-Echeverria 1995)such that in colder years they utilize a smaller portion of the shelf waters and in warm years havebeen observed as far north as the Bering Strait and the Chukchi Sea.

Pollock larvae concentrate in the water mass under the seasonal thermocline (Nishiyama et al.1986). More productive year classes of pollock coincide with better nursery conditions of theirlarvae, which were related to a well developed thermocline (picnocline), large biomass of copepodnauplii, and low abundance of predators (Bailey et al. 1986; Nishiyama et al. 1986; Shuntov et al.1993). As mentioned above, the first two factors are related to the warm conditions of the BeringSea epipelagic layer. Age-1 pollock may also be distributed throughout the cold pool area andmove between water masses. During cold conditions, predation pressure on age-1 pollock isintense by their major piscine predators (adult pollock, arrowtooth flounder and Pacific cod). Asthe cold pool was reduced, predation on age-1 pollock increased due to overlapping distributionsof Greenland turbot, yellow Irish lords and thorny sculpins (Wyllie-Echeverria and Ohtani 1999).The total biomass of the first group of predators was much higher in the 1980s than the secondgroup (Audin et al. 2002) and has remained higher until the present time despite some declines inwestern Bering Sea walleye pollock and cod stocks. In addition, the second group of predators iscomprised of relatively small-sized fish (except Greenland turbot) and age-1 pollock could avoidpredation through higher growth rates during warm conditions. In the relatively warm 1980s,strong year-classes of pollock were found to occur synchronously throughout the Bering Sea(Bulatov 1995) and coincide with above-normal air and bottom temperatures and reduced icecover (Quinn and Niebauer 1995, Decker et al. 1995). These favorable years of production are theresult of high juvenile survival and have been shown to be related to how much cold water habitatis present (Ohtani and Azumaya 1995), the distribution of juveniles relative to the adult populationto avoid predation (Westpestad et al. 2000) and enhanced rates of embryonic development inwarmer water (Haynes and Ingell 1983). Strong year-classes of pollock were also observed in theeastern Bering Sea in the 1990s (Stepanenko 2001), which may be related to the higher frequencyof El Nino events, which contributed to heat transport throughout the region (Hollowed et al.2001). However, there were no strong year classes of walleye pollock in the western Bering Seain the 1990s. This could be explained by a general cooling of the Bering Sea climate and theoceanographic regime during a period of less intensive Pacific water inflow in the 1990s. Thepelagic layer heat budget may need to be on the level of the late 1970s-1980s for the pollockreproduction conditions to improve in the Bering Sea on the whole. The average location andintensity of the winter position of the Aleutian Low pressure field has been linked to the dynamicsof the annual ice conditions (Glebova 2001). Examination of average 10-day period maps of sealevel pressure reveals that six distinct atmospheric patterns characterize the Bering Sea. Glebova(2001) found that 3 of these types were associated with winter, with the stated “4th type” mostfrequently present. This 4th type results in a decrease in air temperature through the transfer ofcold Arctic air into the Bering Sea through the prevalence of persistent northeast winds. Therelationship between the number of 10-day periods characterized by the 4th type of atmosphericcirculation pattern and the percent of ice cover for 1984-2000 is shown in Figure 12.5.5.

Page 115: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–115 DO NOT CIRCULATE OR CITE

The position of the boundary between the western (with the well defined cold interlayer; CIL) andeastern (with weak CIL) water masses of sub-Arctic structure likely affects the coupling of theprocesses of ice formation in the Okhotsk and Bering Sea (Plotnikov and Yurasov 1994, Plotnikov1996). When the boundary is displaced far to the east, ice conditions in both seas developed thesame trend, whereas under a westward shift, the Bering Sea is characterized by an eastern varietyof ice propagation, and disparate trends develop. The location of the “boundary” development isrelated to the state of the Alaska Gyre and the Western Subarctic Gyre, which, in turn, depend onatmospheric processes. A notable period of high distinction between the trend of ice conditions inthe Sea of Okhotsk and the Bering Sea was observed in 1985-1991 (Shuntov 2001) when a large-scale reconstruction of the water circulation pattern in the North Pacific led to a strengthening ofthe western sub-Arctic gyre and a weakened Alaska Gyre (Radchenko 1994, Hollowed andWooster 1992). Since 1992, winters of reduced ice formation have been observed in both the Seaof Okhotsk and the Bering Sea. Increased differences in east-west ice formation were not observedin the second half of 1990s, even after 1998, when a general increase in ice conditions occurred(Shuntov et al. 2002). However, low ice cover in the Bering Sea (Figure 12.5.5) and heavy ice inthe Sea of Okhotsk were noted in the first years of the 21st century. These observations maysignal the beginning of a sequential period of regionally distinct winter ice conditions in both searegions.

(Insert Figure 12.5.5 here)

During warm periods, favorable environmental conditions after the seasonal ice retreat, which cansignificantly affect an increase in the Bering Sea biological productivity. In contrast, physicalfactors during cold periods adversely affect an increase in the zooplankton growth and biomass,and the viability of the pelagic fish juveniles feeding on it. The “Oscillating Control Hypothesis”predicts that the southeastern Bering Sea pelagic ecosystem function alternates between primarilybottom-up control in cold regimes and primarily top-down control in warm regimes (Hunt et al.2002). Late ice retreat (late March or later) leads to an early, ice-associated bloom in cold water(e.g., 1995, 1997, 1999), whereas no ice, or early ice retreat before mid-March, leads to an open-water bloom in May or June in warm water (e.g., 1996, 1998, 2000). Zooplankton, in particularcrustacean populations, are sensitive to water temperature. In years when the spring bloom occursin cold water, low temperatures limit the production of zooplankton, the survival of larval andjuvenile fish, and their recruitment. Such phenomenon may be important for large piscivorousfish, such as walleye pollock, Pacific cod, and arrowtooth flounder. When continued over decadalscales, this leads to bottom-up limitation and a decreased biomass of piscivorous fish.Alternatively, in periods when the bloom occurs in warm water, zooplankton populations shouldgrow rapidly, providing plentiful prey for larval and juvenile fish. In the southeastern Bering Sea,important changes in the biota since the mid-1970s include a marked increase in the biomass oflarge piscivorous fish and a concurrent decline (due to predation) in the biomass of forage fish,including age-1 walleye pollock, particularly over the southern portion of the shelf (Hunt et al.2002).

Examination of the spatial distributions of forage fishes including herring, capelin, eulachon andjuvenile cod and pollock indicate temperature related differences (Brodeur et al. 1999, Wyllie-Echeverria and Ohtani 1999). Annual capelin distributions exhibit an expanded range in yearswith a larger cold pool and contract in years of reduced ice cover. Although the productivity of

Page 116: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–116 DO NOT CIRCULATE OR CITE

capelin stocks in relation to temperature is not known, population growth of this relatively cold-water dwelling fish is not expected under the conditions of a warm regime. As discussed above,capelin biomass grew when the abundance of walleye pollock and Pacific herring were at a lowlevel in the western Bering Sea (Naumenko et al. 1990) possibly due to a reduction in predationpressure of these species on capelin larvae. The eastern Bering Sea herring stocks exhibitedimproved recruitment during warm years (Williams and Quinn 2000), similar to herring stocksthroughout the Pacific coast of the American continent where the timing of spawning has alsobeen shown to be temperature related (Zebdi and Collie 1995). In the western Bering Sea, Pacificherring have also demonstrated a dependence on reproductive success related to the thermalconditions of coastal waters. However, herring stock increase and large-scale fishery restorationare related to the “historically most abundant” (Naumenko 2001) year class which appeared in theanomalously cold 1993. Generally, strong herring year classes have appeared in the westernBering Sea in years with high SST in May but the lowest SST in June (Naumenko 2001). After2000, herring biomass decreased in the western Bering Sea but still exceed the average levelinherent for the last warm period (1977-1989). In general, the distributions of all forage speciesfrom trawl surveys during a cold year (1986) were found to be more widespread with greateroverlap among species than during a warm year (1987) (Brodeur et al. 1999).

Time-series of recruitment and stock biomass have been examined for evidence that climatic shiftsinduce responses in the production of groundfish species in the Bering Sea and North PacificOcean (Hollowed and Wooster 1995; Hollowed et al. 2001). Even though results from thesestudies can be highly variable, strong autocorrelation in recruitment, associated with thesignificant change in climate in 1977, was observed for salmonids and some winter-spawningflatfish species. Substantial increases in the abundance of Pacific cod, skates, flatfish and non-crabbenthic invertebrates also took place on the Bering Sea shelf in the 1980s as discerned from trawlsurvey CPUE (Conners et al. 2002). This warm period was characterized by larger researchcatches and a change in the benthic invertebrate species composition from a system largelydominated by crabs to a more diverse mix of starfish, ascidians and sponges.

In the southwestern Bering Sea, transition from the relatively warm period of 1977-1989 to thesubsequent cool period also showed a notable response in the groundfish community. Theproportion of Pacific cod decreased from 80% in 1985 to 12% - 26.3% in 1990s while sculpin(8.2% in 1985) and flatfish (9.3% in 1985) proportions increased by 15.1%-31.5% and 24.2-39.6%, respectively (Gavrilov & Glebov 2002). Anthropogenic factors can also notably affect thebenthic communities’ state and dynamics. It should be recognized that large fishery removals ofred king crab occurred in the 1970s, which may have contributed to this reorganization of thebenthos in the eastern Bering Sea. The climatic change related to recruitment success for winter-spawning flatfish may be associated with cross-shelf advection of larvae to favorable nurseryareas, instead of water temperature (Wilderbuer et al. 2002).

Ice conditions and water temperatures can influence not only fish stock distributions andabundance but also fishery effectiveness. Coldwater effects have been observed in the behavior offlatfish species which may cause changes in the annual operation of the fishery. Because coldwater causes slower metabolism in high latitude fish stocks, it has been observed that spawningmigrations of yellowfin sole may be delayed in cooler years (Wilderbuer and Nichol 2001), whichcan alter the temporal and spatial characteristics of the fishery. In addition, high catch rates have

Page 117: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–117 DO NOT CIRCULATE OR CITE

been obtained by targeting yellowfin sole in proximity of the retreating ice edge, which has a hightemporal variability, and in warm years only occurs in areas north of their spring-time distribution.The catch process can also be affected as it is believed that flatfish bury themselves in muddysubstrate during cold years whereby they are less vulnerable to herding by the sweep lines ofbottom trawls (Somerton and Munro 2001). This would result in lower catch rates during coldyears for shelf flatfish species. These temperature related behavior effects may also occur forother commercial species, in particular pelagic fish, which react to avoid capture (Sogard and Olla1998).

For some commercially important fish stocks, in particular Pacific salmon, similarities in stocktrends driven by the large-scale climate factors were observed both regionally and for the wholeNorth Pacific Ocean (Beamish and Bouillon 1993, Klyashtorin and Sidorenkov 1996, Jaenicke etal. 1998, Finney et al. 2002). However, the variability in the magnitude of salmon runs withinregions and between species and years is very large suggesting that these fluctuations may becaused by environmental conditions operating on individual stocks during the various stages oftheir life cycle, as well as from the abundance of the parental generation. It has been documentedthat increased salmon production due to enhanced ocean survival is a response related to climatechange and varies synchronously, but in opposite modes, between Bering Sea and west coast U.S.and Canada salmon stocks (Mantua et al. 1997, Hare and Francis 1995, Hollowed et al. 2001).Within the Bering Sea and adjacent areas, opposite trends were observed for the East-Kamchatkaand Bristol Bay sockeye salmon but similar stock abundance growth was observed forsoutheastern Alaska and East-Kamchatka pink salmon (odd years) in the1990s. Throughout theircentury long exploitation, Alaska salmon stocks have had periods of high and low productionwhich persist for many consecutive years before abruptly reversing to the opposite productionstate. These production regimes coincide with low frequency climate changes in the North PacificOcean and the sub-arctic Bering Sea (Pacific Decadal Oscillation and the Aleutian Low PressureIndex). In the 1930s and early 1940s, and then again in the late 1970s, Bering Sea salmon catchesreached high levels during warm temperature regimes in their oceanic habitat. It is hypothesizedthat improved feeding conditions may prevail during warm oceanic regimes (Hare and Francis1995). There also is evidence of an upper thermal tolerance for salmon species, which has setlimits on their distributions (Welch et al. 1995), but it is doubtful that this effect would happen inthe Bering Sea because the historical temperature range is much lower there.

The three species of crab, which inhabit the Eastern Bering Sea shelf (red king crab, Tanner craband snow crab), exhibit highly periodic patterns of increased abundance. Rosenkranz et al. (2001)investigated five hypotheses on factors affecting year class strength of Tanner crab in Bristol Bayin order to understand these patterns. They determined that anomalously cold bottomtemperatures may adversely effect the Tanner crab reproductive cycle and that northeast windsmay also promote coastal upwelling, which advects larvae to regions of fine sediments favorablefor survival upon settling. Incze at al. (1987) linked low densities of copepods inside the 70 misobath of Bristol Bay with low abundance of Tanner crab larvae. An examination of recruitmentpatterns of red king crab in relation to decadal shifts in climate indicate that the Bristol Bay stocksare negatively correlated with the deepening of the Aleutian Low and warmer water temperatures(Zheng and Kruse 2000). Red king crab were also moderately exploited during the late 1970s,which contributed to the population decline.

