Fisheries and aquaculture in the Newfoundland and Labrador Seas, Northeastern Canada

From The Encyclopedia of Earth
Jump to: navigation, search


February 9, 2010, 3:54 pm
May 7, 2012, 1:11 pm

This is Section 13.4 of the Arctic Climate Impact Assessment
Lead Authors: Hjálmar Vilhjálmsson, 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,Thomas Wilderbuer

320px-Northeast Canadian ecosystem.JPG Fig. 13.24. Geographical overview of the Northeast Canadian ecosystem. The map also shows the NAFO statistical areas referred to in the text. (Source: ACIA)

Fisheries in ACIA Region 4 may be subdivided into those near the coast of Greenland, those near the coast of Canada, and those in the deep waters of Baffin Bay and Davis Strait between Greenland and Canada. The whole area is within the fisheries convention area of NAFO (North Atlantic Fisheries Organization) (Fig. 13.24) and the stocks are currently managed by the coastal state or by NAFO.

Along the northeast coast of Canada the study area extends southward to the central Grand Bank (46º N) in order to assess climate-driven impacts on marine ecosystems that are comparable to those considered for the northeast Atlantic (Section 13.2 (Fisheries and aquaculture in the Newfoundland and Labrador Seas, Northeastern Canada)) and around Iceland (Section 13.3 (Fisheries and aquaculture in the Newfoundland and Labrador Seas, Northeastern Canada)). This southward extension reflects the presence of the Labrador Current, which transports cold water southward from Davis Strait, the Canadian Archipelago, and Hudson Bay. The median southerly extent of sea ice (Sea ice in the Arctic) is on the northern Grand Bank at approximately 47º N[1] and bottom water temperatures on the northern Grand Bank are below 0ºC for long periods. The southerly extent of cold conditions is also indicated by the regular presence of polar cod along the northeast coast of Newfoundland and their occasional occurrence on the northern Grand Bank[2].

Fish have dominated the history of Newfoundland since the time of British colonization. The British interest in Newfoundland after its “discovery” during the Cabot voyage of 1497 was due to the incredibly large amounts of codfish. Exploitation of this fishery by the British reduced its dependence on Iceland for fish, a dependence that was creating difficulties. The French also saw the value of Newfoundland’s fishery, and possession of the island became an important part of the colonial wars of the 18th century[3]. As an inducement for France to enter the revolutionary war on the side of the American colonies, Benjamin Franklin offered a share of the Newfoundland fishery to the French as bounty once the war was won[4]. Indeed, until the late 1800s, when a cross-island railroad was built, fishing was Newfoundland’s only industry. There was then a series of diversification programs, which have continued in one form or another until the present day. Although in the early 1970's Newfoundland had the world’s largest hydroelectric plant (in Labrador), and despite many attempts to diversify the economy with both small-scale industries (e.g., cement production, knitting mills, a shoe factory, a chocolate factory) and numerous large-scale industries (e.g., the Churchill Falls hydroelectric station, a petroleum refinery, a third paper mill, iron mines in Labrador) in the twenty years following Confederation with Canada in 1949, none of these made any difference to the dominance of the fishery in Newfoundland[5]. However, what did change in Newfoundland with Confederation, and after revisions in Canadian federal/provincial intergovernmental arrangements, was the emergence of extremely large government, health, and education sectors which, as shares of GDP (Gross Domestic Product), eclipsed the fishery. By 1971, the fish and fish processing sectors accounted for less than 5% of Newfoundland GDP. By 2001, their contribution was 3.5%.

Ecosystem Essentials (13.4.1)

320px-Total catch of northeastern Canada.JPG Fig. 13.25. Total catch of selected species off northeastern Canada, 1960–2002 (data from NAFO STATLANT 21A online database www.nafo.int/activities/FRAMES/AcFrFish.html [6]). (Source: ACIA)

The ecosystem off northeastern Canada has been characterized by a relatively small number of species, a few of which have historically occurred in high abundance[7]. The dominant fodder fish has been capelin, with polar cod more prominent to the north and sand lance (Ammodytes dubius) more prominent to the south on the plateau of Grand Bank. Herring is found only in the bays and adjacent waters. These four species of planktivorous fish feed mainly on calanoid copepods and larger crustaceans, the latter predominantly hyperiid amphipods to the north and euphausiids to the south. The dominant piscivorous fish has been Atlantic cod, but Greenland halibut and American plaice (Hippoglossoides platessoides) have also been important. Snow crab and northern shrimp have been the dominant benthic crustaceans. The top predators are harp seals and hooded seals which migrate into the area from the north during late autumn and leave in spring. Other important predators include whales, most of which migrate into the area from the south during late spring and leave during autumn. The most important are humpback (Megaptera novaeangliae), fin, minke, sei, sperm, and pilot whales (Globicephala melaena). Additional immigrants from the north during the winter include many birds, such as thick-billed murre, northern fulmar (Fulmarus glacialis), and little auk. Additional immigrants from the south during summer include short-finned squid, fish such as mackerel and bluefin tuna (Thunnus thynnus), and birds such as greater shearwater (Puffinus gravis) and sooty shearwater (P. griseus).

The Labrador/Newfoundland ecosystem has experienced major changes since 1980. Atlantic cod and most other demersal fish, including species that were not targeted by commercial fishing, had declined to very low levels by the early 1990's[8]. In contrast, snow crab[9] and especially northern shrimp[10] increased considerably in abundance during the 1980's and 1990's and now support the most important fisheries in the area. Harp seals increased in abundance between the early 1970's and the mid-1990's[11]. Capelin have been found in much reduced quantities in offshore acoustic surveys since the early 1990s, but indices of capelin abundance in the inshore surveys have not shown similar declines, leaving the status of capelin uncertain and controversial[12]. Atlantic salmon, the major anadromous fish in the area, has declined in abundance, due in part to lower survival at sea[13].

The waters of eastern Newfoundland have been fished for centuries, primarily for Atlantic cod but with an increasing emphasis on other species during the latter half of the 20th century. These fisheries have undoubtedly had an influence on both the absolute abundance of some species and the abundance of species relative to one another. However, the role of the fisheries in structuring the ecosystem is often difficult to distinguish from the role of changes in the physical environment. The area cooled during the last three decades of the 20th century, with particularly cold periods in the early 1970's, early to mid-1980's, and early 1990's. This cooling, which was associated with an intensification of the positive phase of the North Atlantic Oscillation[14], may have played an important role in the dramatic decline in Atlantic cod and other demersal fish, and the increase in crustaceans, especially northern shrimp.

Fish Stocks and Fisheries (13.4.2)

Catches are from official NAFO statistics (as of February 2004) or from relevant assessment documents if there is a difference between the two[15]. Figure 2 provides an overview of developments in the main fisheries off Newfoundland and Labrador since 1960.

Atlantic Cod

320px-Atlantic cod catch 1960-2002.JPG Fig. 13.26. Catch of Atlantic cod, 1960–2002[16]. (Source: ACIA)

The distribution of Atlantic cod off Canada has historically been from the northern Labrador Shelf southward to beyond the limit of this study, although during the 1990's there were few cod off Labrador. Atlantic cod tends to occur on the continental shelf, but has been found at depths of at least 850 meters (m) on the upper slope off eastern Newfoundland[17].

The European fishery for Atlantic cod off eastern Newfoundland began in the late 15th century. For the first few centuries fishing was by hook and line, so the cod were exploited only from late spring to early autumn and only in shallow water along the coast and on the plateau of Grand Bank to the southeast of the island. There is evidence that local inshore overexploitation was occurring in the 19th century[18], but improvements in gear and an increase in the area fished tended to compensate for local reductions in catch rate. Annual landings increased through the 18th and 19th centuries to about 300,000 tons (t) in the early 20th century. The deep waters were refugia until the 1950's, when larger vessels with powered gurdys were introduced to exploit cod in deep coastal waters and European trawlers started to fish the deeper water on the banks. Landings increased dramatically in the 1960's as large numbers of trawlers located and exploited the overwintering aggregations on the edge of the Labrador Shelf and the Northeast Newfoundland Shelf. At the same time, the numbers of large cod in deep water near the coast of Newfoundland are thought to have declined quickly as the longliner fleet switched to synthetic gillnets. Catches peaked at 894,000 t in 1968, and then declined steadily to only 143,000 t in 1978. Following Canada’s declaration of a 200 nautical mile EEZ (Exclusive Economic Zone) in 1977, the stock recovered slightly and catches were between 230,000 t and 270,000 t for most of the 1980's. However, catches fell rapidly in the early 1990's as the stock declined to very low levels. A moratorium on directed fishing was declared in 1992 (Fig. 13.26). A small cod-directed inshore fishery was opened in 1998 but closed in 2003. 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[19], Lear and Parsons[20], Hutchings and Myers[21], Lear[22], Neis et al.[23], and Ferguson[24].

Greenland Halibut

320px-Greenland halibut catch 1960-2002.JPG Fig. 13.27. Catch of Greenland halibut, 1960–2002 (data from NAFO STATLANT 21A online database www.nafo.int/activities/FRAMES/AcFrFish.html; Darby et al[25]). (Source: ACIA)

Greenland halibut (also called Greenland turbot) is distributed off West Greenland from Cape Farewell northward to about 78º N and then southward off eastern Canada to beyond the limit of this study. It is a deepwater species, occurring at depths from about 200 m to at least 2,200 m off West Greenland[26].The history of the fishery is complicated by temporal and spatial variation in effort and catch by different fleets and by alleged underreporting of landings. For details of the fisheries, refer to Bowering and Brodie[27], Bowering and Nedreaas[28], and NAFO[29].

The fishery off eastern Newfoundland dates back to the mid-19th century[30]. Annual catches from longlines were less than 1,000 t until the early 1960's, when catches began to increase substantially. Landings from offshore trawlers, mainly from European countries, also increased after the mid-1960's. Catches in SA 2 + Div. 3KL fluctuated around 25,000 t to 35,000 t from the late 1960's to the early 1980's, after which there was a gradual decline to about 15,000 t in 1986. Landings increased dramatically in 1990 with the arrival of many non-Canadian trawlers that fished deep waters on the northern Grand Bank (see Fig. 13.24 for location). Catches over the next four years were high (estimated at between 55,000 and 75,000 t in 1991[31]), declined substantially in 1995 due to an international dispute, and increased again in the late 1990's under NAFO (Northwest Atlantic Fisheries Organization) quotas that maintained catches well below those of the early 1990's (Fig. 13.26).

The fishery to the north (NAFO SA 0), which has been conducted primarily with otter trawlers in the second half of the year[32], reported an average annual catch of 2,100 t between 1968 and 1989 (including a high of 100,00 t in 1972). Catches increased dramatically to 14,500t in 1990 with increased effort by Canada, but declined to about 4,000 t from 1994 onward. These landings came mainly from off southeastern Baffin Island. The fishery expanded even further north into Baffin Bay in the mid- to late 1990s[33]. This fishery, which extended to 73º N in 2002[34], has been limited by sea-ice cover in September through November.

Capelin

320px-Capelin catch 1960-2002.JPG Fig. 13.28. Catch of capelin, 1960–2002 (data from NAFO STATLANT 21A online database www.nafo.int/activities/FRAMES/AcFrFish.html). (Source: ACIA)

Before the start of a commercial offshore fishery in the early 1970's, capelin were fished on or near the spawning beaches. Annual catches, used for local consumption, may have reached 20,000 to 25,000 t[35]. Offshore catches by foreign fleets increased rapidly, peaking in 1976 at about 250,000 t, and then declined rapidly. This offshore fishery continued at a low level until 1992. Catches in the offshore fishery were taken at different times of the year in different areas. The spring fishery was dominated by large midwater trawlers operating in Div. 3L. During the autumn, the offshore fishery first occurred in Div. 2J, off the coast of Labrador, and gradually moved south into Div. 3K as the capelin migrated toward their overwintering area (see Fig. 13.24 for NAFO statistical areas). This fishery was also dominated by large midwater trawlers, which mostly took feeding capelin that would spawn the following year. During the late 1970's, as the foreign fishery declined, Canadian fishers began fishing mature capelin near the spawning beaches to supply the Japanese market for roe-bearing females. This fishery expanded rapidly, exhibited highest catches during the 1980's, and declined over the 1990's. Catches in the inshore fishery have generally been lower than from the offshore fishery. The total international catch of capelin off Newfoundland and Labrador from 1960 to 2002 is shown in Fig. 13.26.

Herring

Herring in the Newfoundland and Labrador area are at the northern extent of their distribution. Stocks are coastal in distribution and stock abundance is low compared to other stocks in the Atlantic. A peak catch of 30,000 t occurred in 1979, supported by strong year classes from the 1960's. Recruitment since the 1960's has been lower. Stock sizes in the late 1990's were less than 90,000t and annual catches less than 10,000 t[36].

Polar Cod

Polar cod is broadly distributed through the Arctic and in cold waters of adjacent seas. It occurs on the shelf from northern Labrador to eastern Newfoundland, with the average size of individuals and the size of aggregations decreasing from north to south[37]. There has been no directed fishery for polar cod off eastern Canada, but a small bycatch was reported in the Romanian capelin fishery in 1979[38], and it is likely that small quantities were also taken in other years and by other countries.

Northern Shrimp

320px-Northern shrimp catch 1960-2002.JPG Fig. 13.29. Catch of northern shrimp, 1960–2002 (data from NAFO STATLANT 21A online database www.nafo.int/activities/FRAMES/AcFrFish.html [39]). (Source: ACIA)

Northern shrimp is distributed off West Greenland from Cape Farewell northward to about 74º N and then southward off eastern Canada to beyond the limit of this study. The depth of highest concentration tends to vary from area to area but is generally between 200 to 600 m. A fishery with large trawlers began off northeastern Canada in the late 1970's[40]. For the first decade most of the catch was taken from two channels in the central and southern Labrador Shelf, but in the late 1980's there was an increase in effort and landings both to the south on the Northeast Newfoundland Shelf and to the north off northern Labrador. Catches increased above 25,000 t by the mid- 1990's. New survey technology introduced in 1995 indicated that commercial catches were very small relative to survey biomass, and quotas were increased considerably in the late 1990's. Total landings rose to more than 90,000 t by 2000 (Fig. 13.29). Much of the increase in catch from 1997 onward was from a new fleet of small (<100 feet) vessels that fished with bottom trawls mainly on the mid-shelf. In the 1990's fishing also expanded to Div. 3L[41].

Snow Crab

320px-Snow crab catch 1960-2002.JPG Fig. 13.30. Catch of snow crab, 1960–2002 (data from NAFO STATLANT 21A online database www.nafo.int/activities/FRAMES/AcFrFish.html [42]). (Source: ACIA)

Snow crab is distributed from the central Labrador Shelf at approximately 55º N southward off eastern Canada to beyond the limit of this study. The depth distribution extends from approximately 50 to 1,400 m, but most of the fishery occurs at 100 to 500m. The fishery off eastern Newfoundland began in the late 1960's as a small bycatch fishery, but soon expanded into a directed fishery with crab traps (pots) along most of the inshore areas of eastern Newfoundland (Div. 3KL)[43]. During the late 1970's and early 1980's there was an increase in effort and an expansion of fishing grounds. Catches in Div. 3KL reached almost 14,000 t in 1981, but then declined. In the mid-1980's there was expansion of the fishery to the area off southern Labrador (Div. 2J) and new entrants gained access to supplement declining incomes from the groundfish fisheries. The number of participants and the area fished expanded further during the 1990's, and total catches rose quickly, reaching almost 55,000 t in 1999. Quotas and landings were reduced for the next two years following concerns that the resource may have declined.

Commercial catch rates in Div. 3KL increased during the late 1970's to a peak in about 1981, declined to their lowest point by 1987, and then increased in the late 1980s and early 1990's to a level comparable to that in the early 1980's[44]. Catch rates remained high to the end of the 1990's, despite the substantial increase in fishing effort and landings (Fig. 13.30). This partly reflects an increase in the area fished, although there must also have been an increase in productivity.

Marine Mammals

320px-Harp and hooded seal harvest 1952-2002.JPG Fig. 13.31. Catch of harp and hooded seals, 1952–2002. The catches also include seals taken in the Gulf of St. Lawrence[45]. (Source: ACIA)

Harp seals summer in the Canadian Arctic or Greenland but winter and breed in Canadian Atlantic waters. There are two major breeding groups: the first breeding in the Gulf of St. Lawrence and the second breeding off southern Labrador and northeast Newfoundland[46]. The total population increased from less than 2 million in the early 1970's to more than 5 million in the mid-1990's[47]. The increase was largely due to a reduction in the hunt after 1982[48]. The population stabilized when the hunt was increased in the mid-1990's. Reported Canadian catches of harp seals include harvests off the coast of Newfoundland/Labrador (the “Front”) and in the Gulf of St. Lawrence. Seals caught in both areas belong to the same population: the Northwest Atlantic Harp Seals. The proportion of the population that occurs in the two areas varies among years, as does the relative number of seals caught in each area. Catches from both areas are combined in official statistics and so those presented here are combined “Front” and Gulf of St. Lawrence catches (Fig. 13.31).

Hooded seals are less abundant than harp seals. Whelping occurs on pack ice off northeast Newfoundland, in Davis Strait, and in the Gulf of St. Lawrence. Pups migrate into arctic waters and remain there as juveniles. Adults migrate south in the autumn and return to the Arctic in April[49]. The harvest of hooded seals (“Front” and Gulf of St. Lawrence combined) is shown in Fig. 13.31.

There has been no commercial whaling in the area since the late 1970's. Using north Atlantic population estimates, assumed growth rates, and an assumed proportion of the total population in the Newfoundland and Labrador area, Bundy et al.[50] estimated population abundances of 33,000 for humpback whales, 1,000 for fin whales, 5,000 for minke whales, 1,000 for sperm whales, 1,000 for sei whales, and 9,000 for pilot whales.

Past Climatic Variations and Their Impact on Commercial Stocks (13.4.3)

Atlantic Cod

250px-Catch of atlantic cod 1960-2000.JPG Fig. 13.32. Catch of Atlantic cod by NAFO statistical division, 1960–2000[51]. (Source: ACIA)

The severe decline in Atlantic cod in the Newfoundland and Labrador area seems to have occurred from north to south. On the northern and central Labrador shelf (Div. 2GH), catches of 60,000 to 90,000 t were reported for the period 1965 to 1969, but catches declined to less than 5,000 t for most of the 1970's and early 1980's, and to less than 1,000 t in the latter half of the 1980's (Fig. 13.32). There are no analyses of factors that contributed to the decline in this area.

In the area from southern Labrador to the northern Grand Bank, the Div. 2J+3KL stock (the so-called “northern cod”) collapsed in the 1970's in response to severe overfishing. The stock recovered slightly in the 1980's but collapsed to even lower levels in the late 1980's and early 1990's.There is controversy as to whether there was a rapid but progressive decline from the mid-1980's onward or a precipitous decline in the early 1990's[52]. Many studies[53] have concluded that the final stock collapse was entirely due to fishing activity (landed catch plus discards). However, several authors have pointed to ways in which the decline in water temperature and increase in sea-ice cover might have contributed to the collapse, either directly by reducing productivity[54] or indirectly by affecting distribution[55].

Despite many studies on this cod stock, there are few uncontested demonstrations 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 the northern limit of the species’ range in North America[56]. However, there have been conflicting reports of whether such a relationship can be detected[57]. Part of the problem is that recruitment is also positively influenced by the number and size of spawners in the population (the spawning stock biomass or SSB[58]). Both temperature and SSB declined from the 1960's to the 1990's, increasing the difficulty of demonstrating a temperature effect. A reported positive relationship between recruitment and salinity[59] was subsequently supported[60] and later rejected[61] as data for additional years became available.The negative effect of temperature on individual growth has been well documented[62]. Additional aspects of cod biology that changed during the early 1990's, possibly in response to changes in the physical environment, include a delay in arrival on traditional inshore fishing grounds in early summer[63], a concentration of distribution toward the shelf break in autumn[64], a move to deeper water in winter[65], and an apparent southward shift in distribution[66].

Of much interest is the possibility that an increase in natural mortality contributed to the rapid disappearance of cod in the early 1990's. The sharp decline in survey abundance indices occurred during a period of severe cold and extensive sea-ice cover. A considerable decline in the condition of the cod occurred at the same time, especially in the north[67]. Steep declines in abundance also occurred among other groundfish in the 1980's and 1990's[68], and while there have been some suggestions that these declines were caused by captures during fishing for cod and other species[69], there is no direct evidence of large removals. In the case of American plaice, a species studied in detail, Morgan et al.[70] demonstrated that the declines were too large to have resulted from fishing alone. The contribution of increased natural mortality to the decline in cod and other demersal fish in this area during the last two decades of the 20th century, and particularly during the early 1990's, remains unresolved[71].

The northern cod stock was still at a very low level a decade after the moratorium on directed fishing[72]. Recruitment to ages 0 to 2 remained very low, possibly due in part to a very small spawning stock biomass; juveniles in the offshore areas appeared to show very high mortality, possibly due in part to predation by harp and hooded seals; and a directed fishery during 1998 to 2002 targeted the inshore aggregations, resulting in increased mortality on the larger fish. The unquantified impacts of low spawning stock biomass, high predation, and fishing make it difficult to establish whether some aspect of ocean climate has had a role in impeding recovery.

Greenland Halibut

The status of Greenland halibut in the northwest Atlantic has been uncertain because the stock structure is still unclear, the fish have extensive ontogenetic migrations, there appear to have been shifts in distribution, the fisheries have undergone many changes in fleet composition and in areas and depths fished, and individual research surveys have only covered part of the distribution range. Nevertheless, evidence suggests that the biomass of Greenland halibut on the western side of the Labrador Sea declined substantially during the 1980's, with the decline off Baffin Island and northern Labrador (Div. 0B and 2GH) most pronounced in the first half of the decade and the decline off southern Labrador and eastern Newfoundland (Div. 2J3K) to the south most pronounced in the latter half of the 1980's and the early 1990's[72]. Evidence for a decline in biomass in Div. 2J3K is also seen in the declining success of the gillnet fishery in the 1980's. The history of the fish exploited during the 1990's by the new deep-water trawler fishery to the south on the northern Grand Bank (Div. 3L) is less clear. At least some of these fish may have migrated into the area from the shelf to the north[73], in which case the decline in Div. 2J3K was partly due to a southward shift in distribution.

Reasons for the declines in biomass and shift in distribution remain unclear. Bowering and Brodie[74] drew attention to the decline in water temperatures on the shelf in the early 1990's, but thought it unlikely that such a change would in itself have affected the distribution and abundance of Greenland halibut because this species occupies relatively deep water. Also, much of the shift in distribution must have occurred in the latter half of the 1980's, a period during which water temperatures were low but not as low as during the early 1990's.

Variability in the physical environment had no observed effect on either size at age[75] or maturity at size or age[76] between the late 1970's and mid-1990's.

Capelin

The relationship between capelin biology and the physical environment has been extensively studied in the Newfoundland and Labrador area. Of particular relevance to this assessment is the observation that many aspects of capelin biology changed during the 1990's and, initially, it appeared that these were the result of changes in water temperature. However, water temperatures in the latter half of the 1990's returned to normal while the biological changes exhibited by capelin did not revert to earlier patterns. There are many environmental variables that are linked to capelin biology which may be relevant in the event of global climate change and these are briefly described in the rest of this section.

Mean fish length of the mature population was smaller during the 1990's[77]. These small sizes have been attributed to smaller fish sizes at age with fewer older and more younger fish in the population. Condition (calculated as a relationship between length and weight and regarded as a measure of “wellbeing”) of capelin was generally higher in the 1980's than the 1990's. Condition was not related to temperature[78].