Page 118: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–118 DO NOT CIRCULATE OR CITE

Given our present state of knowledge of complex marine ecosystems such as the Bering Sea, it isnot possible to predict with any certainty what the effects of future atmospheric forcing, in thiscase increased sea surface temperature, may have on commercial fish and invertebrate species.Evaluation of a future state of nature, like global warming, would require knowledge of the futurevalues of many ocean-atmosphere parameters to adequately describe how these changes would bemanifested in upper trophic level commercial stocks. These include: storm activity and frequency,wind direction and intensity, shelf stratification characteristics, effects on circulation and transportactivity, sea level pressure (location and intensity of the Aleutian Low pressure system) andprecipitation as well as the predictions of sea surface temperature. Although consideration of seasurface temperature alone is an oversimplification of the climatic effect on ocean-atmosphereinteraction through global warming, five future climate scenarios have been posited for the BeringSea. Average annual values for these scenarios range from 1.9 to 7.0 °C for 1990-1994 and areprojected to increase over a range from 40% to two and a half times during the next century.These increases range from 2.7 to 3.1°C over the projection and model significant increases inmean annual sea surface temperature. Mean annual sea surface temperature anomalies for the past18 years have ranged from +/- 1°C.

A hypothesis may be entertained about future Bering Sea conditions if we allow ourselves toextrapolate trends which seem to characterize the decadal scale variability of the key physicalfactors which influence the Bering Sea ecosystem. It can then be predicted that the followingphenomena and processes could be expected given the background of future air temperatureincrease and general warming of the active layer of the upper pelagic zone:

- an increase of the zonal or regional type repetition of atmospheric circulation for the early2000s and for the period of the next eleven-year cycle of solar activity;

- an increase of the storm activity of wind-induced turbulence for the corresponding period;- a gradual increase of the water exchange with the Pacific Ocean, the achievement of the

level, close to the maximal one, in 2015-2020;- low ice conditions, intensified by an increase in the air temperature, for the next 10 to 20

years.

As described for the other three sea areas reported in this chapter, available information from therecent warm period in the Bering Sea suggests that primary productivity, and therefore carryingcapacity, would be enhanced in the warming episode. Because mixed layer depth and currentinformation is not available for this scenario, it is unknown to what extent this increase wouldoccur due to uncertainties in the availability of nutrient renewal essential to sustainedphytoplankton and zooplankton blooms. Spring blooms are associated with the ice edge and awarming period with less ice could delay the springtime onset of primary production (Hunt et al.2002). Specific conditions of high water column stability, which are formed at the sea ice edgeduring the spring ice retreat, support intensive spring bloom and photosynthetic activity ofplanktonic algae.

A hypothesis regarding the dependence of the phytoplankton cells floating/sinking process onwater column stability in the thin surface layer must be tested relative to the coccolithophoridbloom phenomenon, which occurred in the late 1990s. According to recent studies onphytoplankton sinking velocity, heavier diatom algae cells sink quicker than flagellates, which arelighter in a weight (Huisman and Sommeijer 2002). It follows then that it is possible that iceless

Page 119: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–119 DO NOT CIRCULATE OR CITE

winters would create unfavorable conditions for diatom algae bloom. Therefore, a forecastedclimate warming can restrain biological production in the Bering Sea until the beginning of theprojected increase in ice cover (after 2010). Alternatively, a warmer period would speed up theprocess of thermal stratification so that the bloom, which is not ice-reliant, would commencesooner. However, an alternate hypothesis postulates that nutrient supply quickly runs out during ashort and intensive bloom and photosynthesis slows down. Pacific water inflow can be anothersource of nitrogen and phosphorus in addition to the upwelling and nutrient recycling in the activepelagic layer. Expected intensification of water exchange with the Pacific Ocean gives a higherexpectation on primary production growth in the Bering Sea ecosystem in the forthcoming warmperiod. Starvation, advection and predation are usually listed as determinates of years class inmarine species so it follows that increased primary productivity is usually associated withimproved survival for juveniles of most species (Cushing 1969) and subsequent contribution tothe adult spawning stock.

12.5.5 Possible influence of future climate warming

To predict the relationship between future climate change and commercial species distribution,abundance and harvest patterns, we must first assume that future management policies will be thesame as the present. That is, target fishing mortality values will be imposed which maintain thefemale spawning stock at a minimum of 40-60% of the unfished level (depending on species).Furthermore, when stocks are assessed at below this level, harvest is reduced proportionally torebuild the spawning stock to the target level. Given this management system we would expectthat fisheries for species which respond favorably to warmer climates would realize larger catchesand may shift to areas of increased abundance or expanded habitat and those species which arenegatively impacted by the change would expect smaller quotas, contracted areas of operation andeven vastly different locales.

From the literature documenting the response of Bering Sea species to the past warm period, itseems reasonable to assume that there will be increases in the abundance of many groundfishspecies under the future climate scenarios. Pollock, Pacific cod, Pacific halibut (Clark and Hare2002), skates, some flatfish species, salmon, eastern Bering Sea herring and Tanner crab all couldbenefit from the warming (Table 12.5.1), although it is not conclusively clear what mechanismsare responsible for the increase in most cases. Strickland and Sibley (1984) propose a possiblenorthward expansion of pollock feeding and spawning habitat under a global warming scenariodue to reduced ice cover and stratification of the water column and increased food supply.

Page 120: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–120 DO NOT CIRCULATE OR CITE

Table 12.5.1Estimated effect of different physical phenomena and processes on the stock conditions in thewestern Bering Sea and forecasted stock dynamics under the expected conditions of late 2000s –2020s.

Group

Watertemperaturewarming in

upperpelagic layer

Increase ofwindstress,zone

transportrepetition

Growth ofwater

exchangewith Pacific

ocean

Mild icecondi-tions

Prevalent biologicaleffects related to thephysical environment

changes Key reference:1 Adult Pollock + 0 + + + (food supply) Shuntov et al. 19932 Juvenile Pollock + - + + - (predation) Nishiyama et al. 19863 Pacific cod + 0 + 0 + (food supply) Bakkala 19934 Pacific herring, WBS - 0 - + - (competition) Naumenko 20025 Pacific herring, EBS + 0 - + - (competition) Wespestad 19876 Pacific salmon + - + 0 + (food supply) Hollowed et al. 20017 Cephalopods - 0 + 0 - (predation) Sinclair et al. 19998 Capelin - - - - - (predation) Wespestad 19879 Arctic cod - + - - - (competition) Wyllie-Echeverria &

Ohtani 199910 Pacific halibut - 0 0 0 - Clark et al. 199911 Greenland turbot - 0 - 0 - (competition) Livingston et al. 199712 Arrowtooth flounder 0 + 0 0 + (food supply) Wilderbuer et al. 200213 Small flatfish + + 0 - + (food supply) Wilderbuer et al. 200214 Skates 0 0 0 0 + (food supply) Borets 199715 Sculpins + 0 0 - - (competition) Borets 199716 Atka mackerel + - + + + (food supply) ?17 Mesopelagic fish 0 0 + 0 - (predation) Radchenko 199418 Tanner crab + + 0 - Rozenkranz et al. 200119 King crab + + 0 -

-(predation by

flatfishes)Haflinger & McRoy

198320 Shrimp + - 0 - - (predation) Ivanov 200121 Benthic Epifauna + 0 + - - (predation) Conners et al. 200222 Benthic Infauna + 0 + - - (predation) Livingston et al. 199723 Jellyfish + - + 0 - (competition) Brodeur et al. 199924 Euphausiids + 0 + + - (predation) Shuntov 200125 Copepods + 0 + + - (predation) Shuntov 200126 Phytoplankton + + + - - (greasing) Shuntov 2001

Effects: + - evident positive; - - evident negative: 0 – no evident or unclear effect.Expected trend in stock abundance:

- positive

- negative

- no expected or uncertain trend

It is believed that capelin distributions would move northward to colder waters and forego largeareas of spawning habitat, which would be available in colder years. Similarly, Arctic cod and 5species of marine mammals which are associated with the ice-edge would be restricted to waters

Page 121: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–121 DO NOT CIRCULATE OR CITE

in the Chukchi Sea for large parts of the year. Red king crab decreased in abundance during thepast warming period and may not do well under the future warming scenario. Deep-water speciessuch as Greenland turbot may only be affected during their larval and juvenile states when theyare in shallow water. Greenland turbot recruitment may be enhanced in colder years, perhaps dueto a decrease in overlap with adult pollock or to other factors (Livingston et al. 1997). In the sameway, predation by large piscivorous fish may affect the pandalid shrimp stocks. It wasdocumented that these shrimp species are among the preferred prey of Pacific cod during summerfeeding in the northern Bering Sea, especially for recruit age classes (Napazakov et al. 2001).Predation on shrimp stocks would likely increase with the enhancement of Pacific codrecruitment. Annual consumption of gonatid squid species was estimated in the Bering Sea in late1980s at 4.2 million tons (Radchenko 1992). There is low expectation for squid stock growthunder a warming regime due to increases of their main predators (for except of Greenland turbot)abundance. Some growth in the abundance of sculpins and western Bering Sea herring occurredin the 1990s and was related to the weakening of interspecific competition in the pelagic andbenthic fish communities (Naumenko et al. 2001). It is expected that these stocks mightexperience future decreases due to increases in pollock stock size.

Certain flatfish species inhabiting the Bering Sea shelf have been linked to better productivity inwarmer climates but are not expected to shift their distributions in response to ocean temperature.These species undertake a wintertime migration to deeper, warmer waters (where some spawn)and then migrate during spring and early summer to the mid shelf area where they feed on thebenthic infauna. It is expected that site fidelity is important for these species for feeding purposesand that they would tolerate moderate thermal changes. Species which successfully developedniche preferences as habitat specialists would be expected not to change much in response toclimate change whereas other “colonizing“ species which exhibit a more diverse diet might beinclined to shift with the changing conditions. These ecological characteristics are expected toinfluence how fisheries would change during the warming period. The total fishery catch duringthe expected warm climate and oceanographic regime will grow to the extent allowable undercurrent management practices. This conclusion also corresponds to historical data for the westernBering Sea fishery (Figure 12.5.6). The following attempts to forecast this increase are based onthe previously achieved maximum fishery harvest, assuming current management philosophiespersist into the future.

(Insert Figure 12.5.6 here)

The total maximum pollock catch was 4.07 million t in the Bering Sea in 1988, and averaged 3.55million t for 1986-90 (Fadeyev and Wespestad 2001). The total walleye pollock biomass wasestimated at 20 million t in the entire Bering Sea in that period (Shuntov et al. 1997).Conventional wisdom assumes that the stock reduction in the 1990s was due to a decrease inproductivity in response to environmental conditions and not overfishing. Therefore, the level ofthe total pollock harvest in the favorable period of late 1980s can be taken as the reference pointfor the prediction of future catch during the forthcoming warm period. As a rule, an increase in thepollock fishery stock forms by several average and strong year classes, as it was in the second halfof 1960s, or one super-strong year class. A super-strong generation appeared in the Bering Sea in1978 and ensured a stabilization of stock abundance and the development of the large-scalepollock fishery at the end of 1980s (Stepanenko 2001). The eastern Bering Sea population of

Page 122: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–122 DO NOT CIRCULATE OR CITE

pollock in recent years nearly doubled from 1995-2001 and supports an annual catch of more than1 million t with strict regulations. The stable condition of this population provides for the expectedincrease in future abundance with the occurrence of the favorable environmental conditions.However, it should be noted that the 1978 year class occurred 13 years after the first strong yearclass in 1965 (Wespestad 1993) and no cohorts of this strength have been observed since.Therefore, there is no expectation of a swift increase in stocks size and catch of pollock in the nearfuture.

Pacific cod harvest ranged from 33 100 – 117 650 t and averaged at 65 210 t in the western BeringSea in the period of Dutch seine trawls and, to a lesser degree, long-line fisheries in 1981 - 2001.The catch for the whole Bering Sea for that period totaled 207 110 t. This is a relatively low codcatch relative to the total estimated biomass at 3.27 million t. Adult Pacific cod may be regardedas the main predator for commercially important fish (e.g. pollock and herring) as well ascrustaceans, in particular Tanner crab and shrimp. Relative to their weight, one unit of Pacific codbiomass consumes about 1.11 biomass units of Tanner crab juveniles, 1.12 - of shrimp, 0.8 - ofwalleye pollock, 0.39 – of squid, 0.31 – of herring on the western Kamchatka shelf during the 6months of the warm season (Chuchukalo et al. 1999). Pacific cod attracts a specific attention inaspect of initiation of additional fishery efforts for regulation of predation in the fish communities.If the Pacific cod stock attains the same abundance in the Bering Sea as during the mid-1980s, it isspeculated that the total harvest of this fish could be increased (western and eastern parts of theBering Sea) to the level of the 1980s-1990s to about 350 000 t.