Spawning occurs most often on fine gravel and grain size and beach orientation have been shown to explain 61% of the variation in egg concentration among beaches[79]. Water temperature is also a determinant of capelin spawning. The lowest and highest recorded temperatures for beach spawning in Newfoundland are 3.5 and 11.9ºC, with beach spawning ceasing when temperatures exceed 12.0ºC[80]. Capelin eggs are very cold- and salinity-tolerant, surviving down to -5 ºC and in salinities from 3.4 to 34[81].The rate of egg development in the beach gravel is directly related to average incubation temperatures, which in turn are determined by water temperature, maximum and minimum air temperature, and hours of sunlight (Solar radiation)[82].

Some capelin that move close to spawning beaches eventually spawn in deeper water adjacent to beaches. This demersal spawning can occur simultaneously with intertidal spawning when temperatures are suitable as well as when water temperatures at the beach–water interface become too warm. Egg mortality among these demersal eggs has been observed to be higher. Reproductive success may have been lower during the 1990's because the water temperatures encountered when the capelin reached the spawning beaches would have increased the incidence of demersal spawning[83]. Historically, the spawning of capelin off Newfoundland beaches in June and July was a predictable event. In the early 1990's, spawning was later, and 80% of the variation in spawning time (1978–1994) was significantly and negatively related to mean fish size and sea temperatures experienced during gonadal maturation[84]. Spawning on Newfoundland beaches continued to be delayed through 2000 despite sea temperatures having returned to normal. However, mean lengths of capelin continued to be small.

There are historical reports of capelin occurring outside their normal distribution range. Unusual appearances in the Bay of Fundy and on the Flemish Cap were attributed to cooler water temperatures while occurrences in Ungava Bay coincided with warming trends[85].

In the early 1990's, capelin distribution occurred more to the south, centered on the northern Grand Banks. Originally attributed to the colder water temperatures[86], this shift within the normal distribution area continued through 2000. Because capelin did not return to their usual pattern of seasonal distribution as water temperatures increased, this suggests that factors other than water temperature were also operating. Outside their normal distribution area, capelin occurred on the Flemish Cap and eastern Scotian Shelf in the early 1990's and occasionally during earlier cold periods. Capelin continued to appear on the Flemish Cap and on the eastern Scotian Shelf through 2000. In this case, capelin appear to be gradually declining in abundance as the waters warm. For mature capelin offshore during spring, Shackell et al.[87] concluded that temperature was not a proximate cue during migration but that seasonal temperatures moderated offshore capelin migration patterns through the regulation of growth, maturation, food abundance, and distribution.

Capelin typically move up and disperse throughout the water column at night, descending and aggregating at greater depths during the day. However, during spring surveys throughout the 1990's they remained deeper in the water column and exhibited reduced vertical migration[88]. This change in vertical distribution was not related to the several factors tested, including temperature and predation, but may have been linked to feeding success[89].

Recruitment of beach-spawning capelin is partly determined by the frequency of onshore winds during larval residence in the beach gravel[91]. Capelin assessments have been especially problematic since the early 1990's, resulting in considerable uncertainty in the status of the stock. However, there is no evidence to indicate that exploitation has had a direct effect on population abundance[90], suggesting that any variations in abundance are due to environmental factors. It is not known whether some changes in biology such as condition and distribution have affected abundance, however, spawning time and increased demersal spawning may be contributing to poor survival.

Thus, 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 1990's, there is no clear indication of what external factor(s) has (have) influenced the changes. Earlier studies concluded that temperature was an important factor for some changes, but it now seems unlikely that temperature is the sole factor, given that water temperatures have returned to normal.There are suggestions that changes in food supply (zooplankton) may be affecting capelin biology but the exact mechanisms have not been identified.

Herring

Recruitment is positively related to warm overwintering water temperatures and high salinities[91]; these conditions seldom exist in this region and so, large year classes rarely occur.

Polar cod

The distribution of polar cod off eastern Newfoundland expanded to the south and east during the cold period of the early 1990's[92].

Northern shrimp

The shrimp resource off northeastern Canada has increased in density and expanded in distribution since the mid-1980s. There is no indication that increased catches have negatively affected the resource[93].

There is much support for the hypothesis that the increase in northern shrimp off northeastern Canada was, at least in part, a consequence of a reduced predation pressure by Atlantic cod and other groundfish[94]. Nevertheless, there is evidence that other factors were involved. For example, Lilly et al.[95] noted that the increase in shrimp density on the Northeast Newfoundland Shelf might have started in the early 1980's, a time when the biomass of Atlantic cod was increasing following its first collapse in the 1970's. Parsons D. and Colbourne[96] found that catch per unit effort in the shrimp fishery on the central Labrador Shelf was positively correlated with sea-ice cover six years earlier. They suggested that cold water or sea-ice cover itself was beneficial to the early life history stages of shrimp in that area.

Snow Crab

The increased productivity of snow crab in the 1990's may have been caused, at least in part, by the release in predation pressure from Atlantic cod and other demersal fish[97]. However, the relationships between Atlantic cod and snow crab have not yet been explored to the same extent as for Atlantic cod and northern shrimp. A preliminary examination of the influence of oceanographic conditions on snow crab productivity has shown a negative relationship between ocean temperature and lagged catch rates[98].This has been interpreted to indicate that cold conditions early in the life cycle are associated with the production of strong year classes of snow crab in this area.

Marine Mammals

Trends in [[population]s] of marine (Marine biodiversity) mammals over recent decades appear to be influenced mainly by the commercial harvest. As populations of harp seals have increased in abundance, changes in biological characteristics indicate that density-dependence may be operating[99]. Density-independent influences may also regulate harp seal populations. Harp seals whelp on sea ice and mortalities may vary according to sea-ice conditions in this critical period. 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 seal populations increased. It has been estimated that 74% (about 3 million t) of the total annual consumption by four species of seals in eastern Canada occurred off southern Labrador and Newfoundland[100]. Predation by harp seals has been implicated in the lack of recovery of the northern cod stock[101], and predation on cod by hooded seals may be large[102].

Aquaculture

Salmonid aquaculture does not occur in the ACIA (Arctic Climate Impact Assessment (ACIA)) (Arctic Climate Impact Assessment) part of Newfoundland because the water is too cold in winter. The main species cultured is blue mussel. Production of this species has grown over the last twenty years such that, in 2002, around 1,700t were raised in the whole of Newfoundland.

Possible Impact of Climate Change on Fish Stocks (13.4.3)

Two recent papers[103] discussed the possible influence of climate change on ecosystems and fisheries off eastern Canada. Frank et al.[104] predicted shifts in the ranges of several groundfish stocks because of redistribution of [[population]s] and changing recruitment patterns. Stocks at the southern limit of a species’ distribution should retract northward, whereas those near the northern limit should expand northward. Frank et al.[105] did not make predictions specifically for Labrador and eastern Newfoundland, but events during the decade following publication of their paper were in many respects opposite to these general predictions.The changes off Labrador and eastern Newfoundland were unprecedented and not predicted, and illustrate the uncertainty of predictions, even on a regional scale and in the relatively short term. Shuter et al.[106] had the advantage of witnessing the dramatic changes that occurred in the physical and biotic environment during the 1990's. They concluded that greenhouse gas accumulation will lead to a warmer, drier climate and, for the fisheries of Atlantic Canada, this will result in a “decrease in overall sustainable harvests for coastal and estuarine populations due to decreases in freshwater discharge and consequent declines in ecosystem productivity”. For fisheries in the Arctic, they predicted “increases in sustainable harvests for most fish populations due to increased ecosystem productivity, as shrinkage of ice cover permits greater nutrient recycling”.

As the relative importance of fishing and environment is difficult to determine for any species or group of species, it is not surprising that the importance attributed to each has varied for different studies. It is also not surprising, given the differences among species in the magnitude of fishery removals relative to stock size, that opinion favors fishing as the dominant factor for some species and environment for others. For demersal fish, there are many statements to the effect that declines were caused entirely by overfishing, but there is evidence that changes in oceanographic properties contributed to changes in distribution and declines in productivity. For crab and especially shrimp, it has been suggested that increases in biomass were simply a consequence of a release in predation pressure from Atlantic cod and perhaps other demersal fish, but again there is evidence that changes in oceanographic factors contributed to an increase in reproductive success. For capelin, most information supports the hypothesis that fishing had little impact on population dynamics, and that environmental factors were the primary determinant of stock size, well-being (growth and condition), distribution, and timing of migrations. For polar cod, fishing may be dismissed as a contributor to changes in distribution and biomass.

An important constraint on predicting changes in fish stocks and the fisheries that exploit them off Labrador and eastern Newfoundland is uncertainty about the direction and magnitude of change in important oceanographic variables. For surface air temperature, some model outputs project a cooling over the central North Atlantic, and it is not clear where the Labrador/ Newfoundland region lies within the gradation from significant warming in the high Arctic to cooling over the central North Atlantic. In addition, there is no model for downscaling the output of general circulation models to specifics of the Labrador/Newfoundland area. As an example of the importance of regional models, many large-scale models project an increase in air temperature over the Norwegian and Barents Seas, while simulations with one specific regional model[107] indicate that sea surface temperature in that area may decline in the next 20 years before increasing later in the century. Another concern is that natural variability in a specific region, such as the Labrador Shelf, may be greater than variability in the global mean[108]. Thus, a warming trend in shelf waters off Labrador and Newfoundland might be accompanied by substantial annual variability, such as was witnessed during the last three decades of the 20th century, and it is even possible that the amplitude of that variability could increase. For biota, 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 as important to know how the North Atlantic Oscillation will behave (especially the intensity and location of the Icelandic Low) as it is to know that global temperature will rise.

In the absence of region-specific information on likely future developments of climate, all predictions of climate-driven changes in the marine ecosystem off Newfoundland and Labrador can only be highly tentative. The following subsections describe the changes that seem most likely under three different scenarios: no change or even cooling of climate, moderate warming, and considerable warming.

No Change

As temperatures were generally below the long-term average during the ACIA (Arctic Climate Impact Assessment (ACIA)) baseline period (1981–2000), no change from present conditions or even a cooling are likely to favor the current balance of species in the system. This implies a predominance of commercial invertebrates like northern shrimp and snow crab and cold water species of fish such as Greenland halibut, polar cod, and capelin.

Moderate Warming

The moderate warming scenario (an increase of 1 to 3 ºC) assumes that there will be a gradual warming of the shelf waters off Labrador and Newfoundland. Using the events in West Greenland during the first half of the 20th century[109] as a spatial/temporal analogue, there is likely to be better recruitment success and northward expansion of Atlantic cod and some other demersal fish that live mainly on the shelf. Capelin is also likely to shift northward. If zooplankton abundance is enhanced by warmer water, capelin growth is likely to improve. It is possible that many existing capelin spawning beaches will disappear with the projected rise in sea level[110]. Depending on the increase in sea level, storm events, and the availability of glacial deposits, some beaches may move and new beaches be formed, while others may disappear completely. While beach-spawning capelin can adapt to spawning on suitable sediment in deeper water, survival of eggs and larvae appears to be adversely affected[111], suggesting that a rise in sea level is likely to result in reduced survival and recruitment for capelin. A warming of sea temperatures is likely to retard recruitment to snow crab and northern shrimp, so these species might experience gradual reductions in productivity. Thus, a gradual, moderate warming of sea temperature is likely to promote a change back to a cod–capelin system from the present system where snow crabs and northern shrimp are the major commercial species. In addition, both cod and capelin are also likely to become more prominent off central Labrador than they were during the 1980's.

A gradual warming of shelf waters is also likely to promote a shift of more southerly species into the area. For example, haddock is likely to become more abundant on the southern part of Grand Bank, and expand into the study area. Migrants from the south, such as short-finned squid, mackerel, and bluefin tuna, are likely to occur more regularly and in greater quantities than in the 1980's and early 1990's.

The simple scenario of a gradual change back to a cod–capelin system under moderate warming conditions is uncertain. This is because the influence of oceanographic variability in the past is still not clear, and because it is likely that the dynamics of some species are now dominated by a different suite of factors than was the case in the past. It is highly likely that the ecosystem off northeastern Canada changed substantially as a consequence of fishing during the first four centuries after the arrival of European fishers, changed even further with the increasingly intensive fishing of the 20th century, and has changed dramatically from the 1960's onward. The magnitude of these changes is such that it would be difficult to predict accurately the future state of this ecosystem even without the added complications of climate change. Thus, the system could remain in its current state, could revert to some semblance of an historic state (or at least the state of the early 1980's), or could evolve toward something previously unseen.

Changes in sea ice (see Tables 9.2 and 9.3) are likely to have a negative impact on harp seals, the most important marine (Marine biodiversity) mammal predator in the area. Sea-ice duration is projected to shorten and it is not known whether harp seals would be able to adjust their breeding time to accommodate this change. A decrease in sea-ice extent is unlikely to affect harp seals because they would probably shift their distribution with the sea ice. However, thinner sea ice may be deleterious, resulting in increased pup mortality. Increases in regional storm intensities (see Table 9.1) are likely to result in higher pup mortalities if such storms occur during the critical period shortly after birth[112]. Changes in seal abundance are likely to cascade through the ecosystem, since seals are important predators on many fodder fish and commercially important groundfish[113], and are thought by some to be important in impeding the recovery of cod[114] and thus maintaining the present balance within the ecosystem.

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

A reduction in sea-ice cover (see Tables 9.2 and 9.3) is also likely to negatively impact upon Greenland halibut fisheries that are conducted through fast ice. For example, a fishery that was developed in Cumberland Sound on Baffin Island in the late 1980's has developed into a locally important enterprise[115]. The fishery is conducted with longlines set through ice over deep (600–1,125 m) water, with the season extending in some years from mid-January to June. Since the mid-1990s, the season has been shorter, typically from early February to May[116]. To date, attempts at fishing during the open water season have not proved successful. The catches have been small and the fish appear dispersed. It is unclear whether the fish would be present in commercial concentrations in the winter/spring if sea ice were not present. Even if they were, the absence of sea ice would certainly affect the conduct of the fishery.

Considerable Warming

Since a warming of 4 to 7ºC is beyond any recorded experience in the Newfoundland–Labrador area, a meaningful discussion of the considerable warming scenario is not practicable. In very general terms, such a shift could favor cold-temperate species such as cod, improve conditions for more southern species such as haddock and herring, and even lead to the formation of demersally spawning stocks of capelin in addition to beach spawning stocks. However, there are likely to be other changes, such as a freshening of the surface layer due to freshwater from melting sea ice further north, which would be likely to reduce primary production in the area.

The Economic and Social Importance of Fisheries (13.4.5)

From an economy based primarily on the fishery, Newfoundland has, along with most of North America, moved to a service economy. By 1971, for instance, the fishing and fish processing sectors accounted for less than 5% of Newfoundland’s GDP, whereas the service sector accounted for more than half. Mining accounted for 11% and construction 18% (although that included construction 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 accounted for slightly more than 5%, service industries had grown to 68%, mining had fallen to less than 6%, and construction to 8%[117]. By 2001, the fishery harvesting and processing sectors accounted for only 3% of GDP, the service industries remained constant at 68%, construction had slipped to 4.7%, and conventional mining to 3%. Oil production, a new industry in Newfoundland, already accounted for 8.4% of the provincial GDP, with every prospect of growing (the 2001 data were from the Newfoundland Statistics Agency). Mining was also expected to see resurgence with the potential opening of a large nickel mine in Labrador.

While the fishery may not be of great importance to the overall Newfoundland economy, it continues to dominate completely the economy in rural areas, and perhaps even more importantly, its culture. After fifty years, there is still a daily Fisheries Broadcast on radio and when the Canadian Broadcasting Company decided to cancel the weekly television program Land and Sea, which often focuses on the fishery, public pressure forced the crown corporation to continue the program. With the fishery in deep trouble in 1989, the dominant newspaper in the province, The Evening Telegram, commented in an editorial entitled “Too Many Fishermen?” on 1 June 1989, that “Newfoundland’s fishery must eventually be expanded and diversified so that it can employ more people, not fewer...”

With the spectacular change in fisheries employment that accompanied the collapse of the northern cod stock, there has been a sharp reemphasis on economic diversification. The two areas paraded as holding the hope of the future are tourism and information technology[118]. Progress has been made in both areas (e.g., fishing vessels converted to tour boats for whale watching, and many bed and breakfast establishments), but has been uneven, and some government policies have been inconsistent with the promotion of these industries. For instance, despite its interest in developing tourism, the Newfoundland government decommissioned or privatized a substantial number of the parks in the province’s extensive parks system[119]. How many of those sold remain as parks is unknown.

The real problem with the emphasis on tourism and information technology is that it is happening at the time as it is happening throughout the world. Why should Newfoundland have an advantage over the rest of the world in either field? It is too early to tell whether this diversification will be successful.

The story of the Newfoundland fishery does not end with the collapse of the groundfishery, its catastrophic consequences for many families, and the serious pressures it placed on the government. Two critical sets of changes have occurred since 1992: (1) the fishery management process has evolved and (2) shellfish have replaced groundfish as the main components of fishery landings in Newfoundland.

Fishery management by restricting total allowable catches (TAC's) began in Newfoundland in the mid-1970's. Following a particularly dramatic race to the fish in 1981, the government imposed, at the industry’s request, enterprise allocations on the offshore groundfish fleet in 1982. These allocations were divisible and transferable in the year in which they were assigned. Government emphasized that these were rights to fish as opposed to property rights, which could be permanently sold. 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 have also been assigned to this inshore sector. These allocations are not transferable.The federal government has relinquished the licensing of fishers in favor of the provincial Professional Fish Harvester’s Certification Board. This board was established as part of a professionalization program and it licenses harvesters either as apprentices or in one of two professional classes. The federal government, in turn, in 1996 established “core” fishing enterprises for the inshore fisheries. Senior level professional fishers who met certain conditions were declared the heads of core fishing enterprises. Approximately 5,500 of these were established and the government claimed that no additional core licenses would be issued; the only way that fishers could obtain such status would be to buy out the core license of an existing core fisher. In an attempt to reduce the number of fishers, the federal government bought approximately 1,500 core licenses, claiming that these will not be reissued. The final major changes in the management system are that (1) species license fees (access fees) are no longer nominal, for most important species they are based on the anticipated gross income from the fishery; and (2) there is now an extensive system of public consultations before recommendations concerning total allowable catches are made to the Minister of Fisheries and Oceans. For a discussion of the current fishery management system, see Schrank and Skoda[120].

The value of landings in all Newfoundland fisheries in 1991 was Can$ 282,838,000. This fell during the first year of the moratorium to Can$ 194,745,000 and then doubled to Can$ 388,700,000 by 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 the value of landings was for cod, in 2000 46% was accounted for by crab and a further 30% by shrimp. With some cod stocks reopened on a limited basis for commercial fishing, cod accounted for 9% of the total landings in 2000, although this had earlier (in 1996) fallen to less than one-half of one percent.

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

Since all major species are under quota, total allowable catches for shrimp and crab have been increasing rapidly, along with the number of harvesting and processing licenses for shrimp and crab.The number of shrimp harvesting licenses rose from 19 in 1986 and 57 in 1991 to 438 in 2000. The numbers of crab licenses for those years were 274, 721, and 3,333[121]. There has been much controversy as to whether the old error of issuing too many licenses has occurred for crab. Even with the exodus from the province, in February 2003 there was still a 17.5% seasonally adjusted unemployment rate in the province (where the national figure was 7.4%) with continued pressure to open closed plants and increase licenses for crab fishers. While the number of crab fishing licenses has increased substantially, the increase in the number of crab processing plants has been modest.

Past Variations in the Fishing Industry and their Economic and Social Impacts (13.4.6)

The Evening Telegram editorial, referred to in the "The economic and social importance of fisheries" section, appeared within the context of a fishery that had long been in trouble. In 1967, a provincially financed report supported the trend away from a seasonal inshore fishery in Newfoundland toward a capital-intensive yearround offshore fishery. The report also noted that the “number of people dependent on the fishery should be reduced.”[122] This has been a recurring theme. In 1970, the federal fisheries department appealed to the Canadian cabinet to permit the department to establish regulations that would lead to a 50% reduction in the number of |Atlantic Canadian (meaning mainly Newfoundland) fishers.

The authority to effect these changes was denied[123]. The fishery faced repeated crises, was repeatedly studied, and the conclusion was repeatedly drawn that too many people were dependent on it. One study estimated that of the then 35,000 licensed fishers, only 6,000 could be supported unsubsidized by the fishery at a better than poverty-level income. The same study concluded that for every dollar of fish landed, there was a dollar of subsidy[124]. In 1976, with the extension of coastal states’ fishery jurisdiction to 200 nautical miles from shore, the two Canadian departments concerned (fisheries and regional economic expansion) both published reports stating that there was sufficient extra capacity in the industry that no significant employment benefits could be expected from the expanded jurisdiction[125]. Two years later, a provincial government report made a similar point by stating that the Canadianized fishery, when fully developed, could employ only 9,000 inshore fishers[126]. Yet, in response to popular pressure, both federal departments, as well as the provincial government, licensed and subsidized a tremendous expansion in the physical capacity of the industry: from 13,636 registered fishers in 1975 to 33,640 in 1980; from 9,517 registered inshore vessels in 1976 to 19,594 in 1980; from a fish freezing capacity of 181,000 t in 1974 to 467,000 t in 1980; from net fishers’s unemployment insurance benefits of US$ 30,724,000 in 1976/77 to US$ 66,060,000 in 1980/81; and from outstanding loans of the provincial Fisheries Loan Board (to finance inshore vessels) from US $36,869,000 in 1976/77 to US $78,558,000 in 1980/81[127]. By 1981 the expansion had stopped and the two federal departments agreed that no further expansion of fish processing facilities would be built with federal financing[128]. However, the damage had already occurred and, in the face of the anti-inflation recession of the early 1980's in the United States (where most of the Newfoundland fish production was sold), the market for Newfoundland fish products shrank dramatically and most Newfoundland fish processing companies faced bankruptcy.

As a result, the industry was financially, but not structurally, reorganized and the massive industrial closures implicit in bankruptcy were averted[129]. Yet, starting in 1982, inshore groundfish catches fell and after 1986 offshore catches followed. By 1992, the situation was so bad that a moratorium on the commercial catching of the formerly massive northern cod stock was put into place. Shortly after, nearly all Newfoundland groundfisheries were closed and the moratorium was extended to non-commercial fishing[130]. The closure of most of these groundfisheries continues, in whole or in part, to the present. The closure of the Newfoundland groundfisheries is reputed to have involved the largest mass layoff of labor in Canadian history. In social terms (due to the mass layoff), in biological terms (due to the decimation of the fish stock), and in governmental financial terms (due to billions of dollars spent on income maintenance for fishers and fish plant workers) the moratoria were disasters.

The response of government, industry, and the public to the moratoria indicates what might happen with climate change. Although the cause of the stock destruction in Newfoundland waters may be debated, the dramatic effect on the fish population is incontrovertible. Should significant changes in environmental conditions occur, and should these changes have substantial effects on commercial fish stocks, then the Newfoundland experience may provide a template for what might be expected to happen elsewhere. Moreover, the Newfoundland experience may also indicate the need for alternative policies.