From 1981-1991, herring fisheries in the southeastern Bering Sea, in the vicinity of the Alaskacoast, harvested about 30 000 t and in the southwestern Bering Sea some 17 000 t from a biomasslevel of nearly 500 000 t (Figure 12.5.7). The same level of harvest can be expected for the nextwarm period. The western Bering Sea “fat herring” fishery will decline during the next decade,but the Alaska roesac herring fishery is expected to increase.

(Insert Figure 12.5.7 here)

The Pacific salmon fishery recently surpassed its top harvest level in the north-eastern Kamchatkaarea that was determined by the historical record of pink salmon catch in odd years (82 300 - 83600 t in 1997 and 1999). Undoubtedly, some decline will occur there since the local stocks ofother Pacific salmon species are not as abundant. Chum salmon coastal catch did not exceed 12200 t and sockeye salmon - 7000 t. However, these relatively high catches were made in the lastyears. Proceeding from the 5-years rate of this increase in 1990s, total chum and sockeye salmonharvest can reach about 20% and 12% surplus at the end of the 2020s. Eastern Bering Sea salmonproduction is dominated by sockeye salmon which contributed 41 million fish from a total 42million of Pacific salmon catch in the Bristol Bay in 1993. It was the historical record after the 37million fish (101 550 t) in 1983 (The Bering Sea 1996). Presently, the Bristol Bay sockeye stockhas declined and is in a period of low production. According to stock dynamics observed in the1960s-1990s (Chigirinsky 1994), low productivity periods can continue 15-20 years with anaverage annual sockeye catch at about 20 000 t. However, if we conclude that the warmconditions are favorable for Pacific salmon growth and survival during the marine life cycle, wecan expect that their annual catch will be close to that of the previous warm period 1977-1993, i.e.about 110 000 t. The proportion of sockeye salmon will increase from the mid-2000s to 2020s

Page 123: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–123 DO NOT CIRCULATE OR CITE

from 25-30% to 50-55% of the total and the proportion of pink salmon will decrease accordingly.Chum salmon will traditionally take a third place; chinook will go fourth and coho fifth in thegeneral Pacific salmon range.

Although flatfish biomass may increase in future warm years, the catch is likely to remain low dueto bycatch and market constraints. The Atka mackerel of the Aleutian Islands, skates, smelt andSaffron cod of the southeastern shelf are other potential stocks in good condition. Development ofnew markets for these fishery products could increase future fisheries harvest. There are no dataavailable for discussion of a future crab fishery in the Bering Sea since it is not well understoodwhich environmental conditions provide better survival and growth of crab larvae and juvenilestages. Further, the reasons for the sharp crab stock decrease in the Bering Sea in 1980s are notunderstood as there is some debate over whether the decline was from overfishing orenvironmental. Arctic cod, capelin, sculpin, mesopelagic fish, shrimp and squid fisheries areundeveloped at present in the Bering Sea, and no precondition exists to develop these fisheriesunder the expected warm climate regime. Some bycatch of Commander squid and sculpins duringthe trawl fishery on pollock in the Dutch seine fishery on groundfish is utilized, but the total valueof this harvest is insignificant.

12.5.6 The economic and social importance of the fisheries of the Bering Sea

In comparison to other areas of the Arctic, the commercial fisheries of the North Pacific, includingthe Sea of Okhostk, and the Bering Sea are relative newcomers. Near-shore artesanal fisheries byindigenous people have occurred for centuries in the Bering Sea (Ray and McCormick-Ray 2004;Frost 2003) (See Chapter 9). The first documented commercial exploitation of groundfish datesback to 1864, when a single schooner fished for Pacific cod in the Bering Sea (Cobb 1927),although salmon were part of commerce during earlier times. In 1882, American sailing schoonersbegan a regular handline cod fishery. As recorded in Russian literature, the California-basedfishermen ceased to sail to fish in the Sea of Okhotsk after the cod shoals near the ShumaginIslands in the Gulf of Alaska were discovered. In the western Bering Sea, the early Russianfisheries were poorly developed and limited to near shore subsistence fishing by aboriginalpeoples and settlers (Ray and McCormick-Ray 2004). Already at this early date, however, it wasunderstood that the Bering Sea contained a rich resource of fish. The herring fishery areaexpanded north to the Bering Strait and operated during two weeks in May when herring migratednear the coasts. The Pacific salmon fishery annually yielded 12 million fish of which 2 - 4 millionfish were from the Yukon delta area, while the remainder was caught by Russian developmentcompanies and Japanese corporations operating concessions on Russian rivers (Netboy 1973).

In contrast to the slow development of the early fisheries, the hunting of marine mammalsdeveloped rapidly. In the western Bering Sea, the fur seal harvest ranged from 20 000-50 000animals on the Commander Islands. A Russian-American Company was mainly responsible forthe hunting and fur purchase operations in the Eastern Bering Sea, Gulf of Alaska, and AleutianArchipelago regions from 1786 – 1862. The sea-otter harvest totaled 201 403 animals during theexistence of the Russian-American Company of which nearly one third was purchased bymerchants from the aboriginal peoples. Other marine mammal harvests included sea lion huntingon St. George Island (the Pribilof Archipelago), which yielded 2 000 animals per year and walrus

Page 124: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–124 DO NOT CIRCULATE OR CITE

hunting which annually yielded 300 - 2 000 animals until it was reduced in the 1830s due to adeclining population. Due to overexploitation, the fur seal breeding-grounds disappeared from thePribilof Islands, Unalashka Island, and adjacent areas in 1830 - 1840. From 1743 to 1823, 2 324364 fur seals, 200 839 sea-otters, about 44.2 t of walrus tusk and 47.8 t of baleen were harvestedfrom the Aleutian Arc, other islands, and the Alaska coast. The first protective measures on furseal populations from Japanese and American illegal sealers were set by Russia in 1893. There isan illustration of this circumstance in the R. Kipling ballad “The rhyme of the three sealers”:

“Now this is the Law of the Muscovite, that he proves with shot and steel When ye come by his isles in the Smoky Sea ye must not take the seal”.

In 1911, a three-sided treaty was concluded between Russia, USA, and Japan, which established asealing prohibition on the high seas in exchange for compensation paid from harvests in therookeries (Miles et al. 1982).

Large-scale commercial exploitation of the Bering Sea fish stocks was developed slowly.Between 1915 and 1920, as many as 24 American vessels fished Pacific cod. Annual harvestsranged from 12 – 14 thousand t (Pereyra et al. 1976). Small and infrequent halibut landings weremade by USA and Canadian fishermen in 1928 – 1950, which increased sharply and exceeded 3300 tons from 1958 – 1962 (Dunlop et al., 1964). In the early 1970s, halibut catch fell to a low of130 t before recovering to a high in 1987, and then slowly declined. The International PacificHalibut Commission (IPHC), established by Canada and USA in 1923 to manage the halibutresource, determined that factors such as overexploitation by the setline fishery, juvenile halibutbycatch, and adverse environmental conditions led to the decline in abundance (NRC, 1996). Inthe western Bering Sea, the exploitation of groundfish resources was mainly by small-scale coastaloperations. Information on groundfish abundance was lacking until the first Soviet PacificIntegrated Expedition in 1932-1933. This expedition covered the entire Bering Sea area anddescribed the eastern shelf and continental slope as more productive fishing grounds incomparison to the narrower western ones. Therefore, Soviet fisheries concentrated their efforts inthe eastern Bering Sea after 1959. Newly organized Soviet fishing on the eastern Bering Sea shelfand in the Gulf of Alaska provided about 600 000 t of Pacific ocean perch, yellowfin sole, herring,cod, crabs and shrimps by the mid-1960s (Zilanov et al., 1989).

The Japanese and Russian fleets expanded rapidly in 1959-1965, with vessels from the Republicof Korea and other nations also participating in later years. These fishery efforts were added to thesolely Japanese fishery efforts, which actively operated there since 1930s, especially after theWorld War II. By 1960 169 vessels from Japan were present on the Bering Sea fishing groundsalong with 50-200 vessels from the Soviet Union (Alverson et al., 1964). Significant growth offishing effort led to overfishing of several stocks. The Soviet walleye pollock fishery began in theearly 1970s after the decline of some commercially valuable fish stocks. Before that, walleyepollock was not regarded in the Soviet fishery as a target species. The Japanese mothershipoperations had three to five conventional catcher/trawlers and as many as eight pair trawlersassociated with each mothership (Alverson et al., 1964). The catch was processed at sea with thefrozen products transported to respective countries for food. Japanese catches were mostlyprocessed aboard motherships into fish meal.40 Female walleye pollock in spawning condition 40 Livers were extracted for vitamin oil.

Page 125: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–125 DO NOT CIRCULATE OR CITE

soon became an important source of roe-bearing fish, which were processed into valuableproducts. The increase in product value, combined with an increase in pollock abundance after thesecond half of 1960s led to the gradual growth of its catch: up to 550 000 t in 1967 and 1 307 000 tin 1970 (Fadeev and Wespestad, 2002). Groundfish catches were primarily by vessels from thesecountries until 1986, when U.S. fishing vessels participating in joint ventures with foreignprocessing vessels took a larger proportion of the catch. By 1990 the distant water fleets werephased-out of the eastern Bering Sea (U.S. economic zone) and U.S. fishing vessels became thesole participants in the fishery. Some fishing occurs under license from the Russian Federation inits EEZ

FisheriesUS fisheries off Alaska constitute more than half of landings and about half the value of nationallandings of fish and shellfish from federal waters (NMFS 2003). Depending on species,approximately 90% of the landings in Alaska are from the Bering Sea/Aleutian Islands area of thisstudy. All of the groundfish, crab and salmon in the US EEZ of the Bering Sea are caught bydomestic fishing entities (Hiatt et al. 2002). In the Russian EEZ the preponderance of the harvestsare taken by domestic fleets with a decreasing portion being harvested under agreements withneighboring states. In 1997 it is estimated that the Russian Far East fisheries account for 70% ofthe Russian Federation total fisheries production (Conover 1999, Zilanov 1999) however thisproportion may be decreasing due to the declines in pollock, crab, herring and other species notbeing offset by the increases in Pacific salmon.

In the Bering Sea, walleye pollock is the major harvest by volume and value with Pacific cod,flatfish, salmon and crabs constituting most of the remainder. Total wholesale (raw fish landings)value for groundfish harvests in the Eastern Bering Sea is approximately $426 million (2001).Total primary processed value was approximately $1.4 billion.Crab harvests, predominantly fromthe Bering Sea/Aleutian Islands area, amounted to $124 million even at the generally lowpopulation abundances noted earlier (Hiatt et al. 2003). Pacific salmon, a large portion of whichcomes from the Bristol Bay region and the Yukon River area, showed Bering Sea catch valuedbetween $122 million (2001) and $179 million (2000) (Link et al. 2003). The CDQ Program,which allocates 10% of the total Bering Sea TAC to 65 coastal communities organized into sixCDQ corporations, earns more than $40 million dollars annually (NPFMC 2003a). No separatevalue is assigned in this study to recreation or subsistence harvests in the Bering Sea due to lack ofadequate analyses despite their local and cultural significance. Economic value data for theRussian Far East are difficult to locate (Pautzke 1997). Press reports for product value estimatetotal 2001 production to account for $3.0 billion (Pacific Rim Fisheries Update, May, 2002, p. 3).Since the transition to a market economy began in the early 1990s and the Soviet stylemanagement of fisheries has changed, it appears that there are significant tracking and reportingdifficulties with less fish being landed to avoid taxation and fees. Instead harvests may betransferred at sea or transported directly to foreign markets by fishing vessels (Velegjanin 1999).Thus, production and value data must be treated with caution until a more robust accountingsystem is developed.