The decline in the cod fisheries was better understood after the reports by Alverson[131] and Harris L.[132]. That major problems were developing in the groundfishery was no longer debatable. In 1990, in response to the decline of the fishery, the federal government introduced the Atlantic Fisheries Adjustment Program (AFAP)[133]. The emphasis was on the word “adjustment”. People were to be retired from the fishery, rural communities were to receive money to help them diversify their economies away from the fishery, and steps were to be taken to increase scientific understanding of the declining fish stocks. But only a few hundred fishers left the industry. With the shock of the total closure of the commercial northern cod fishery in July 1992, the federal government, anticipating that the fishery would revive in two years, created the Northern Cod Adjustment and Recovery Program (NCARP)[134]. Again there was an emphasis on people adjusting out of the fishery. This program called for early retirement of fishers, buybacks of fishing licenses, training of fishers for other trades, and income maintenance payments to fishers and fish plant workers. A third of the 9,000 northern cod fishers and half the 10,000 plant workers affected by the northern cod shutdown were expected to leave the fishery. In fact, only 1,436 took early retirement and 876 fishers sold their licenses. Fishers were not convinced that the shutdown would continue for long, and believed that the government would support them until the fishery reopened; the relatively uneducated, potential low end laborers did not see a need to leave an industry in which they were skilled and for which they were trained from an early age. One reason for low educational levels (until 1991, less than half Newfoundland’s adult population had completed high school) was the fishing tradition. Boys started fishing with their fathers at a young age and looked forward to leaving school as soon as possible to join the family fishing enterprise. Boys and girls with little interest in fish harvesting could work for life in the local fish plant, in jobs which mostly required little formal education.

However, the fish did not return after two years, and have still not returned a decade after the start of the moratorium. As NCARP was ending, a new adjustment program began. The Atlantic Groundfish Strategy (TAGS)[135] was to be a five-year program of income maintenance and adjustment (license buybacks and retirements) in which a 50% reduction in fishing capacity was anticipated. Again, there was very little movement of people out of the fishery. Their reluctance to abandon the fishery was for the same reasons as under the NCARP program.

As TAGS drew to a close at the end of the 1990's, the government took a harder line. The post-TAGS program did not resemble its predecessors: income maintenance was severely cut and many people were removed from the program. With government financial support gone, or going, and the fish still not returned, an exodus from the fishery finally occurred.

Between 1986 and 1991, the Newfoundland population stagnated, at least partially from a dramatic drop in family size. From the highest birth rate of Canadian provinces, by 1991 Newfoundland had the lowest. Also, there had always been modest migration out of the province. But in the five years between 1991 and 1996 (from the year after the start of AFAP to halfway through TAGS) the population actually fell by 3%. With the continuing moratorium and the change in government policy, the exodus increased significantly and between 1996 and 2001 there was a further drop of 7%. Census figures for 2001 are from Statistics Canada 2002[136], while those for 1991 and 1996 are from Statistics Canada 1999[137].

Even though such a population drop in a province over a decade is dramatic, this value of 10% actually hides 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 groundfish processing plant. Newfoundland groundfish operators had been operating under an Enterprise Allocation scheme since 1982. With the drop in fish stocks toward the end of the 1980's, enterprise allocations were cut and, in response, a number of fish plants closed. One of the first to close, in 1990, was that in Trepassey. The result was that a town with 1,375 inhabitants in 1991 had shrunk by more than 35% to 889 in 2001. Many rural communities in Newfoundland have seen population declines since 1990 of 15 to 30%[138].

Economic and Social Impacts of Climate Change: Possible Scenarios (13.4.7)

Climate change is likely to cause changes in the size of fish stocks. The effects are unlikely to be greater than the historical changes described in Past variations in the fishing industry and their economic and social impacts above. With human society, responses to impulses are not “natural” in the sense that a climatic change “causes” a human response. The human response is determined by the magnitude of the stimulus plus the political response of the society. In this sense, societal responses to climate change will not be qualitatively different from society’s responses to past changes. The political system will respond, and the details of that response are impossible to predict. But models exist from past experience. However, many of the federal and provincial interventions in the fishery since 1992 have appeared ill thought out, often unfair, and have raised controversy.

A recent case provides an illustration. The groundfish plant in Twillingate was once owned by the largest fish company in Newfoundland, Fishery Products International, Ltd., but had been sold to another operator in the mid-1980's[139]. With the northern cod moratorium in 1992, the plant was shut and remained shut until 2002 when it opened as a shellfish plant with more technologically sophisticated equipment than had been used for groundfish. The Marine Institute, a branch of Memorial University, introduced a course to teach fish plant workers to use the new equipment, charging more than CAN$ 400 per person. For unemployed people receiving (un)employment insurance the fee was paid by the federal government. Others had to pay for themselves. Most unemployed people without employment insurance could not afford the fee. Most working people, wanting higher paying jobs in the fish plant, or former fish plant workers wanting to return to the industry, would need to quit their jobs to take the course, unless they were granted time off, which is unlikely in unskilled trades. But if they quit their jobs and completed the course, there was no guarantee of a job in the Twillingate, or any other technologically advanced, fish plant[140]. Thus, every aspect of the long adjustment process which started with the decline of the groundfishery in the 1980's has been characterized by a deep sense of unfairness.

The Newfoundland experience shows that a “catastrophic” event concerning the fishery leads to severe adjustment problems, and that the adjustment period may be very long, but that it also raises new potential for a successful industry. The issues seem to be:

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

In terms of predicting the socio-economic effects of longterm climate change, this is one case where it is easier to prescribe than to predict. The Newfoundland experience has shown reactions to expanding fish [[population]s] and to shrinking fish populations. In neither case was the reaction, in terms of government action or political pressure, appropriate. During the expansion period of the late 1970's, the fishery expanded too much, with excessive and ultimately largely immobile labor and capital entering the industry. Whereas a properly managed fishery would have restricted the expansion of production factors, the expansion was almost without letup until stopped by a general economic crisis. It was understood at the time that employment expansion was an incorrect response. Should fish stocks off Newfoundland increase over the next 20 or 50 years, care should be taken (1) to restrict by government regulation the magnitude of any expansion of capital and labor in the fishery; or (2) to ensure that such economic incentives are in place that excessive growth does not occur; or (3) to combine the two.

Should fish stocks decrease over the next 20 or 50 years, then it should be clear from the start that endless subsidies will not be forthcoming. License buyouts, even generous license buyouts of core enterprises for instance, would help. While these payments are subsidies, they are limited in scope and time and would have the effect of permanently shrinking the factor base of the industry. In the 1990's, such a policy would have been much cheaper and much less stressful for the fishing families affected, than the offering of income maintenance payments.

A gradual warming of shelf waters is likely to lead to increased opportunity for aquaculture. Warmer temperatures and shorter periods of sea-ice cover are likely to enable mussel farming to be more productive. Warmer waters are also likely to promote the development of Atlantic cod farming. If inshore waters become sufficiently warm, it is likely to be possible to farm Atlantic salmon along the east coast of Newfoundland. This is presently impossible because water temperature in winter falls below the lethal temperature for salmon.

Ability to Cope with Change (13.4.8)

Climate change will affect all aspects of the fishery: the range of existing species, the relative [[population]s] of different species, and the economic circumstances of people who depend on the fishery for a living. How ready the economic and social systems of Newfoundland are to cope with these changes is not clear. When the Newfoundland fishery was revitalized after the declaration of the 200 nanometer (nm) limit, its economic structure overexpanded with largely immobile capital and labor and resulted in disaster. When the cod fishery started to decline in the late 1980's, several years passed before many of the necessary adjustments occurred. Whether the situation will be any different in response to climate change depends on whether lessons have been learned, and whether the social and political systems are prepared to adjust. Both the expansion of the groundfishery in the late 1970's and the failure of the fishery to adjust to decimated stocks in the early and mid-1990's were largely due to the subsidies. During the 1970's the expansion was mainly financed by the federal and provincial governments. Despite the efforts of the federal government to adjust fishers out of the industry in the 1990's, the adjustment programs became income maintenance programs, which in effect encouraged fishers to remain in the industry in the hope that the fish would return. It was only when the subsidies were substantially reduced after 1998 that a significant number of fishers left the industry (based on the assumption that departures from the fishery are reflected in the census figures).

Subsidies are not the results of whimsical acts of governments or politics but are responses to real social and economic concerns. As long as the government considers the survival of small rural communities a major priority, subsidies to the fishery (the primary industry in these communities) will continue. While subsidies exist, the response of people to changes in the industry will be slow.Without subsidies, economic forces will require change, probably rapid change if the fishery is declining. If the biological base of the fishery is expanding, there is always the possibility that the industry will overexpand without government help. Government financial assistance would virtually ensure that overexpansion would occur.

To the extent that adjustments induced by climate change cause human suffering, the government can be expected to ameliorate the situation and ease the necessary transitions. But there is strong precedent for transition programs being transformed into short-run income maintenance programs. If that were to happen again, the process of adjustment is likely to be as long, painful, and wasteful as before.

Thus, 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.

Concluding Comments (13.4.9)

The ecosystem off the northeast coast of Canada is under the influence of the Labrador Current, which carries cold water south from Davis Strait, the Canadian Archipelago, and Hudson Bay. As a result, climate impacts in this ecosystem can be compared to impacts on comparable ecosystems in the Northeast Atlantic and Iceland. Historically, the dominant demersal species were cod, Greenland halibut, and American plaice, the dominant invertebrates were northern shrimp and snow crabs, the dominant pelagic fish was capelin, and the dominant top predators were harp seals and whales.

The Labrador/Newfoundland ecosystem experienced major changes in the 1980's and 1990's. Atlantic cod and most other demersal fish, including species that were not targeted by commercial fishing, had declined to very low levels by the early 1990's. In contrast, snow crab and especially northern shrimp surged during the 1980's and 1990's and now support the most important fisheries in the area. Harp seals increased in abundance between the early 1970's and the mid-1990's. Capelin have been found in much reduced quantities in offshore acoustic surveys since the early 1990's, but indices of capelin abundance in the inshore surveys have not experienced similar declines, leaving the status of capelin uncertain and controversial.

The relative importance of overfishing and the environment on changes in cod and Greenland halibut has not been determined, although fishing is generally accepted as the most important factor affecting cod abundance. Ocean climate is thougfht to have had an impact on the lack of cod recovery, although this has not been quantified. Exploitation has not been shown to have affected any aspect of capelin biology in this area. Although there have been several changes in capelin biology since the early 1990's, there is no clear indication of what external factor(s) has (have) influenced the changes. A combination of reduced predation and favorable environmental conditions probably contributed to the success of northern shrimp and snow crab. Harp seals increased because of reduced commercial harvesting.

Changes of the magnitude that have occurred in the biological components of the ecosystem since the early 1980's are unprecedented and together with the lack of regional predictions of changes in the ocean due to climate change, make predictions of biological responses to climate change highly speculative.

If there is no change from the present state or even a cooling, it is likely that the current balance of species will persist.

With a moderate, gradual warming, there is likely to be a change back to a cod–capelin system with a gradual decline in northern shrimp and snow crab. Cod and other demersal, shelf-dwelling species and capelin are likely to move north. Many existing capelin spawning beaches are likely to disappear as sea levels rise. If there is an increase in demersal spawning by capelin in the absence of new spawning beaches, capelin survival is likely to decline. Seals are likely to experience higher pup mortality as sea ice thins. Increases in regional storm intensities are also likely to result in higher pup mortality. A reduction in the extent and duration of sea ice is likely to permit fishing further to the north. A reduction in sea-ice cover is likely to shorten Greenland halibut fisheries that are conducted through fast ice.

If a more intense regional warming occurs as a consequence of extensive climatic warming, then fpredicting the responses of the biological community to these changes must occur in the absence of historic precedence and be completely speculative. Such an event is likely to improve conditions for cold-temperate species such as cod, improve conditions for more southern species such as haddock and herring, and even result in the formation of demersally spawning stocks of capelin.

Although the fishery in Newfoundland has accounted for 5% or less of provincial GDP since 1971, it dominates the economy and culture in rural areas. The cod fishery expanded rapidly in the 1980's and then contracted rapidly in the 1990's, the latter in response to the fishing moratorium. The social and economic effects of changes in fish stocks due to climate change are likely to be less than the historical changes experienced in the latter part of the 20th century in Newfoundland and Labrador.

Past experience suggests that the political system will respond but that the details of the fresponse are impossible to predict. It is, however, possible to prescribe directions that governments should follow in the event of expansions or contractions of ffish stocks resulting from climate change. If fish stocks off Newfoundland increase over the next 20 or 50 years, care should be taken (1) to restrict by government regulation the magnitude of any expansion of capital and labor in the fishery; (2) to ensure that such economic incentives are in place that excessive growth does not occur; or (3) to combine the two. If fish stocks decrease over the next 20 or 50 years, then it should be clear from the start that endless subsidies will not be forthcoming. License buyouts, even generous license buyouts of core enterprises for instance, would help. While these payments are subsidies, they are limited in scope and time and would have the effect of permanently shrinking the factor base of the industry.

Aquaculture in Newfoundland and Labrador is relatively small but there is interest in expansion, especially with the lack of recovery of cod stocks. A gradual warming of shelf waters is likely to lead to increased opportunity for aquaculture. Warmer temperatures and shorter periods of sea-ice cover are likely to enable mussel farming to be more productive. Warmer waters are also likely to promote the development of Atlantic cod farming and the farming of Atlantic salmon along the east coast of Newfoundland.

Chapter 13: Fisheries and Aquaculture
13.1 Introduction (Fisheries and aquaculture in the Newfoundland and Labrador Seas, Northeastern Canada)
13.2 Northeast Atlantic – Barents and Norwegian Seas
13.3 Central North Atlantic – Iceland and Greenland
13.4 Newfoundland and Labrador Seas, Northeastern Canada
13.5 North Pacific – Bering Sea
13.6. Synthesis and key findings
13.7. Research recommendations

References

  1. ^ DFO, 2003a. Northern (2J+3KL) Cod. Department of Fisheries and Oceans, Canada, Stock Status Report 2003/018.-- Lilly, G.R., P.A. Shelton, J. Brattey, N.G. Cadigan, B.P. Healey, E.F. Murphy and D.E. Stansbury, 2001. An Assessment of the Cod Stock in NAFO Divisions 2J+3KL. Department of Fisheries and Oceans, Canadian Stock Assessment Secretariat Res. Doc. 2001/044, 148p.
  2. {{note|91}Carscadden, J.E., K.T. Frank and W.C. Leggett, 2000. Evaluation of an environment-recruitment model for capelin (Mallotus villosus). ICES Journal of Marine Science, 57:412–418.-- Leggett,W.C., K.T. Frank and J.E. Carscadden, 1984. Meteorological and hydrographic regulation of year-class strength in capelin (Mallotus villosus). Canadian Journal of Fisheries and Aquatic Sciences, 41:1193–1201.