Fishing fleet and fishersVirtually every fishing vessel in the Bering Sea fleet is homeported outside of the region. Vesselsmust be of requisite size of vessels to weather the environmental conditions and to have adequate

Page 126: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–126 DO NOT CIRCULATE OR CITE

scale efficiencies to operate in the area. These factors plus the lack of deep water moorings as wellas other support services make the eastern North Pacific a largely “distant water” fishery. Overall,the number of vessels eligible to fish for the increasing stocks of groundfish in the federal watersof the Bering Sea region has decreased since the mid-1990s from a total of 464 [1995] vessels to398 [2001]. This is the case for all groundfish vessel classes and types. For 2001 there are 163hook and line [longline] vessels, 81 pot vessels, and 162 trawl vessels fishing of whichapproximately 20 are at-sea capture/processors for pollock. The overall decrease in number comesas a result of rationalization programs for pollock under the American Fisheries Act 1998 and theNorth Pacific Fishery Management Council’s license limitation program for all species41 (Hiatt etal. 2003). With respect to other sectors, there are approximately 274 eligible Bering Sea/AleutianIslands crab fishing vessels, 2 500 catcher longliners42 mostly involved in halibut/ sablefish andPacific cod fisheries and some 5 200 salmon fishing vessels of various types (NRC 1999)

Employment in the groundfish harvesting sector (at-sea catching and processing on land as well asmotherships) amounts to 4 000 full-time equivalent jobs including skippers, fishing crew,processing crew and home office staff for 2001 (NMFS 2003b). With a few exceptions, most ofthis employment is in the form of relatively small corporations. NPFMC license limitationregulations basically limit the size and ability to grow of existing catching entities. Thus, few largeintegrated harvesting and processing companies exist. Still, even the smaller entities are dealingwith multi-million dollar investments with substantial annual operating expenses, e.g., a typicalcatcher vessel with length of about 120-130 feet would require a family owner or small business tohave a fair market worth of $2.5-3.5 million (NRC 1999).

In the western Bering Sea, the picture is somewhat similar to that of the Alaska EEZ. A large partof the harvesting capacity is located in the southern parts of the area as are financial and other fishsupply and repair services. The number of fishing vessels has declined drastically following theend of the Soviet era distant water fishing due to other nation’s extending their EEZs as well asefforts to renew the fishing fleet and to reorganize it on market economy terms (Zilanov 1999).Between 1990 and 1999 the Russian fishing fleet decreased by nearly 44% in number. Most of thefleet became private in the form of joint stock companies (56.7%), or were transferred tocooperatives (kolkhozes; 23.7%), private companies (12.5%) or joint Russian-foreign ventures(2.4%]) (Zilanov 1999). In the Russian Far East, this opened the opportunity for small and mid-scale fisheries to develop while some large entities under Soviet style fisheries have transformedand remained dominant forces (Platonov, pers. comm.). Likewise, total employment in thefisheries sector nation-wide fell from 550 000 in 1990 to 398 000 in 1998. Contributing to thedecline in employment in the Russian Far East was a considerable exodus of population from theregion as persons assigned to duties there returned to families and friends in their home regions.

41 This figure does not include halibut/sablefish IFQ vessels.42 Including Alaska state-water vessels.

Page 127: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–127 DO NOT CIRCULATE OR CITE

The land side of the fishing industryApproximately 70% of the Bering Sea harvests are processed on shore in a relatively few [8]groundfish processing plants near Dutch Harbor/Onalaska (NMFS 2003b). Recent efforts arebeing made to locate processing facilities on Adak Island in the Western Aleutians. During periodsof high abundance of king crabs and snow crab in the Bering Sea crab processing is performed onthe Pribilof Islands. Salmon tendering and processing is focused around Bristol Bay although notexclusively. Sites where processing occurs require significant infrastructure for processing as wellas for providing services to the fishing fleet. Given the remote nature of the Bering Sea fishprocessing, the communities in which these activities take place are highly dependent on thefishing industry for economic activity with government services and tourism being distant rivals.Most of the groundfish activity in processing is adjacent to the densest aggregations of groundfishand where it is possible with a catcher vessel with refrigerated sea water holds to make relativelyrapid trips to maintain product quality. Still, for some species and products, e.g., high gradesurimi, it is difficult for shoreside processors to compete.

Employment in Alaskan shoreside processing for groundfish is estimated at 3 525 full-timeequivalents (NMFS 2003b). The number of processing jobs onshore in the Bering Sea hasincreased as much as 50% between the early 1990s and the present because of policies decreasingthe amount allocated to at-sea processing versus onshore processing. Much of this employment oflabor is not of Alaska residents. Rather, the work force tends to be an ethnically diverse set ofwork permit holders from other parts of the globe, mostly West Coast United States, Mexico andPhilippines. Through time, Alaskan communities in the Bering Sea region are being transformedas workers stay on and move up the corporate ladder.

In the Russian Far East the style of fisheries has been a significant fraction processed at sea onlarge catcher processors with the rest being shorebased processing, cold storage, etc.. Withdomestic demand low in terms of the ability to compete with global market price and other tax andregulatory hurdles on shore, a substantial incentive is present to process offshore and to exportdirectly (Velegjanin 1999). This has contributed to a sizeable drop in domestic consumption aswell as employment in shoreside processing and other services to the fishing industry. It appearsthat the period of transition to a market economy has been difficult but that the learning curve istrending upward with new management institutions and experience. Still, without the fullcooperation of the fishing industry with management and tensions over allocation of revenuesbetween the Far East and Moscow, the regional administrations have a long way to go before theindustry stabilizes.

Fisheries communitiesThe North Pacific fishing communities surrounding the Bering Sea are different from the NorthAtlantic models. There is not a history of small fishing coastal communities. In the eastern BeringSea some 65 communities exist with a total population of around 27 500 inhabitants. They arefrequently inhabited by a large percentage of Native Alaskans but not exclusively (NPFMC2003a). Until they became participants in the Community Development Quota Program, they hadlimited coastal subsistence fisheries as well as some developed small-scale commercial fisheriesfor salmon and halibut. Involvement in the groundfish and crab fisheries has provided valuableincome and employment as well as a role in management of the fisheries offshore. The main locusof the fish processing on Akutan and Dutch Harbor/Onalaska had been important for crab, halibut

Page 128: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–128 DO NOT CIRCULATE OR CITE

and some salmon fisheries. It was not until foreign and domestic investment was encouraged inshoreside processing of groundfish in the late 1980s that these communities were transformed.Loss of access to fishing in the US zone prompted Japanese investment in processing so that rawfish could be purchased at low prices and benefits gained in value-added processing fromshorebased plants.

The history of the purchase of Alaska from Russia in 1867 and its status as a territory untilStatehood in 1959 was that of a domestic colony. In particular, fishing interests in WesternWashington and Oregon were some of the prime early investors in Alaskan fisheries. Ownershipof the highly seasonal Alaskan canneries was mostly external to Alaska. Salmon fishing broughtlabor from down south. Halibut fisheries could be developed as soon as ice-making andrefrigeration technologies permitted catching and transport of fish to southern markets. Early crabfishing interests were based out of Seattle. Thus, the fisheries of Alaska have strong personal,financial and service connections to Seattle due to the laws of comparative advantage. Alaska is ahigh cost area for living and carrying out a business (NRC 1999). In the federal water fisheries,residents of other states may not be discriminated against in management regulations, whichfurther enforces the long, mostly cooperative, relationship between fishing interests in Washingtonand Oregon and those in Alaska. Overall dependence on fisheries varies by community but inAlaska as a whole fisheries is a distant second to oil production in terms of revenue from resourceextraction and for some cities with onshore processing, fisheries a the prime source of locallanding tax revenue.

Similar to Alaska, small Native Russian settlements existed around the western Bering Sea. Withthe colonization by Russians, larger towns sprang up and, later, during the Soviet Union era thesegrew as bases for resource development and national defense. Population in the Far East’s 7administrative regions is concentrated in coastal cities and it has declined slowly over the decadeof the 1990s (Zilanov 1999). Several large cities account for the majority of the population so thatmuch of the Russian coastline is undeveloped. Fisheries tend to be dominated by fishing interestsin Vladivostok and Nakhodka. Increasingly stronger demands are being made by other regionalfishing bases for more autonomy in management and greater allocations to proximate users.

MarketsThe relatively low populations of the Bering Sea region do not constitute a very large local marketfor the large-scale fisheries. Thus, both Alaska and the Russian Far East look to distant markets athome and abroad. For Alaska the prime markets are Japan, Korea and China with Europeproviding entry for some products. Over 90% of Alaskan fish is exported. Korea and now Chinawith their relatively low wage labor have served as processing centers for some products that arere-exported, i.e. imported back in some value added form. For the Russian Far East, exports havestarted to play an increasing role in the fisheries economy. During the Soviet Union era as muchas 80% of the Far Eastern fish products were processed and sent on to domestic markets in thewestern more populous parts of the country. The rest was exported or taken under fisheriesagreements with neighboring states to obtain hard currencies badly needed by the centralgovernment. While low effective demand, i.e., domestic consumers, is not able to payinternational prices for seafood products, many of the higher value species are being exported andlow value species and products are being imported so that now about 50% of the seafoodharvested is destined for export (Zilanov 1999). Clearly, the remote Bering Sea is a major player

Page 129: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–129 DO NOT CIRCULATE OR CITE

in terms of global seafood markets where declines in abundance of Atlantic cod, for example,opens markets for fillets of pollock at the same time demand for pollock surimi products seems tobe slackening as a result of weak Japanese and Korean markets. Similarly, high abundance ofsnow crab in Canada causes market erosion for the same species in the eastern Bering Sea.

Due to the significant price competition from farmed Atlantic salmon, the wild salmon dependentfishing enterprises and communities are facing major adjustments. Even though North Pacificwild salmon stocks are quite abundant at present, the large quantity of farm raised salmon and itsmethod of sale and delivery reduce the price that can be obtained. There is some consideration inAlaska and Russia about starting aquaculture but it is recognized that the investment, organizationand technology may be significant hurdles (Link et al. 2003). Given the experience with salmon,there is concern as well over developments in the farming of halibut, sablefish and cod becomingcompetitive with wild stock harvests.

Management regimeIn Sections 12.5.1 and 12.5.3 the development of an international regime for fisheries managementin the North Pacific is sketched out. Currently, the US and Russian EEZs are the majormanagement jurisdictions although the multilateral conventions for management of the ‘DonutHole’ fishery outside of these boundaries plays an important role in fisheries management.Similarly, the Convention of the North Pacific Science Organization and the North PacificAnadromous Fisheries Convention provide frameworks for scientific exhange and cooperation.Even though main covering activities fall somewhat south of the ACIA boundaries, theConvention on High Seas Driftnets constrains fisheries on the high seas that are considered tohave potential to intercept salmon of Russian and United States origin as well has to have negativebycatch affects for Dall’s porpoise and some seabird species. Bilateral agreements such asbetween Canada and the United States for salmon and halibut management and between Russiaand Japan for salmon also exist.

At the national level, the Magnuson-Stevens Fishery Conservation and Management Act is theprime legislation guiding fisheries management in federal waters. In Alaska this means all watersbeyond three nautical miles from the state’s baselines and out to 200 n.m. is under federaljurisdiction. Other relationships exist such as federal management for halibut in all waters due tothe international treaty and Alaskan state jurisdiction (with federal oversight) over crabs ascreatures of the continental shelf and salmon which are harvested within state waters (Miles et al.1982a). The waters off Alaska constitute one of the nation’s eight fishery management region.This is administered by the regional office of the National Marine Fisheries Service, withmanagement decision making taking place in the North Pacific Fishery Management Council – anadvisory body to the regional director and thereby to the Secretary of Commerce. The federalregulations essay to develop a decision process that is comprehensive, transparent and open toparticipation by all interested parties (NMFS 2003b),

The main tools for fishery management are Fishery Management Plans that set forth the rules andregulations for management of each species or species complexes. Under current managementapproach, TAC is set on an annual basis in the Stock Assessement Fishery Evaluation process(e.g., NPFMC 2002). In this process ecosystem considerations are explicitly made available at thetime of the TAC setting process in the form of a Chapter of the SAFE document discussing

Page 130: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–130 DO NOT CIRCULATE OR CITE

ecosystem trends and relationships to fishing as well as in the environmental assessments requiredunder the National Environmental Policy Act. All of the meeting of the Council and its AdvisoryCommittee and Scientific and Statistical Committee are public meetings. Thus, any interestedparty can be present and participate in deliberations of Plan Development Teams setting TACs toany of the deliberations on other matters.

The NPFMC has developed some fairly innovative approaches to management. Scientific adviceis rigorously adhered to in the setting to TACs and conservative harvest limits are applied. A capof two million metric tons has been set on total removals in the fishery even when allowablecatches might be considerably above that cap. Bycatch is counted against TAC and targetfisheries can be closed if the bycatch limit is reached before the target fishery TAC. Larger boatsare obligated to carry and pay for one or more observers to gather scientific information aboutharvests. Species like halibut, salmon and herring are considered prohibited species in thegroundfish and other non-target species fisheries. Finally, significant areas of the fishing groundsare closed to trawling to protect habitat necessary for other species, e.g., red king crab savings area(Witherell et al. 2000). In addition, much of the current NPFMC work is on developing spatiallyexplicit relationships between fisheries and fish habitats under the Essential Fish HabitatProvisions of the MSFCMA (NPFMC 2003b). Further, there is a Council emphasis onrationalization of fisheries through share-based management systems like the IFQ program forhalibut and sablefish [and as proposed for Bering Sea and Aleutian Islands crabs] or using acooperative approach as done for pollock under the American Fisheries Act from 1999.