Citation

Committee, I. (2012). Fisheries and aquaculture in the Newfoundland and Labrador Seas, Northeastern Canada. Retrieved from http://editors.eol.org/eoearth/wiki/Fisheries_and_aquaculture_in_the_Newfoundland_and_Labrador_Seas,_Northeastern_Canada
  1. Anon, 2001. Sea Ice Climatic Atlas, East Coast of Canada, 1971–2000. Canadian Ice Service, Ministery of Public Works.
  2. Lilly, G.R. and M. Simpson, 2000. Distribution and Biomass of Capelin, Arctic Cod and Sand Lance on the Northeast Newfoundland Shelf and Grand Bank as Deduced from Bottomtrawl Surveys. Department of Fisheries and Oceans, Canadian Stock Assessment Secretariat Res. Doc. 2000/091, 40p.-- Lilly, G.R., H. Hop, D.E. Stansbury and C.A. Bishop, 1994. Distribution and abundance of polar cod (Boreogadus saida) off southern Labrador and eastern Newfoundland. ICES C.M.1994/O:6, 21p.
  3. for the historical background of Newfoundland, see: Chadwick, St. J., 1967 Newfoundland: Island into Province. Cambridge University Press.-- Innis, H.A., 1954.The Cod Fisheries. University of Toronto Press, revised edition.-- Lounsbury, R.G., 1934.The British Fishery in Newfoundland, 1634–1763.Yale University Press.
  4. Burnett, E.C., 1941.The Continental Congress. Macmillan.
  5. Letto, D.M., 1998. Chocolate Bars and Rubber Boots: the Smallwood Industrialization Plan. Blue Hill Publishers.
  6. ;also see: Anon, 2004a. Report of the joint ICES/NAFO Working Group on Harp and Hooded Seals. ICES CM 2004/ACFM:6.-- Darby, C., B. Healey, J.-C. Mahé and W.R. Bowering, 2004. Greenland halibut (Reinhardtius hippoglossoides) in Subarea 2 and Divisions 3KLMNO: an assessment of stock status based on Extended Survivors Analysis,ADAPT, and ASPIC analyses, with stochastic projections of potential stock dynamics. Northwest Atlantic Fisheries Organization, Scientific Council Research Doc. 04/55, Serial No. N5008.-- Dawe, E., D.Taylor, D. Orr, D. Parsons, E. Colbourne, D. Stansbury, J. Drew, P. Beck, P. O’Keefe, P.Veitch, E. Seward, D. Ings and A. Pardy, 2004. An assessment of Newfoundland and Labrador snow crab in 2003. Department of Fisheries and Oceans, Canadian Science Advisory Secretariat Res. Doc. 2004/024.-- Hvingel, C., 2004.The fishery for northern shrimp (Pandalus borealis) off West Greenland, 1970–2004. Northwest Atlantic Fisheries Organization, Scientific Council Research Doc. 04/75, Serial No. N5045.-- Lilly, G.R., P.A. Shelton, J. Brattey, N.G. Cadigan, B.P. Healey, E.F. Murphy, D.E. Stansbury and N. Chen, 2003. An assessment of the cod stock in NAFO Divisions 2J+3KL in February 2003. Department of Fisheries and Oceans, Canadian Science Advisory Secretariat Res. Doc. 2003/023, 157p.-- Murphy, E.F. and C.A. Bishop, 1995.The status of 2GH cod, 3LNO haddock, 3Ps haddock and 3Ps pollock. Canadian Department of Fisheries and Oceans, Atlantic Fisheries Research Doc. 95/33.-- Orr, D.C., P.J.Veitch and D.J. Sullivan, 2003. Northern shrimp (Pandalus borealis) off Baffin Island, Labrador and Northeastern Newfoundland. Department of Fisheries and Oceans, Canadian Science Advisory Secretariat Res. Doc. 2003/050.-- Orr, D.C., P.J.Veitch and D.J. Sullivan, 2004. Divisions 3LNO northern shrimp (Pandalus borealis) – interim monitoring update. Northwest Atlantic Fisheries Organization, Scientific Council Research Doc. 04/65, Serial No. N5026.-- Stenson, G.B., B.P. Healey, B. Sjare and D.Wakeham, 2000. Catch-atage of northwest Atlantic harp seals, 1952–1999. Department of Fisheries and Oceans, Canadian Stock Assessment Secretariat Res. Doc. 2000/079.
  7. Bundy, A., G.R. Lilly and P.A. Shelton, 2000. A Mass Balance Model of the Newfoundland-Labrador Shelf. Canadian Technical Report of Fisheries and Aquatic Sciences 2310, xiv + 157p.-- Carscadden, J.E., K.T. Frank and W.C. Leggett, 2001. Ecosystem changes and the effects on capelin (Mallotus villosus), a major forage species. Canadian Journal of Fisheries and Aquatic Sciences, 58:73–85.-- Livingston, P.A. and S.Tjelmeland, 2000. Fisheries in boreal ecosystems. ICES Journal of Marine Science, 57:619–627.
  8. Atkinson, D.B., 1994. Some observations on the biomass and abundance of fish captured during stratified-random bottom trawl surveys in NAFO Divisions 2J and 3KL, autumn 1981–1991. NAFO Scientific Council Studies, 21:43–66.-- Gomes, M.C., R.L. Haedrich and M.G.Villagarcia, 1995. Spatial and temporal changes in the groundfish assemblages on the north-east Newfoundland/Labrador Shelf, north-west Atlantic, 1978–1991. Fisheries Oceanography, 4:85–101.
  9. DFO, 2002a. Newfoundland and Labrador snow crab. Department of Fisheries and Oceans, Canada, Science Stock Status Report C2-01.
  10. DFO, 2002b. Northern shrimp (Pandalus borealis) - Div. 0B to 3K. Department of Fisheries and Oceans, Canada, Science Stock Status Report C2-05.
  11. DFO, 2000c. Northwest Atlantic harp seals. Department of Fisheries and Oceans, Canada, Science Stock Status Report E1-01.
  12. DFO, 2000b. Capelin in Subarea 2 + Div. 3KL. Department of Fisheries and Oceans, Canada, Science Stock Status Report B2-02.-- DFO, 2001. Capelin in Subarea 2 + Div. 3KL - Update. Department of Fisheries and Oceans, Canada, Science Stock Status Report B2-02.
  13. DFO, 2003b. Newfoundland and Labrador Atlantic salmon 2003 stock status update. Department of Fisheries and Oceans, Canada, Stock Status Update Report 2003/048.-- Narayanan, S., J. Carscadden, J.B. Dempson, M.F. O’Connell, S. Prinsenberg, D.G. Reddin and N. Shackell, 1995. Marine climate off Newfoundland and its influence on Atlantic salmon (Salmo salar) and capelin (Mallotus villosus). In: R.J. Beamish (ed.). Climate Change and Northern Fish Populations, pp. 461–474. Canadian Special Publication of Fisheries and Aquatic Sciences 121.
  14. Colbourne, E.B. and J.T. Anderson, 2003. Biological response in a changing ocean environment in Newfoundland waters during the latter decades of the 1900s. ICES Marine Science Symposia, 219:169–181.-- Colbourne, E.B., S. Narayanan and S. Prinsenberg, 1994. Climate changes and environmental conditions in the Northwest Atlantic, 1970–1993. ICES Marine Science Symposia, 198:311–322.-- Mann, K.H. and K.F. Drinkwater, 1994. Environmental influences on fish and shellfish production in the Northwest Atlantic. Environmental Reviews, 2:16–32.-- Narayanan, S., J. Carscadden, J.B. Dempson, M.F. O’Connell, S. Prinsenberg, D.G. Reddin and N. Shackell, 1995. Marine climate off Newfoundland and its influence on Atlantic salmon (Salmo salar) and capelin (Mallotus villosus). In: R.J. Beamish (ed.). Climate Change and Northern Fish Populations, pp. 461–474. Canadian Special Publication of Fisheries and Aquatic Sciences 121.
  15. e.g.: NAFO, 2001a. STATLANT 21 reported catches by stock tabulated against STACFIS estimates, 1985–1999. Northwest Atlantic Fisheries Organization, Scientific Council Summary Document 01/5, 23p-- NAFO, 2001b. Scientific Council Report, Northwest Atlantic Fisheries Organization. June 2001, 339p.
  16. data from: Murphy, E.F. and C.A. Bishop, 1995.The status of 2GH cod, 3LNO haddock, 3Ps haddock and 3Ps pollock. Canadian Department of Fisheries and Oceans, Atlantic Fisheries Research Doc. 95/33.-- Lilly, G.R., P.A. Shelton, J. Brattey, N.G. Cadigan, B.P. Healey, E.F. Murphy, D.E. Stansbury and N. Chen, 2003. An assessment of the cod stock in NAFO Divisions 2J+3KL in February 2003. Department of Fisheries and Oceans, Canadian Science Advisory Secretariat Res. Doc. 2003/023, 157p.
  17. Baird, J.W., C.R. Stevens and E.F. Murphy, 1992. A review of hydroacoustic surveys conducted during winter for 2J3KL cod, 1987–1992. Canadian Atlantic Fisheries Scientific Advisory Committee, Res. Doc. 92/107, 14p.
  18. Cadigan, S.T., 1999. Failed proposals for fisheries management and conservation in Newfoundland, 1855–1880.In: D. Newell and R.E. Ommer (eds.). Fishing Places, Fishing People:Traditions and Issues in Canadian Small-scale Fisheries, pp. 147–169. University of Toronto Press.
  19. Templeman,W., 1966. Marine resources of Newfoundland. Fisheries Research Board of Canada Bulletin 154.
  20. Lear,W.H. and L.S. Parsons, 1993. History and management of the fishery for northern cod in NAFO Divisions 2J, 3K and 3L. In: L.S. Parsons and W.H. Lear (eds.). Perspectives on Canadian Marine Fisheries Management. pp. 55–89. Canadian Bulletin of Fisheries and Aquatic Sciences 226.
  21. Hutchings, J.A. and R.A. Myers, 1995.The biological collapse of Atlantic cod off Newfoundland and Labrador: an exploration of historical changes in exploitation, harvesting technology, and management. In: R. Arnason and L. Felt (eds.).The North Atlantic Fisheries: Successes, Failures, and Challenges, pp. 39–93.The Institute of Island Studies, Charlottetown, Prince Edward Island.
  22. Lear,W.H., 1998. History of fisheries in the Northwest Atlantic: the 500-year perspective. Journal of Northwest Atlantic Fishery Science, 23:41–73.
  23. Neis, B., D.C. Schneider, L. Felt, R.L. Haedrich, J. Fischer and J.A. Hutchings, 1999. Fisheries assessment: what can be learned from interviewing resource users? Canadian Journal of Fisheries and Aquatic Sciences, 56:1949–1963.
  24. Hutchings, J.A. and M. Ferguson, 2000.Temporal changes in harvesting dynamics of Canadian inshore fisheries for northern Atlantic cod, Gadus morhua. Canadian Journal of Fisheries and Aquatic Sciences, 57:805–814.
  25. Darby, C., B. Healey, J.-C. Mahé and W.R. Bowering, 2004. Greenland halibut (Reinhardtius hippoglossoides) in Subarea 2 and Divisions 3KLMNO: an assessment of stock status based on Extended Survivors Analysis,ADAPT, and ASPIC analyses, with stochastic projections of potential stock dynamics. Northwest Atlantic Fisheries Organization, Scientific Council Research Doc. 04/55, Serial No. N5008.
  26. Bowering,W.R. and W.B. Brodie, 1995. Greenland halibut (Reinhardtius hippoglossoides). A review of the dynamics of its distribution and fisheries off Eastern Canada and Greenland. In: A.G. Hopper (ed.). Deep-water Fisheries of the North Atlantic Oceanic Slope, pp. 113–160. Kluwer Academic.
  27. Bowering,W.R. and W.B. Brodie, 1995. Greenland halibut (Reinhardtius hippoglossoides). A review of the dynamics of its distribution and fisheries off Eastern Canada and Greenland. In: A.G. Hopper (ed.). Deep-water Fisheries of the North Atlantic Oceanic Slope, pp. 113–160. Kluwer Academic.
  28. Bowering,W.R. and K.H. Nedreaas, 2000. A comparison of Greenland halibut (Reinhardtius hippoglossoides (Walbaum)) fisheries and distribution in the Northwest and Northeast Atlantic. Sarsia, 85:61–76.
  29. NAFO, 2001b. Scientific Council Report, Northwest Atlantic Fisheries Organization. June 2001, 339p.
  30. Bowering,W.R. and W.B. Brodie, 1995. Greenland halibut (Reinhardtius hippoglossoides). A review of the dynamics of its distribution and fisheries off Eastern Canada and Greenland. In: A.G. Hopper (ed.). Deep-water Fisheries of the North Atlantic Oceanic Slope, pp. 113–160. Kluwer Academic.-- Bowering,W.R. and K.H. Nedreaas, 2000. A comparison of Greenland halibut (Reinhardtius hippoglossoides (Walbaum)) fisheries and distribution in the Northwest and Northeast Atlantic. Sarsia, 85:61–76.
  31. NAFO, 2001b. Scientific Council Report, Northwest Atlantic Fisheries Organization. June 2001, 339p.
  32. Bowering,W.R. and W.B. Brodie, 1995. Greenland halibut (Reinhardtius hippoglossoides). A review of the dynamics of its distribution and fisheries off Eastern Canada and Greenland. In: A.G. Hopper (ed.). Deep-water Fisheries of the North Atlantic Oceanic Slope, pp. 113–160. Kluwer Academic.
  33. Treble, M.A. and R. Bowering, 2002.The Greenland Halibut (Reinhardtius hippoglossoides) Fishery in NAFO Division 0A. NAFO SCR Doc. 02/46. Serial No. N4658. 9p.
  34. M.A.Treble, Fisheries and Oceans Canada, pers. comm., 2003
  35. Templeman,W., 1968. Review of some aspects of capelin biology in the Canadian area of the northwest Atlantic. ICES Rapports et Procès- Verbaux des Réunions, 158:41–53.
  36. DFO, 2000a. East and Southeast Newfoundland Atlantic Herring. Department of Fisheries and Oceans, Canada, Science Stock Status Report B2-01.
  37. Lear,W.H., 1979. Distribution, size, and sexual maturity of Arctic cod (Boreogadus saida) in the northwest Atlantic during 1959–1978. Canadian Atlantic Fisheries Scientific Advisory Committee, Res. Doc. 79/17.
  38. Maxim, C., 1980. Capelin (Mallotus villosus) Catch, Effort, and Biological Characteristics in the Romanian Fishery in Division 2J, September–October 1979. Northwest Atlantic Fisheries Organization Scientific Council Research Doc. 80/4, Serial No. N030.
  39. also see: Hvingel, C., 2004.The fishery for northern shrimp (Pandalus borealis) off West Greenland, 1970–2004. Northwest Atlantic Fisheries Organization, Scientific Council Research Doc. 04/75, Serial No. N5045.-- Orr, D.C., P.J.Veitch and D.J. Sullivan, 2003. Northern shrimp (Pandalus borealis) off Baffin Island, Labrador and Northeastern Newfoundland. Department of Fisheries and Oceans, Canadian Science Advisory Secretariat Res. Doc. 2003/050.-- Orr, D.C., P.J.Veitch and D.J. Sullivan, 2004. Divisions 3LNO northern shrimp (Pandalus borealis) – interim monitoring update. Northwest Atlantic Fisheries Organization, Scientific Council Research Doc. 04/65, Serial No. N5026.
  40. Orr, D., D.G. Parsons, P.J.Veitch and D.J. Sullivan, 2001a. Northern Shrimp (Pandalus borealis) off Baffin Island, Labrador and Northeastern Newfoundland – First Interim Review. Department of Fisheries and Oceans, Canadian Stock Assessment Secretariat Res. Doc. 2001/043.-- Anon, 2004a. Report of the joint ICES/NAFO Working Group on Harp and Hooded Seals. ICES CM 2004/ACFM:6.-- Stenson, G.B., B.P. Healey, B. Sjare and D.Wakeham, 2000. Catch-atage of northwest Atlantic harp seals, 1952–1999. Department of Fisheries and Oceans, Canadian Stock Assessment Secretariat Res. Doc. 2000/079.
  41. Orr, D.C., D.G. Parsons, P.Veitch and D. Sullivan, 2001b. An Update of Information Pertaining to Northern Shrimp (Pandalus borealis) and Groundfish in NAFO Divisions 3LNO. Northwest Atlantic Fisheries Organization, Scientific Council Research Doc. 01/186, Serial No. N4576.
  42. also see: Hvingel, C., 2004.The fishery for northern shrimp (Pandalus borealis) off West Greenland, 1970–2004. Northwest Atlantic Fisheries Organization, Scientific Council Research Doc. 04/75, Serial No. N5045.-- Orr, D.C., P.J.Veitch and D.J. Sullivan, 2003. Northern shrimp (Pandalus borealis) off Baffin Island, Labrador and Northeastern Newfoundland. Department of Fisheries and Oceans, Canadian Science Advisory Secretariat Res. Doc. 2003/050.-- Orr, D.C., P.J.Veitch and D.J. Sullivan, 2004. Divisions 3LNO northern shrimp (Pandalus borealis) – interim monitoring update. Northwest Atlantic Fisheries Organization, Scientific Council Research Doc. 04/65, Serial No. N5026.