In the EEZ of the Russian Far East, the basic management system and issues are similar to thosedescribed in 12.2.6 with the exception of the reciprocal fishing agreements. The regionaladministration is subject to central control for setting allocations and for Border Guardenforcement. Based on comments about the implementation of enforcement in western Russia, itappears that the US Coast Guard and the Boarder Guards have developed a more effective form ofcooperation in enforcement in the Bering Sea in particular with respect to the fishing zoneboundary and High Seas driftnet fishing. As noted earlier in this presentation of Bering Seafisheries, the scientific basis for setting allocations in Russia is similar to that of the Alaska region.Significant concerns have been expressed about how well such allocations are being followed andenforced (Velegjanin 1999). Similarly, the role of the central as opposed to the regional fisheryadministrations in setting quotas and allocating them is being challenged. For the past severalyears, significant portions of the total allowable harvest are being auctioned to the highest bidders.This innovative effort has been controversial.

12.5.7 Variations in Bering Sea fisheries and socio-economic mpacts: Possible scenarios

The predominant changes over the years in the commercial fisheries of the Bering Sea have beenin the distribution of the harvests among nations and sub-national user groups. Secondarily,changes in species composition of the catch due to environmental conditions and fishing pressureshave also had impacts on those employed in the fishing industry and their communities. While thelatter changes are of primary interest in this study, it is very much worth noting how extensive theadjustments have been to changing claims to jurisdiction in the Bering Sea (Miles et al., 1982a).The tremendous dislocation of fishing fleets from Japan and then-Soviet Union post EEZ

Page 131: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–131 DO NOT CIRCULATE OR CITE

extension, clearly shows that major adjustments can be made but with considerable hardship.Simlarly, the response of the US fishing industry assisted by favorable government incentivesshowed how quickly it can respond to opportunity. The question here is how fully occupiedfisheries can respond to sustainable and precautionary management.

Fisheries in the Bering Sea are largely a post World War II phenomenon where the technology andscale of enterprise necessary to fish in the inhospitable, enormous expanses of the remote BeringSea shelf developed. Along with these developments in mothership operations, food processingtechnology came development of new markets for species like walleye pollock that was previouslynot a target species despite what appear to have been large quantities of pollock available.Relatively little is known about the fisheries ecosystem of the Bering Sea prior to the developmentof intensive industrial scale fisheries in the area. Attention has been given to the early whalingactivity in the North Pacific that took its toll on the more valuable and easier to harvest species. Itis uncertain how removal of this biomass of whales may have affected controls in the Bering Seaecosystem (NRC 1996). However, it is difficult to dismiss the loss of whales as insignificant. Ofcourse, the decline of the North Pacific whaling was offset by effort being redirected toward otherparts of the ocean including the Antarctic. For some whaling communities where rending andprocessing occurred onshore, the displacement of effort meant the end of whaling as a source ofemployment and income.

Fisheries development after WW II tended to target the highest value species first. Despite effortsto develop sufficient scientific information and international management under INPFC there is noquestion that some stocks like Pacific ocean perch and other longer-lived, high value species wereoverfished. Still the opportunity to fish on previously unfished stocks of very large size and extentresulted in significant employment and income benefits for those who did assume the risks. Withthe development of coastal state management, the question of how to manage properly these large-scale fisheries developed. In the early period, most observers do not consider reported harvests tobe an accurate representation of actual catches. The valuable but limited joint scientific survey ofstocks provides some information. For all practical purposes and certainly relative to the efforts toseparate anthropogenic-induced change versus environmental variability induced change, theperiod of record is extremely limited. A further compounding factor is the nature of the BeringSea itself – a large, remote, difficult area to characterize and monitor. Thus, the linking ofscientific advice to fisheries management objectives has been a process of successive refinement.The ability to assess the range of natural variability in stock sizes is very imprecise and how theecosystems function is only now being modeled with a significant degree of sophistication to startto puzzle out answers (NRC 2003).

Above it is noted that some eight periods of alternating cold/warm sea temperatures can beidentified in the instrumental record. The extent to which these have altered population sizes andconcentrations is difficult to tease of out of the data for the reasons mentioned above.Furthermore, population sizes may have been affected by both high levels of fishing for some highvalue species, and low levels of fishing for species with low market value or with high levels ofbycatch. Fishery management is generally thought to mediate for overfishing and to manage tomaintain abundance of desired species. Since the mid- to late-1970s warmer temperatures andassociated atmospheric and sea surface circulation may have favored salmonids, winter-spawningflatfish, walleye pollock, Pacific cod and Pacific halibut but have been detrimental to capelin,

Page 132: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–132 DO NOT CIRCULATE OR CITE

Pacific herring, shrimps, and several species of large crab. Fisheries have fully developed onthose species that are at high levels of abundance and have essentially exited those whoseabundance is low (NMFS 2003).

The US fishing industry in the Bering Sea survived changes in relative abundance in speciesduring the growth phase by having the luxury of shifting from crab fishing to walleye pollockfishing and Pacific cod fisheries, for one example. This has altered the conditions for traditionalcrab processing ports in the Pribilof Islands but has contributed to the growth of the groundfishprocessing in Dutch Harbor/Onalaska and Akutan. The big question is what would happen to theindustry if there were a pronounced shift to a coldwater period? Fisheries management isattempting to rationalize effort in these fisheries to increase efficiency, reduce bycatch ofprohibited species and to encourage capture of value added through higher quality products andutilization rates. This tends to reduce flexibility of movement, such as took place as the domesticfisheries developed. So far there is little advance planning for how fishery management couldoperate in a transition between such cold and warm regimes. For most of the groundfish speciesmanagement under quotas, the expectation is that small or large year classes would be detected inthe assessments and that quotas would rise and fall to prevent overfishing. For species with shortlife spans, such an approach may not be as effective, yet the high natural variability is somethingthat managers consider. For exceptionally long-lived species like rockfish, experience shows thatvery conservative harvest rates may need to be applied and no-take marine reserves have beensuggested as a tool to insure against loss of older highly productive fish.

This is a non-trivial issue as can be seen in the massive buildup of the red king crab fleet in thelate 1970s to harvest anomalously large quantities of a Bristol Bay stock that subsequently crashed– likely due to recruitment failure from environmental conditions. Additional effort entered thecrab fleet with the strong stocks of snow crabs in the 1980s and 1990s. Sharp decline in thesefisheries, again associated with environmental conditions, has caused severe problems foroperators with high debt service and relatively few assets, However, those with low debt andsubstantial assets are more capable to ride out the low production intervals. These problems in theeastern Bering Sea crab fisheries provide incentives to seek out other pot gear fishingopportunities so that now other fixed gear operators in Pacific cod are being squeezed by the entryof crab vessels into their traditional fisheries. This fishing behavior domino effect is highlypredictable even if the underlying phenomena driving it are not.

Another effect noted above is less favorable conditions for pinnipeds under warmer conditions.This effect appears to be indirect through the food web as opposed to direct through predation,although there may be interacting effects. This has a pronounced effect on fisheries in the EasternBering Sea under United States jurisdiction where spatial extent and timing of walleye pollock,Atka mackerel and Pacific cod fisheries have been modified as a precautionary measure to protectSteller sea lions (NRC 2003). In this way, environmental conditions affecting non-target speciescan be significant enough in terms of management for endangered and threatened species that itincreases fishing costs and reduces revenues for fishing.

From the standpoint of fishing dependent communities, the many subsistence fishing villages onthe shores of the Bering Sea experience climate variability directly. The approximately 65 CDQcommunities in the Eastern Bering Sea have connections with climate variability directly through

Page 133: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–133 DO NOT CIRCULATE OR CITE

subsistence fishery activities and through the participation in the industrial fisheries through theirpartners. Industrial fisheries in the Bering Sea are quite dependent on large-scale shore basedprocessing plants that are operating, like the fishery itself, under difficult conditions. This isbecause catcher vessels that make deliveries to the shoreside plants must now run farther offshoreto fishing grounds because of the closed areas to buffer fisheries sea lion competition for prey. At-sea processors are somewhat more adaptable to changing environmental conditions because theyare capable to follow the fish and fishing conditions and deliver to various ports.

Salmonids present an interesting case with fairly well documented aggregate North/South shifts inproduction under relatively warm and cold periods (Beamish and Bouillon 1993, Hare and Francis1995, Mantua et al. 1997). This does not explain all sources of variability by any means but it hasbeen used successfully in gaining a better management understanding. Exacerbating these trendsnow is the downward pressure on market price due to decline in the Asian market and competitionfrom farmed sources of Atlantic salmon. Even at high levels of abundance fishing for wildsalmonids in the Bering Sea has become marginal at best. This may force fundamental change inthe structure and practices of salmon fishing in the area. In addition, the extremely low returns tothe Yukon River make survival precarious for Alaskan and Canadian native peoples dependent onthe abundance of migrating salmonids. This has brought disaster relief in the form of federal andstate loans and welfare programs. Based on recent studies (Kocan et al. 2002) the decline inYukon stocks may be due to factors beyond control of fishery managers but may relate to warmerenvironmental conditions. Already the low levels of salmon have renewed calls for reducing thebycatch of salmon in trawl fisheries in the Bering Sea. Even though the salmon bycatch rates havebeen reduced, more salmon are desired by Yukon and other peoples. The trawl industry that hasbeen pushed from low bycatch to higher bycatch areas due to Steller sea lion protection measureshas taken a proactive real time measures to further avoid salmon bycatch.

Edge of the ice and extent and timing of the melting of ice as well as the development of the coldpool can have confounding positive and negative effects on fisheries through the tendency toconcentrate or disperse certain species or to contribute to the productivity of the Bering Seaecosystem. Direct impacts on crab pot loss from ice edge shifts have been noted in the opiliofisheries during some cold years. The economic consequences of these types of variability areconsidered part of the risks of fishing in the Bering Sea. At the present time, it is possible to speakonly in general terms about the true effects of climate variability on fisheries in the Bering Seafrom a socio-economic perspective. Any such analysis is derivative of better scientificunderstanding of the interrelationships in the ecoystems and an increased ability to predict.Complicating this analysis is the difficulty of knowing what the dynamics of the fisheries are dueto the short period of record. Also complicating matters is that external market forces arecurrently exhibiting enormous impacts on value of the fisheries and these pressures may be ofmore importance than variability in landings or overall fish abundance.

At the industrial scale of fishing and processing that is characteristic of groundfish and crabfisheries in the Bering Sea the social effects are derivative of the broader economic trends. Lowerprices and quantities generate fewer jobs and jobs that pay less. However, high global marketprices for species like red king crab may offset declines in stocks when other sources of supplydecrease, e.g., in Russian waters, or increase as in the case of red king crab in northern Norway.Rationalization through the economic system or through fishery management systems may allow

Page 134: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–134 DO NOT CIRCULATE OR CITE

for greater long-term stability with less overall investment in harvesting and processing. Feweroperators earning a greater return on investment are more likely to be able to absorb swings inabundance due to environmental or other conditions. It is difficult to assess impact on consumersas the global trade in fisheries tends to find ways to satisfy market demands. However, withrespect to fishery dependent communities and small family-owned enterprises as found in thesalmon fisheries, the impacts can be devastating as the high costs of fishing may exceed the priceavailable (Link et al. 2003). With most assets tied up in ownership of a salmon fishing vessel, gearand a limited entry area permit and with no place to sell, this is a formula for disaster. Manyoperations face no choice but bankruptcy. In communities where there are numerous such entities,there appear to be few alternatives.

12.5.8 Ability to cope with change

The last several decades of Bering Sea fisheries have been built around fairly consistent warmwater favored species although there are some exceptions between the Western and Eastern BeringSea that may be worth investigating further. There is no question that coastal states have benefitedmore in recent years than distant water fishing nations. However, the management response to atransition to a cold phase has not been adequately considered nor has the opposite, i.e., continuedprevailing warm periods. The regime shift question is explored in the Table 12.5.7 below.Assuming a shift between a cold and warm regime in the mid-1970s, which as described for theBering Sea above is a mere +/- 1 °C., imagine what the effects might be and anticipate changes inother factors. Salmon increase in number but global market price declines. Groundfish abundanceincreases but Asian market is weak because of other economic conditions. Snow crab declines inthe US but Canadian stocks increase due to possible unfavorable or favorable environmentalconditons

Page 135: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–135 DO NOT CIRCULATE OR CITE

Table 12.5.7. Trends in Abundance and Value in Major Alaska Fisheries (Inflation AdjustedDollars)*

Species Stock 1977 Value 1977 Stock 2001 Value 2001 DiscussionSalmon 200 million

fish500 million $with peakvalue in 1988of 1.18 billion$

175 millionfish

205 million $ A smalldecrease intotal catch buta largedecrease inprice due tocompetitionwith farmraised fish.

Groundfish Very small USharvest

2-3 million $but rapidlyincreasing to1.0 billion $ in1988 as a resultof Americani-zation

1.65 m.t.million harvest

400 million $ Whitefishmarkets strongyet price weakbut $ alsoweak.

Shellfish[primarily crabspecies butsome shrimp inearly years]

Red king crabstrong otherspecies smallharvests

440 million $drops whenRKC bubblebursts butOpilio crabtakes over

Most species atlow levels

125 million $ Strongcompetition inOpilio fromEastern Canadabut weakcompetitonfrom Russia

Pacific halibut Low catchmost likely dueto foreign fleetbycatch

less than 30million $

High levels ofhalibutabundance

150 million $ Strong stocksand good pricevis a vis otherwhite fish

Herring Lowabundance

less than 30million $although valueincreased inmid 1980s /mid 1990s to50+ million $

Lowabundance

less than 30million $

Herring insame situation

Source: ADFG as cited by Pacific Fishing, January 2002.

In previous sections it has been demonstrated that a very small difference in ocean environmentalconditions can be detected as a cold or warm phase in the Bering Sea. While there is not a globalclimate change scenario for the Bering Sea per se, this shift between cold and warm periods doesprovide some working hypotheses about what to expect in the area. At a minimum, it is likelythat the conditions that have prevailed over the last several decades might constitute a baseline forslightly warmer conditions. Therefore it would not be likely that there would be a resurgence ofcrab or shrimp populations or for herring and capelin and other small pelagic species. Theecosystem would continue to be dominated by walleye pollock and Pacific cod and flatfish.Walleye pollock juveniles may continue to occupy the role of coldwater forage fish. Salmonids

Page 136: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–136 DO NOT CIRCULATE OR CITE

would likely remain abundant in the aggregate in northern waters but, in the south off BritishColumbia and Washington and Oregon stocks would be depressed.

Socio-economically this base case would replicate the current system in terms of production offish commodities. Through improvements in fishery management, it may be possible to increaseharvests of certain stocks by managing for recovery to levels of former abundance. However, it isprobably just as likely that unforeseen events or interactions may produce management mistakesthat offset such gains in a dynamic ocean system. Exploitation of underutilized species may befeasible to some degree. Some gains in catching the whole total allowable catch might be possibledue to changes in gear and fishing practices to obtain lower bycatch rates. In order to attainincreases in value added and in utilization rates, the industry may need to be further rationalized.

Some additional factors must be taken into this base case simply going forward. Previously notedare declines in marine mammal and seabird populations. In some cases, fishery interactions, whilemodest and indirect, may justify further efforts to protect the numbers of seabirds and marinemammals under an adverse environmental regime and these may constrain fisheries more thanwould be necessary if that stock were the sole interest of management. Similarly, environmentalgroups may change the level of performance to which they expect fishery management to attain,i.e., no detectable impact standard or negligible effect standard and this would alter themanagement “field of play”.

If a continued warming is postulated, it is expected that there could be a range of temperatures,which would continue to generate positive recruitment and growth scenarios for some of the warmadvantaged species (Table 12.5.1). These conditions would be worse for pandalid shrimp andmost crab species. If walleye pollock stocks increase, their impact as a predator on fish may alsoincrease with unpredictable outcomes. Certainly it cannot be expected that migration paths,timing of spawning, timing of start of primary production and composition of species wouldremain the same. Similarly loss of the sea ice effect may change the early spring ecosystemprocesses but greater surface exposure to winter storm conditions might increase nutrientcirculation and resuspension from shallower waters. So far, there does not appear to be crediblepublished forecasts on what would happen north of the Bering Strait with respect to fisheriesunder a no or low sea ice scenario.

12.5.9 References

Alverson, D.L., A.T. Pruter, and L.L. Ronholt. 1964. A study of demersal fishes and fisheries ofthe northeastern Pacific Ocean. H.R. MacMillan Lectures in Fisheries. Institute of Fisheries,University of British Columbia, Vancouver.

Aydin, K.Y., Lapko, V.V., Radchenko, V.I., and Livingston, P.A. 2002. A comparison of theeastern Bering and western Bering Sea shelf and slope ecosystems through the use of mass-balance food web models. U.S. Dep. Commer., NOAA Tech. Memo. NMFS-AFSC-130. 78pp.

Bailey, K.M., Francis, R.C., and Schumacher, J.D. 1986. Recent information on the causes ofvariability in recruitment of Alaska pollock in the eastern Bering Sea: physical conditionsand biological interactions // Bull. INPFC. 47: 155-165.

Page 137: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–137 DO NOT CIRCULATE OR CITE

Beamish, R.J., and Bouillon, D.R. 1993. Pacific salmon production trends in relation to climate //Can. J. Fish. Aquat. Sci. 50 (5): 1002-1016.

Brodeur, R.D., and Ware, D.M. 1992. Long term variability in zooplankton biomass in thesubarctic Pacific Ocean // Fish. Oceanogr. 1 (1): 32-38.

Brodeur, R.D., Wilson, M.T., Walters, G.E., and Melnikov, I.V. 1999. Forage fishes in the BeringSea: distribution, species associations, and biomass trends. In: T.R. Loughlin and K. Ohtani(eds.). Dynamics of the Bering Sea. Fairbanks: Alaska Sea Grant College Program. AK-SG-99-03. P. 509-536.

Bulatov, O.A. 1995. Biomass variation of walleye pollock of the Bering Sea in relation tooceanological conditions. In: Beamish, R.J. (ed.), Climate change and northern fishpopulations, Canadian Special Publication in Fisheries and Aquatic Science, Vol. 121, pp.631-640.

Chigirinsky, A.I. 1994. Commercial fishing of Pacific salmon in the Bering Sea // IzvestiyaTINRO (TINRO Transactions). 116: 142-151 (in Russian).

Chuchukalo, V.I., Radchenko, V.I., Nadtochii, V.A., Koblikov, V.N., Slabinskii A.M., andTerent’ev, D.A. 1999. Feeding and some features of ecology of Gadidae of the westernKamchatka shelf in summer 1996 // Journal of Ichthyology. 39 (4): 309-321 (translated fromVoprosy Ikhthiologii. 1999. 39 (3): 362-374).

Clark, W. G. and S. R. Hare. 2002. Effects of climate and stock size on recruitment andgrowth of Pacific halibut. North American Journal of Fisheries Management 22:852-862.

Cobb, J.N. 1927. Pacific cod fisheries. US Bureau of Fisheries, Document 1014, Appendix VII tothe Report of the US Commissioner of Fisheries for 1926. U.S. Government Printing Office.Washington, D.C.

Conners, M.E., Hollowed, A.B., and Brown, E. 2002. Retrospective analysis of Bering Seabottom trawl surveys: regime shift and ecosystem reorganization // Progr. Oceanogr. 55 (1-2): 209-222.

Conover, S. 1999. The 1997/98 Directory of Russian Far East Fishing Companies. AlaskaCenter for International Business, Anchorage. 99 pp.

Cushing, D.H. 1969. The fluctuation of year-classes and the regulation of fisheries // Fiskeridir.Skr. Ser. Havunders. 15: 368-379.

Davydov, I.V. 1972. On oceanological fundamentals of yield formation of the separate herringgenerations in the western Bering Sea // Izvestia TINRO (The TINRO Transactions). 82:281 - 307 (in Russian).

Decker, M.B., Hunt Jr., G.L., Byrd Jr., G.V. 1995. The relationships among sea surfacetemperature, the abundance of juvenile walleye pollock, and the reproductive performanceand diets of seabirds at the Pribilof Islands, southeastern Bering Sea. In: Beamish, R. J.(ed.), Climate change and northern fish populations, Canadian Special Publication inFisheries and Aquatic Science, Vol. 121, pp. 425-437.

Dunlop, H.A. et al. 1964. Report of the International Pacific Halibut Commission 1964. IPHC.Seattle, WA.

Fadeev, N.S., and Wespestad, V. 2001. Review of walleye pollock fishery // Izvestiya TINRO(TINRO Transactions). 128: 75-91 (in Russian).

Finney, B. P., I. Gregory-Eaves, M. Douglas and J. P. Smol. 2002. Fisheries productivity in thenortheastern Pacific Ocean over the past 2,200 years. Nature, 416, 729-733.

Francis, R.C., Hare, S.R., Hollowed, A.B., and Wooster, W.S. 1998. Effects of interdecadalvariability on the oceanic ecosystems of the NE Pacific // Fish. Oceanogr. 7(1): 1-21.

Page 138: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–138 DO NOT CIRCULATE OR CITE

Frost, Orcutt. 2003. Bering: The Russian Discovery of America. Yale University Press. NewHaven. 330 pp.

Gavrilov, G.M., and Glebov, I.I. 2002. The composition of demersal ichthyocenosis in the westernpart of Bering Sea in November, 2002 // Izvestiya TINRO (TINRO Transactions). 130:1027-1037 (in Russian).

Ginter, J. 1995. The Alaska community development quota fisheries management program. Oceanand Coastal Management. Vol. 28, Nos. 1-3, pp. 147-163.

Girs A.A. 1974. Many - yearly variations of the atmospheric circulation and long - term trends inthe change of hydrometeorological condition in the Bering Sea area. In: Oceanography ofthe Bering Sea. Proc. Int. Symp. Ch. 25. P. 475- 482.

Glebova, S.Yu. 2001. Coordination of types of atmospheric processes abo0ve the far-eastern seas// Izvestiya TINRO (TINRO Transactions). 128: 58-74 (in Russian).

Hare, S.R., and Francis R.C. 1995. Climate change and salmon production in the northeastPacific Ocean, p. 357-372. In: Beamish, R.J. (ed.), Climate change and northern fishpopulations, Canadian Special Publication in Fisheries and Aquatic Science, Vol. 121, pp.357-372.

Hare, S. R., and N. Mantua. 2000. Empirical evidence for North Pacific (clilmatic) regime shiftsin 1977 and 1989. Prog. Ocean., Vol. 47, no. 2-4, 103-145.

Haynes, E.G., and Ingell, G., 1983. Effect of temperature on rate of embryonic development ofwalleye pollock (Theragra chalcogramma) // Fisheries Bulletin. 81: 890-894.

Hiatt, T. R. Felthoven and J. Terry. 2002. Stock Assessment and Fishery Evaluation Report forthe Groundfish Fisheries of the Gulf of Alaska and Bering Sea/Aleutian Island Areas:Appendix D Economic Status of the Groundfish Fisheries of Alaska, 2001. Alaska FisheriesScience Center, National Marine Fisheries Service, National Oceanic and AtmosphericAdministration, Seattle. 119 pp.

Hollowed, A.B., and Wooster, W.S. 1992. Variability of winter ocean conditions and strong yearclasses of Northeast Pacific groundfish. ICES Mar. Sci. Symp. Vol. 195: 433-444.

Hollowed, A.B., and Wooster, W.S. 1995. Decadal-scale variations in the eastern subarcticPacific: II. Response of Northeast Pacific fish stocks. In: Beamish, R.J. (ed.), Climatechange and northern fish populations, Canadian Special Publication in Fisheries and AquaticScience, Vol. 121, pp. 373-385.

Hollowed, A.B., Hare, S.R., and Wooster, W.S. 2001. Pacific Basin climate variability andpatterns of Northeast Pacific marine fish production // Prog. Oceanogr. 49 (1-4): 257-282.

Huisman, J., and Sommeijer, B. 2002. Maximal sustainable sinking velocity of phytoplankton //Mar. Ecol. Progr. Series. 244: 39-48.

Hunt, G.L., Stabeno, P., Walters, G., Sinclair, E., Brodeur, R.D., Napp, J.M., and Bond, N.A.2002.Climate change and control of the southeastern Bering Sea pelagic ecosystem // DeepSea Res. Part II. 49 (26): 5821-5853.

Huntley, M.E., and Lopez, M.D.G. 1992. Temperature-dependent production of marine copepods:a global analysis // American Naturalist. 140: 201-242.

Incze, L.S., and Paul, A.J. 1983. Grazing and predation as related to energy needs of stage I zoeaeof the Tanner crab Chinoectes bairdi (Brachyura, Majidae) // Biol. Bull. 165: 197-208.

Jaenicke, M.J., Mathisen, O.A., and Radchenko, V.I. 1998. Fluctuations in abundance of pinksalmon (Oncorhynchus gorbuscha) in the North Pacific Ocean // N. Pac. Anadr. Fish Comm.Bull. No. 1: 496-502.

Page 139: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–139 DO NOT CIRCULATE OR CITE

Kocan, R. and P. Hershberger, and J. Winton 2001. Effects of Ichthyophonus on survival andreproductive success of Yukon River Chinook salmon: Final Report for Study 01-200, forthe US Fish and Wildlife Service, Subsistence Management, Fisheries Resource MonitoringProgram. University of Washington, Seattle. WA.

Latif, M., and Barnett, T. 1994. Causes of decadal climate variability over the North Pacific andNorth America // Science. 266, 634-637.

Link, M.R., M.L. Hartley, S.A. Miller, B. Waldrop, J. Wilen, J. Barnett. 2003. An analysis ofoptions to restructure the Bristol Bay salmon fishery. Prepared for Bristol Bay EconomicDevelopment Corporation and the Joint Legislative Salmon Industry Task Force.Anchorage. 104 pp.

Livingston, P.A., Low, L-L., and Marasco, R.J. 1999. In: Tang, Q., and Sherman, K. (eds.). LargeMarine Ecosystems of the Pacific Rim - Assessment, Sustainability, and Management.Blackwell Science. P. 140-162.

Luchin, V.A., Saveliyev, A.V., and Radchenko, V.I. 1998. Long-periodical climatic waves in thewestern Bering Sea ecosystem. In: Climate and interannual variability in the atmosphere –land – sea system in the American – Asian sector of the Arctic. Kozlov, V.F. (ed.).Proceedings of the Arctic regional center. V. 1. Vladivostok: Far-Eastern State UniversityPress. P. 31-42 (in Russian).

Luchin, V.A., Semiletov, I.P., and Weller, G.E. 2002. Changes in the Bering Sea region:atmosphere – ice – water system in the second half of the twentieth century // Progr.Oceanogr. 55 (1-2): 23-44.

Mantua, N. J. 2002. Large Scale Climate Variability and the carrying capacity of Alaska’soceans and watersheds. The status of Alaska’s Oceans and watersheds 2002.

Mantua, N.J., Hare, S.R., Wallace, J.M, and Francis, R.C. 1997. Pacific decadal climateoscillation with impacts on salmon production // Bull. Am. Meteor. Soc. 78: 1069-1079.

Miles, E., S. Gibbs, D. Fluharty, C. Dawson, and D. Teeter. 1982a. The Management of MarineRegions: The North Pacific, California University Press, Berkeley. 656 pp.

Miles, E., J.Sherman, D. Fluharty, S. Gibbs. 1982. Atlas of Marine Use: The North Pacific.California University Press, Berkeley. 121 pp.

Napazakov, V.V., Chuchukalo, V.I., Kuznetsova, N.A., Radchenko, V.I., Slabinskii A.M., andNadtochii, V.A. 2001. Feeding and some features of ecology of Gadidae fish in the westernpart of Bering Sea in the summer – autumn season // Izvestia TINRO (The TINROTransactions). 128: 907-928 (in Russian).

National Marine Fisheries Service 2003. Fisheries of the United States, 2002. Washington, D.C.National Marine Fisheries Service 2003b. Alaska Groundfish Fisheries: Draft Programmatic

Supplemental Environmental Impact Statement. Vol. V. App. A. Juneau, AK. Pp. 484.National Research Council. 1996. The Bering Sea Ecosystem. Committee on the Bering Sea

Ecosystem, Polar Research Board, Commission on Geosciences, environment and resources,National Research Council. National Academy Press, Washington, D.C. 1996.

National Research Council. 2003. Decline of the Steller Sea Lion in Alaskan Waters: UntanglingFood Webs and Fishing Nets. National Academy Press. Washington D.C. 1999.

Natural Resources Consultants, Inc. 1999. Status of Washington-Based commercial fisheries andthe fleets’ future utilization of Fishermen’s Terminal. Port of Seattle. Seattle. Pp. 195.

Naumenko, N.I. 2001. Biology and fisheries on marine herring of the far-eastern seas.Petropavlovsk-Kamchatsky: Kamchatsky Publ. House. 330 pp. (In Russian).

Page 140: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–140 DO NOT CIRCULATE OR CITE

Naumenko, N.I., Balykin, P.A., Naumenko, E.A., and Shaginyan, E.R. 1990. Long-term changesin pelagic fish community in the western Bering Sea // Izvestia TINRO (The TINROTransactions). 111: 49 – 57 (In Russian).

Naumenko, N.I., Balykin, P.A., and Naumenko, E.A. 2001. Long-term fluctuations in the pelagiccommunity in the western Bering Sea. In: Abstracts of PICES Tenth Annual Meeting,Victoria, BC, Canada. pp. 125.

Netboy, A. 1974. The salmon: thier fight for survival. Houghton Mifflin, Boston. MA.Niebauer, H.J., Bond, N.A., Yakunin, L.P., and Plotnikov, V.V. 1999. An update on the

climatology and sea ice of the Bering Sea. In: T.R. Loughlin and K. Ohtani (eds.). Dynamicsof the Bering Sea. Fairbanks: Alaska Sea Grant College Program. AK-SG-99-03. P. 29-59.

Nishiyama, T., Hirano, K., and Haryu, T. 1986. The early life history and feeding habits of larvalwalleye pollock in the southeast Bering Sea // Bull. INPFC. 45: 177 - 227.

NPAFC 2001. North Pacific Anadromous Fish Commission. 2001. Annual Report. Vancouver,BC. Canada.

North Pacific Fishery Management Council. 2002. Stock Assessment Fishery Evaluation.Anchorage, AK.

North Pacific Fishery Management Council. 2003a. Discussion paper on Western AlaskaCommunity Development Quota (CDQ) Program. NPFMC.

North Pacific Fishery Management Council. 2003b. Preliminary draft Environmental ImpactStatement for Essential Fish Habitat identification and conservation in Alaska. NPFMC.Anchorage, AK. Vol. I,II.

Ohtani, K., and Azumaya, T. 1995. Influence of interannual changes in ocean conditions on theabundance of walleye pollock in the eastern Bering Sea. In: Beamish, R.J. (Ed), Climatechange and northern fish populations, Canadian Special Publication in Fisheries and AquaticScience, Vol. 121, pp. 87-95.p.

Overland, J.E. 1981. Marine Climatology of the Bering Sea. In: Hood, D.W., and Calder, J.A.(eds.). The eastern Bering Sea shelf: oceanography and resources. Vol. 1. NOAA.Distributed by the University of Washington Press, Seattle. P. 15-22.

Overland, J.E., Adams, J.M., and Bond, N.A. 1999. Decadal variability of the Aleutian low and itsrelation to the high-latitude circulation // Journal of Climate. 12 (5): 1542-1548.

Polovina, J.J., Mitchum, G.T., and Evans, G.T. 1995. Decadal and basin-scale variation in mixedlayer depth and the impact on biological production in the Central and North Pacific, 1960-88 // Deep-Sea Research, Vol. 42, No. 10. P. 1701-1716.

Pautzke, C. 1997. Russian Far East Fisheries Management – North Pacific Fishery ManagementCouncil Report to Congress. Anchorage, AK. Pp. 68.

Pereyra, W.T., J.E. Reeves, and R.A. Bakala. 1976. Demersal Fish and Shellfish Resources of theEastern Bering Sea in the Baseline Year 1975. Northwest Fisheries Center, National MarineFisheries Service. Seattle. WA.

Quinn, T.J. II, and Niebauer, H.J. 1995 Relation of eastern Bering Sea walleye pollock (Theragrachalcogramma) recruitment to environmental and oceanographic variables. In: Beamish,R.J. (ed.), Climate change and northern fish populations, Canadian Special Publication inFisheries and Aquatic Science, Vol. 121, pp. 497-507.

Radchenko, V.I. 1992. The role of squids in the Bering Sea pelagic ecosystem // Okeanologiya(Oceanology). 32 (6): 1093-1101 (in Russian).

Page 141: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–141 DO NOT CIRCULATE OR CITE

Radchenko, V.I. 1993. Long-term variability in the Bering Sea surface geostrophic circulation andits possible influence on the pelagic fish community. In: Abstracts of PICES Second AnnualMeeting, Seattle, WA, pp. 29-30.

Radchenko, V.I., Khen, G.V., and Slabinsky, A.M. 2001. Cooling in the western Bering Sea in1999: quick propagation of LaNina signal or compensatory processes effect? // Progr.Oceanogr. 49 (1-4): 407-422.

Ray, G.C. and J. McCormick-Ray. 2004. Coastal Marine Conservation: Science and Policy.Blackwell Publishing. Malden, MA.

Rosenkranz, G.E., Tyler, A.V., and Kruse G.H. 2001. Effects of water temperature and wind onyear-class success of Tanner crabs In Bristol Bay, Alaska // Fish. Oceanogr. 10 (1): 1-12.

Sapozhnikov, V.V., Arzhanova, N.V., and Zubarevitch, V.V. 1993. Estimation of primaryproduction in the ecosystem of western Bering Sea // Russian J. Aquat. Ecol. 2 (1), 23-34.

Schumacher, J.D. 1999. Regime shift theory: A review of changing environmental conditions inthe Bering Sea and eastern North Pacific Ocean. Proc. of the Fifth North Pacific RimFisheries Conference, December 1-3, 1999. Anchorage, Alaska. 19 pp.

Shuntov, V.P. 2001. Biology of far-eastern seas of Russia. Vladivostok: TINRO-Center. V. 1. 580pp. (In Russian).

Shuntov, V.P., and Vasilkov, V.P. 1982. Epochs of atmospheric circulation and cyclicreoccurrence of abundance dynamics of Far-Eastern and Californian sardine // VoprosyIkhthiologii (Journal of Ichthyology). 22 (1): 187-199 (in Russian).

Shuntov, V.P., and Radchenko, V.I. 1999. Summary of TINRO Ecosystem Investigations in theBering Sea. In: T.R. Loughlin and K. Otani (Eds.). Dynamics of the Bering Sea. Fairbanks:Alaska Sea Grant College Program. AK-SG-99-03. P. 771-776.

Shuntov, V.P., Volkov, A.F., Temnykh, O.S., and Dulepova, E.P. 1993. Walleye pollock in thefar-eastern seas ecosystems. Vladivostok: TINRO. 426 pp. (In Russian).

Shuntov, V.P., Radchenko, V.I., Dulepova, E.P., and Temnykh, O.S. 1997. Biological resourcesof the Far Eastern Russian economic zone: structure of pelagic and bottom communities, up-to-date status, tendencies of long-term dynamics // Izvestia TINRO (The TINROTransactions). 122: 3-15 (in Russian).

Shuntov, V.P., Dulepova, E.P., and Volvenko, I.V. 2002. Up-to-date status and long-termdynamics of biological resources in the Far Eastern economic zone of Russia // IzvestiyaTINRO (TINRO Transactions). 130: 3-11 (in Russian).

Smith, S.L., and Vidal, J. 1986. Variations in the distribution, abundance, and development ofcopepods in the southeastern Bering Sea in 1980 and 1981 // Cont. Shelf. Res. 5: 215-239.

Sogard, S.M., and Olla, B.L. 1998. Contrasting behavioral responses to cold temperatures by twomarine fish species during their pelagic juvenile interval // Envir. Biol. of Fishes. 53: 405-412.

Somerton, D.A., and Munro, P. 2001. Bridle efficiency of a survey trawl for flatfish // Fish. Bull.99: 641-652.

Stabeno, P.J., Bond, N.A., Kachel, N.B., Salo, S.A., Schumacher, J.D. 2001. On the temporalvariability of the physical environment over the south-eastern Bering Sea // Fish. Oceanogr.10 (1): 81-98.

Stepanenko, M.A. 2001. The state of stock, interannual variability of recruitment, and fisheries ofthe eastern Bering Sea pollock in 1980s – 1990s // Izvestiya TINRO (TINRO Transactions).128: 145-152 (in Russian).

Page 142: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–142 DO NOT CIRCULATE OR CITE

Strickland, R. M., and T. Sibley. 1984. Projected effects of CO2 induced climate change on theAlaska walleye pollock (Theragra chalcogramma) fishery in the Eastern Bering Sea and theGulf of Alaska. University of Washington, Seattle, Wa. FRI-UW-8411. 112p.

Sugimoto, T., and Tadokoro, K. 1997. Interannual - interdecadal variations in zooplanktonbiomass, chlorophyll concentration and physical environment in the subarctic Pacific andBering Sea // Fish. Oceanogr. 6 (2): 74-93.

The Bering Sea ecosystem. 1996. Committee on the Bering Sea ecosystem. Polar Research Board.Commission on Geosciences, Environment and Resources. National Research Council(eds.). Washington, D.C.: National Academy Press. 307 pp.

Trenberth, K.E. 1990. Recent observed interdecadal climate changes in the northern hemisphere /Bulletin of the American Meteorological Society. 71: 988-993.

Van Loon, H., and Shea, D.J. 1999. A probable signal of the 11-year solar cycle in thetroposphere of the northern hemisphere // Geophys. Res. Lets. 26: 2893-2896.

Velegjanin. A.N. 1999. International legal aspects of the Russian Far East seafood trade with theNorth Pacific Rim Countries. Pp 189-198 In. V.M. Kaczynski and D.L. Fluharty, eds.Impacts of populations and markets on the sustainability of ocean and coastal resources:perspectives of developing and transition economies of the North Pacific. Collection ofPapers from the International Conference June 3-4, 1999. School of Marine Affairs.University of Washington. Seattle. 277 pp.

Vinogradov, M.E., and Shushkina, E.A. 1987. Functioning of pelagic communities in the oceanicepipelagic layer. Moscow: Nauka Publ. House. 240 pp. (In Russian).

Welch, D.W., Chigirinsky, A.I., and Ishida, Y. 1995. Upper thermal limits on the oceanicdistribution of Pacific salmon in the spring // Can. J. Fish. Aquat. Sci. 52: 489-503.

Wespestad, V.G. 1993. The status of the Bering Sea pollock and the effect of “donut hole” fishery// Fisheries. 18 (3): 18-24.

Wespestad, V.G., Fritz, L.W., Ingraham, W.J., and Megrey B.A. 2000. On relationships betweencannibalism, climate variability, physical transport and recruitment success of Bering Seawalleye pollock Theragra chalcogramma // ICES J. Mar. Sci. 57: 272-279.

Wilderbuer, T.K., and Nichol, D. 2001. Yellowfin sole. In: Stock Assessment and FisheryEvaluation Document for Groundfish Resources in the Bering Sea/Aleutian Islands Regionas projected for 2003. North Pacific Fishery Management Council, P. O. Box 103136,Anchorage, Ak 99510. P. 207-253.

Wilderbuer, T.K., Hollowed, A.B., Ingraham Jr., W.J., Spencer, P.D., Conners, M.E., Bond, N.A.,and Walters, G.E. 2002. Flatfish recruitment response to decadal climatic variability andocean conditions in the eastern Bering Sea // Progr. Oceanogr. 55 (1-2): 235-247.

Witherell, D., C. Pautzke and D. Fluharty. 2000. An ecosystem-based approach for Alaskagroundfish management. ICES Journal of Marine Science 57, 771-777.

Williams, E.H., and Quinn II, T.J. 2000. Pacific Herring, Clupea pallasi, recruitment in the BeringSea and north-east Pacific Ocean, II: relationships to environmental variables andimplications for forecasting // Fish. Oceanogr. 9 (4): 300-315.

Wooster, W.S., and Hollowed, A.B. 1995. Decadal-scale variations in the eastern subarcticPacific: I. Winter ocean conditions. In: Beamish, R.J. (ed.), Climate change and northernfish populations, Canadian Special Publication in Fisheries and Aquatic Science, Vol. 121,pp. 81-85.

Wyllie-Echeverria, T., 1995. Sea-ice conditions and the distribution of walleye pollock (Theragrachalcogramma) on the Bering and Chukchi shelf. In: Beamish, R.J. (Ed), Climate change

Page 143: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–143 DO NOT CIRCULATE OR CITE

and northern fish populations, Canadian Special Publication in Fisheries and AquaticScience, Vol. 121, pp. 131-136.

Wyllie-Echeverria, T., and Ohtani, K. 1999. Seasonal sea ice variability and the Bering Seaecosystem. In: T.R. Loughlin and K. Ohtani (eds.). Dynamics of the Bering Sea. Fairbanks:Alaska Sea Grant College Program. AK-SG-99-03. P. 435-451.

Zaika, V.E. 1983. Comparative hydrobiont productivity. Kiev: Naukova Dumka Publ. House. 203pp. (In Russian).

Zebdi, A. and Collie, J.S. 1995. Effect of climate on herring (Clupea pallasi) population dynamicsin the Northeast Pacific Ocean. In: Beamish, R.J. (Ed), Climate change and northern fishpopulations, Canadian Special Publication in Fisheries and Aquatic Science, Vol. 121, pp.277-290.

Zheng, J., and Kruse, G. H. 2000. Recruitment patterns of Alaska crabs in relation to decadalshifts in climate and physical oceanography // ICES J. Mar. Sci. 57: 438-451.

Zilanov. V.K. 1999. Marine living resources of the Russian Far East in the time of change:impacts of population, markets, reforms, and climate. Pp. 97-111. In. V.M. Kaczynski andD.L. Fluharty, eds. Impacts of populations and markets on the sustainability of ocean andcoastal resources: perspectives of developing and transition economies of the North Pacific.Collection of Papers from the International Conference June 3-4, 1999. School of MarineAffairs. University of Washington. Seattle. 277 pp.

Page 144: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–144 DO NOT CIRCULATE OR CITE

12.6 Synthesis and Key Findings

It has become evident from modeling experiments that it is by no means easy to predict changesof climate in response to forces, which can and have been measured and even monitored on aregular basis for considerable length of time and on which the models are built. The main reasonfor this is that major natural events take place on far longer time scales than decades or evencenturies and the history of regular monitoring of likely crucial forcing events is not all that long.Furthermore, current climate models do not contain scenarios on ocean temperatures, watermassmixing, upwellings, and other relevant ocean variables such as primary and secondary production,neither globally nor regionally. In consequence it is in reality not possible to accurately evaluatethe effects of climate change on marine fish stocks and eventual socioeconomic consequences ofthat for fisheries in Arctic regions.However, the above statement should not be taken as a categorical declaration that attempts totake on the difficult task of predicting reactions of marine stocks in or near the Arctic to climatechange should not be attempted. On the contrary, the scientific community should rise to thechallenge and base initial studies on past records of apparent interactions, however imperfect andinconclusive they may seem. It is on such bases – and such bases only – that effective futureresearch can and should be planned and carried out.Commercial fisheries in Arctic regions are based on a number of species belonging to ecosystemcomplexes. The dynamics of these ecosystems are poorly understood in many cases. This impartsa significant additional degree of uncertainty to evaluating the response of individual species andstocks to climate change. Indeed, it has historically proven difficult to identify the relativeimportance of fishing and the environment on changes in fish populations and biology. Moreover,current fish populations are different in abundance and biology from those in the past due toanthropogenic effects (exploitation rates). As a result it is uncertain whether current populationswill respond to climate changes as they may have in the past.On balance it appears likely that a moderate warming will improve the conditions for some of themost important commercial fish stocks, as well as for aquaculture, in Arctic regions, cfr. chapter 8.This will likely take place through more primary production due to less ice coverage and a moreextensive habitat areas for sub-Arctic species like cod and herring. There is also some chance thatglobal warming will induce an ecosystem regime shift in some areas with a different speciescomposition. Changing environmental conditions are also likely to be deleterious for some speciesand beneficial for others. Therefore, relative population sizes, rates of fish growth andgeographical distribution of fish stocks will be altered. Adjustments will then have to be made inthe Arctic commercial fisheries. Unless there is a dramatic climatic change over a very short timeperiod, these adjustments are likely to be relatively minor and will probably not entail majoreconomic and social costs.The total effect of climate change on fish stocks is probably going to be of a lesser magnitude thanthe effects of fisheries management. The significant factor in determining the future of fisheries issound resource management practices, which in large part depends upon the properties andeffectiveness of resource management regimes. In all Arctic countries there are currently efforts toimplement management strategies based on precautionary approaches where more emphasis isplaced on taking risk and uncertainty into account in decision-making. In addition, ecosystembased approaches to the management of living marine resources, where also effects of climate

Page 145: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–145 DO NOT CIRCULATE OR CITE

change are being considered. Such on-going adjustments to management regimes is likely toenhance ability of societies to adapt to the effects of climate change.

The economic and social impacts of altered environmental conditions depend crucially on theability of the relevant social structures, not the least the fisheries management system, to generatethe necessary adaptations to the changes. It is unlikely that the impact of the global warming in the21st Century signalled by the scenarios used here on Arctic fisheries will have a significant long-term economic or social impacts on a national scale. Certain regions in the Arctic, i.e. thoseheavily dependent on fisheries or marine mammals and birds in direct competition with a fisherymay, however, be radically affected. Local communities in the north are exposed to a number offorces of change. Economic marginalization, depopulation, globalisation-related factors, andpublic policies in the different countries are most likely going to have a stronger impact on thefuture development of northern communities than climate change, at least in the next few decades.

In this section we have considered the possible effect of global warming on four majorecosystems, i.e, the Northeast Atlantic (Barents Sea), the central North Atlantic(Iceland/Greenland), Northeast Canada (Newfoundand/Labrador) and the North Pacific (BeringSea). There are substantial differences between these regions in that the Barents Sea and Icelandicwaters are of a sub-Arctic/temperate type, while Arctic influence is much greater in Greenlandwaters, NE-Canada and the Bering Sea. It follows, therefore, that climate warming need not affectall of these areas in the same or similar manner. Furthermore, the length of useful time series onhistorical environmental variability and associated changes in the hydrobiological conditions, fishabundance and migrations varies greatly from one region to another. And, finally, there aredifferences in species interactions and variable fishing pressure, which have to be considered.

Due to heavy fishing and stock depletions, the Barents Sea, Icelandic waters and probably also theBering Sea could, through more efficient management, yield larger catches of many fish stocks.For that to happen research must be increased, and more cautious management strategies must beimplemented and complied with. However, a moderate warming could enhance the processes ofrebuilding stocks and, furthermore, generally result in higher sustainable yields of most stocks,i.a. through enlarged distribution areas and increased availability of food in general. On the otherhand, warming could also cause fish stocks to change their migratory range and area ofdistribution. That could, as history has demonstrated, trigger conflict among nations overdistribution of fishing opportunities and require tough negotiations to arrive at viable solutions asregards international cooperation in fisheries management.

Greenland and NE-Canadian waters are very different cases. These are much more Arctic innature and e.g. Greenlandic waters appear not to be able to support more temperate species likecod and herring except during warm period. For Greenland, there are examples from the 20th

Century that practically prove this point. No cod in the first two and a half decades, a large localself-sustaining cod stock from 1930 and until the late 1960s, apparently initiated to a large degreeby larval and 0-group drift from Iceland. With a climatic status quo nothing much can be expectedto change at Greenland. On the other hand, a ‘moderate warming’ like that during the period 1920until the late 1960s could bring about some dramatic changes in species composition, a scenariowhere cod would play the most important role by far. The NE-Canadian case is an extreme

Page 146: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–146 DO NOT CIRCULATE OR CITE

example of a situation where a stock of Atlantic cod (the so-called northern cod), which hadsustained a large fishery for at least two centuries, is suddenly gone. Opinion still differs withregard to how this could have happened, where some believe that inhospitable environment is themain cause while others hold the view that the stock appears to have simply been fished out. Inthis case it is worth noting that the Newfoundland-Labrador ecosystem is open, i.e. there are notemperature barriers, to the south and west. In earlier times of climatic adversities the northerncod could therefore have backed out and then repopulated its earlier distribution areas whenconditions improved again. In the present situation, however, the northern cod is so depleted that,due to its slow growth rate, it will almost certainly take decades to rebuild – even under theconditions of a warming climate.

An evaluation of what could happen should climate warming proceed beyond what here is definedas moderate (+ 1°-3°C), is not attempted. This is beyond the existing range of available data andwould therefore be of limited value. In general terms, however, it is likely that at least some of theecosystems would experience reductions in the present-day commercial stocks which, on the otherhand, might be replaced partially or in full by species from warmer waters.

12.7 Research Recommendations

Historical experience has taught us that the living resources of the sea are by no means unlimitedand must be harvested with caution. Although management of marine resources has improved inthe last decades, the situation of today still leaves a large room for improvement. In order to fillthat gap continued, improved and increased research is needed:

1) The present day monitoring of the physical and biological marine environment must becontinued and increased in many cases. Such basic research is often considered a burden,but is in reality an absolute prerequisite in order to understand biological processes .Fortunately, modern technology has opened up possibilities to automatize many timeconsuming tasks, earlier carried out on expensive research vessels. Thus it is now possibleto e.g. deploy buoys in strategic locations on land and in the sea for continually measuringmany of the host of parameters necessary in marine biological studies, in particular manyof those necessary for doing justice to the task of ACIA. Similarly, the monitoring ofcommercial stocks must continue, applying new technologies as they become available.Generally there is shortage of ship time for seaborne work. Many administrators(Governments) do not seem to understand this, nor the fact that, although computerizationhas enabled far more extensive and deeper analyses of existing and older data, people arestill needed to operate and program the computers.

2) Although modelling of marine processes in general and climate variability in particular isstill more or less in its infancy, such work is the key to improving the understanding ofmost of the scenarios dealt with in the ACIA report. In particular regional applications haveto be developed. Regional effects might differ substantially from what is considered to beaverage global effects. In order to relate physical effects on atmosphere and oceans tospecific biological systems (ecosystems), modelling the regional effects is howevernecessary. Current fisheries management models are based on general assumptions ofconstant environmental factors. The introduction of ecosystem based approaches for

Page 147: Ch12 - Fisheries ACIA Scientific Report - Fisheries ACIA Scientific Report External Review January 2004 12–2 DO NOT CIRCULATE OR CITE ocean climate, it is not known whether the present

Ch12 - Fisheries ACIA Scientific Report

External Review January 2004 12–147 DO NOT CIRCULATE OR CITE

fisheries management will require that both physical factors and biological factors outsidethe specific species are taken into account.

3) It turns out to be almost a very difficult task to estimate the economic consequences ofclimate change on the world fisheries or for any specific region. It is important to invest inthe development of better methods for examining the economic and social consequences ofclimate change, at the global and regional as well as the national and local level.