  43. Taylor, D.M. and P.G. O’Keefe, 1999. Assessment of the 1998 Newfoundland and Labrador Snow Crab Fishery. Department of Fisheries and Oceans, Canadian Stock Assessment Secretariat Res. Doc. 99/143.
  44. DFO, 2002a. Newfoundland and Labrador snow crab. Department of Fisheries and Oceans, Canada, Science Stock Status Report C2-01.
  45. data from: Anon, 2004a. Report of the joint ICES/NAFO Working Group on Harp and Hooded Seals. ICES CM 2004/ACFM:6.-- Stenson, G.B., B.P. Healey, B. Sjare and D.Wakeham, 2000. Catch-atage of northwest Atlantic harp seals, 1952–1999. Department of Fisheries and Oceans, Canadian Stock Assessment Secretariat Res. Doc. 2000/079.
  46. Bundy, A., G.R. Lilly and P.A. Shelton, 2000. A Mass Balance Model of the Newfoundland-Labrador Shelf. Canadian Technical Report of Fisheries and Aquatic Sciences 2310, xiv + 157p.
  47. Healey, B.P. and G.B. Stenson, 2000. Estimating pup production and population size of the Northwest Atlantic harp seal (Phoca groenlandica). Department of Fisheries and Oceans, Canadian Stock Assessment Secretariat Res. Doc. 2000/081.-- Stenson, G.B., M.O. Hammill, M.C.S. Kingsley, B. Sjare,W.G.Warren and R.A. Myers, 2002. Is there evidence of increased pup production in northwest Atlantic harp seals, Pagophilus groenlandicus? ICES Journal of Marine Science, 59:81–92.
  48. Stenson, G.B., M.O. Hammill, M.C.S. Kingsley, B. Sjare,W.G.Warren and R.A. Myers, 2002. Is there evidence of increased pup production in northwest Atlantic harp seals, Pagophilus groenlandicus? ICES Journal of Marine Science, 59:81–92.
  49. Bundy, A., G.R. Lilly and P.A. Shelton, 2000. A Mass Balance Model of the Newfoundland-Labrador Shelf. Canadian Technical Report of Fisheries and Aquatic Sciences 2310, xiv + 157p.
  50. Ibid..
  51. data from: Murphy, E.F. and C.A. Bishop, 1995.The status of 2GH cod, 3LNO haddock, 3Ps haddock and 3Ps pollock. Canadian Department of Fisheries and Oceans, Atlantic Fisheries Research Doc. 95/33.-- Lilly, G.R., P.A. Shelton, J. Brattey, N.G. Cadigan, B.P. Healey, E.F. Murphy, D.E. Stansbury and N. Chen, 2003. An assessment of the cod stock in NAFO Divisions 2J+3KL in February 2003. Department of Fisheries and Oceans, Canadian Science Advisory Secretariat Res. Doc. 2003/023, 157p.
  52. Atkinson, D.B. and B. Bennett, 1994. Proceedings of a Northern Cod Workshop Held in St. John’s, Newfoundland, Canada, January 27–29, 1993. Canadian Technical Report of Fisheries and Aquatic Sciences No. 1999, 64p.-- Shelton, P.A. and G.R. Lilly, 2000. Interpreting the collapse of the northern cod stock from survey and catch data. Canadian Journal of Fisheries and Aquatic Sciences, 57:2230–2239.
  53. Haedrich, R.L., J. Fischer and N.V. Chernova, 1997. Ocean temperatures and demersal fish abundance on the northeast Newfoundland continental shelf. In: L.E. Hawkins and S. Hutchinson (eds.).The responses of marine organisms to their environments - Proceedings of the 30th European Marine Biology Symposium, pp. 211–221. University of Southampton.-- Hutchings, J.A., 1996. Spatial and temporal variation in the density of northern cod and a review of hypotheses for the stock’s collapse. Canadian Journal of Fisheries and Aquatic Sciences, 53:943–962.-- Hutchings, J.A. and R.A. Myers, 1994.What can be learned from the collapse of a renewable resource? Atlantic cod, Gadus morhua, of Newfoundland and Labrador. Canadian Journal of Fisheries and Aquatic Sciences, 51:2126–2146.-- Myers, R.A. and N.G. Cadigan, 1995.Was an increase in natural mortality responsible for the collapse of northern cod? Canadian Journal of Fisheries and Aquatic Sciences, 52:1274 1285.-- Myers, R.A., N.J. Barrowman, J.M. Hoenig and Z. Qu, 1996a. The collapse of cod in Eastern Canada: the evidence from tagging data. ICES Journal of Marine Science, 53:629–640.-- Myers, R.A., J.A. Hutchings and N.J. Barrowman, 1996b. Hypotheses for the decline of cod in the North Atlantic. Marine Ecology Progress Series, 138:293–308.-- Myers, R.A., N.J. Barrowman and J.A. Hutchings, 1997a. Inshore exploitation of Newfoundland Atlantic cod (Gadus morhua) since 1948 as estimated from mark-recapture data. Canadian Journal of Fisheries and Aquatic Sciences, 54(1):224–235.-- Myers, R.A., J.A. Hutchings and N.J. Barrowman, 1997b.Why do fish stocks collapse? The example of cod in Atlantic Canada. Ecological Applications, 7:91–106.
  54. Drinkwater, K.F., 2000. Changes in ocean climate and its general effect on fisheries: examples from the north-west Atlantic. In: D. Mills (ed.).The Ocean Life of Atlantic Salmon: Environmental and Biological Factors Influencing Survival, pp. 116–136. Fishing News Books, Oxford.-- 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.-- Mann, K.H. and K.F. Drinkwater, 1994. Environmental influences on fish and shellfish production in the Northwest Atlantic. Environmental Reviews, 2:16–32.-- Parsons, L.S. and W.H. Lear, 2001. Climate variability and marine ecosystem impacts: a North Atlantic perspective. Progress in Oceanography, 49:167–188.
  55. Rose, G.A., B. deYoung, D.W. Kulka, S.V. Goddard and G.L. Fletcher, 2000. Distribution shifts and overfishing the northern cod (Gadus morhua): a view from the ocean. Canadian Journal of Fisheries and Aquatic Sciences, 57:644–663.
  56. Planque, B. and T. Frédou, 1999.Temperature and the recruitment of Atlantic cod (Gadus morhua). Canadian Journal of Fisheries and Aquatic Sciences, 56:2069–2077.
  57. deYoung, B. and G.A. Rose, 1993. On recruitment and distribution of Atlantic cod (Gadus morhua) off Newfoundland. Canadian Journal of Fisheries and Aquatic Sciences, 50:2729–2741.-- Hutchings, J.A. and R.A. Myers, 1994.What can be learned from the collapse of a renewable resource? Atlantic cod, Gadus morhua, of Newfoundland and Labrador. Canadian Journal of Fisheries and Aquatic Sciences, 51:2126–2146.-- Planque, B. and T. Frédou, 1999.Temperature and the recruitment of Atlantic cod (Gadus morhua). Canadian Journal of Fisheries and Aquatic Sciences, 56:2069–2077.-- Taggart, C.T., J. Anderson, C. Bishop, E. Colbourne, J. Hutchings, G. Lilly, J. Morgan, E. Murphy, R. Myers, G. Rose and P. Shelton, 1994. Overview of cod stocks, biology, and environment in the Northwest Atlantic region of Newfoundland, with emphasis on northern cod. ICES Marine Science Symposia, 198:140–157.
  58. Hutchings, J.A. and R.A. Myers, 1994.What can be learned from the collapse of a renewable resource? Atlantic cod, Gadus morhua, of Newfoundland and Labrador. Canadian Journal of Fisheries and Aquatic Sciences, 51:2126–2146-- Morgan, M.J., P.A. Shelton, D.P. Stansbury, J. Brattey and G.R. Lilly, 2000. An Examination of the Possible Effect of Spawning Stock Characteristics on Recruitment in four Newfoundland Groundfish Stocks. Canadian Stock Assessment Secretariat Res. Doc. 2000/028.-- Myers, R.A., K.F. Drinkwater, N.J. Barrowman and J.W. Baird, 1993. Salinity and recruitment of Atlantic cod (Gadus morhua) in the Newfoundland region. Canadian Journal of Fisheries and Aquatic Sciences, 50:1599–1609.-- Rice, J.C. and G.T. Evans, 1988.Tools for embracing uncertainty in the management of the cod fishery of NAFO divisions 2J+3KL. Journal du Conseil International pour l’Exploration de la Mer, 45:73–81.but see: 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.
  59. Sutcliffe,W.H., R.H. Loucks, K.F. Drinkwater and A.R. Coote, 1983. Nutrient flux onto the Labrador Shelf from Hudson Strait and its biological consequences. Canadian Journal of Fisheries and Aquatic Sciences, 40:1692–1701.
  60. Myers, R.A., K.F. Drinkwater, N.J. Barrowman and J.W. Baird, 1993. Salinity and recruitment of Atlantic cod (Gadus morhua) in the Newfoundland region. Canadian Journal of Fisheries and Aquatic Sciences, 50:1599–1609.
  61. Hutchings, J.A. and R.A. Myers, 1994.What can be learned from the collapse of a renewable resource? Atlantic cod, Gadus morhua, of Newfoundland and Labrador. Canadian Journal of Fisheries and Aquatic Sciences, 51:2126–2146-- Shelton, P.A. and D.B. Atkinson, 1994. Failure of the Div. 2J3KL Cod Recruitment Prediction using Salinity. Department of Fisheries and Oceans, Canadian Stock Assessment Secretariat Res. Doc. 94/66.14p.
  62. Krohn, M., S. Reidy and S. Kerr, 1997. Bioenergetic analysis of the effects of temperature and prey availability on growth and condition of northern cod (Gadus morhua). Canadian Journal of Fisheries and Aquatic Sciences, 54(1):113–121.-- Shelton, P.A., G.R. Lilly and E. Colbourne, 1999. Patterns in the annual weight increment for Div. 2J+3KL cod and possible prediction for stock projection. Journal of Northwest Atlantic Fishery Science, 25:151–159.
  63. Davis, M.B., 1992. Description of the inshore fishery during 1991 for 2J3KL cod as reported by inshore fisherpersons. Canadian Atlantic Fisheries Scientific Advisory Committee, Research Document 93/37.
  64. Lilly, G.R., 1994. Predation by Atlantic cod on capelin on the southern Labrador and Northeast Newfoundland shelves during a period of changing spatial distributions. ICES Marine Science Symposia, 198:600–611.-- Taggart, C.T., J. Anderson, C. Bishop, E. Colbourne, J. Hutchings, G. Lilly, J. Morgan, E. Murphy, R. Myers, G. Rose and P. Shelton, 1994. Overview of cod stocks, biology, and environment in the Northwest Atlantic region of Newfoundland, with emphasis on northern cod. ICES Marine Science Symposia, 198:140–157.
  65. Baird, J.W., C.R. Stevens and E.F. Murphy, 1992. A review of hydroacoustic surveys conducted during winter for 2J3KL cod, 1987–1992. Canadian Atlantic Fisheries Scientific Advisory Committee, Res. Doc. 92/107, 14p.
  66. Kulka, D.W., J.S.Wroblewski and S. Narayanan, 1995. Recent changes in the winter distribution and movements of northern Atlantic cod (Gadus morhua Linnaeus, 1758) on the Newfoundland-Labrador Shelf. ICES Journal of Marine Science, 52:889–902.-- Rose, G.A. and D.W. Kulka, 1999. Hyperaggregation of fish and fisheries: how catch-per-unit-effort increased as the northern cod (Gadus morhua) declined. Canadian Journal of Fisheries and Aquatic Sciences, 56(1):118–127.-- Rose, G.A., B.A. Atkinson, J. Baird, C.A. Bishop and D.W. Kulka, 1994. Changes in distribution of Atlantic cod and thermal variations in Newfoundland waters, 1980–1992. ICES Marine Science Symposia, 198:542–552.
  67. Bishop, C.A. and J.W. Baird, 1994. Spatial and temporal variability in condition factors of Divisions 2J and 3KL cod (Gadus morhua). Northwest Atlantic Fisheries Organization, Scientific Council Studies, 21:105–113.-- Lilly, G.R., 2001. Changes in size at age and condition of cod (Gadus morhua) off Labrador and eastern Newfoundland during 1978–2000. ICES CM 2001/V:15, 34p.
  68. Atkinson, D.B., 1994. Some observations on the biomass and abundance of fish captured during stratified-random bottom trawl surveys in NAFO Divisions 2J and 3KL, autumn 1981–1991. NAFO Scientific Council Studies, 21:43–66.-- Gomes, M.C., R.L. Haedrich and M.G.Villagarcia, 1995. Spatial and temporal changes in the groundfish assemblages on the north-east Newfoundland/Labrador Shelf, north-west Atlantic, 1978–1991. Fisheries Oceanography, 4:85–101.
  69. Haedrich, R.L. and S.M. Barnes, 1997. Changes over time of the size structure in an exploited shelf fish community. Fisheries Research, 31:229–239.-- Haedrich, R.L. and J. Fischer, 1996. Stability and change of exploited fish communities in a cold ocean continental shelf ecosystem. Senckenbergiana Maritima, 27:237–243.-- Haedrich, R.L., J. Fischer and N.V. Chernova, 1997. Ocean temperatures and demersal fish abundance on the northeast Newfoundland continental shelf. In: L.E. Hawkins and S. Hutchinson (eds.).The responses of marine organisms to their environments - Proceedings of the 30th European Marine Biology Symposium, pp. 211–221. University of Southampton.-- Hutchings, J.A., 1996. Spatial and temporal variation in the density of northern cod and a review of hypotheses for the stock’s collapse. Canadian Journal of Fisheries and Aquatic Sciences, 53:943–962.
  70. Morgan, M.J.,W.B. Brodie and D.W. Kulka, 2002.Was overexploitation the cause of the decline of the American plaice stock off Labrador and northeast Newfoundland? Fisheries Research, 57:39–49.
  71. Lilly, G.R., 2002. Ecopath modelling of the Newfoundland Shelf: observations on data availability within the Canadian Department of Fisheries and Oceans. In:T. Pitcher, M.Vasconcellos, S. Heymans, C. Brignall and N. Haggan (eds). Information Supporting Past and Present Ecosystem Models of Northern British Columbia and the Newfoundland Shelf. Fisheries Centre, University of British Columbia. Fisheries Centre Research Reports, 10(1):22–27.-- Rice, J., 2002. Changes to the large marine ecosystem of the Newfoundland-Labrador Shelf. In: K. Sherman and H.R. Skjoldal (eds.). Large Marine Ecosystems of the North Atlantic, pp. 51–103. Elsevier Science.
  72. Bowering,W.R. and W.B. Brodie, 1995. Greenland halibut (Reinhardtius hippoglossoides). A review of the dynamics of its distribution and fisheries off Eastern Canada and Greenland. In: A.G. Hopper (ed.). Deep-water Fisheries of the North Atlantic Oceanic Slope, pp. 113–160. Kluwer Academic.
  73. Ibid.
  74. Ibid.
  75. Bowering,W.R. and K.H. Nedreaas, 2001. Age validation and growth of Greenland halibut (Reinhardtius hippoglossoides (Walbaum)): a comparison of populations in the Northwest and Northeast Atlantic. Sarsia, 86:53–68.
  76. Morgan, M.J. and W.R. Bowering, 1997.Temporal and geographic variation in maturity at length and age of Greenland halibut (Reinhardtius hippoglossoides) from the Canadian north-west Atlantic with implications for fisheries management. ICES Journal of Marine Science, 54:875–885.
  77. Carscadden, J.E.,W.A. Montevecchi, G.K. Davoren and B.S. Nakashima, 2002.Trophic relationships among capelin (Mallotus villosus) and seabirds in a changing ecosystem. ICES Journal of Marine Science, 59:1027–1033.
  78. Carscadden, J.E. and K.T. Frank, 2002.Temporal variability in the condition factors of Newfoundland capelin (Mallotus villosus) during the past two decades. ICES Journal of Marine Science, 59:950–958.
  79. Nakashima, B.S. and C.T.Taggart, 2002.Why are some beaches better for spawning than others? ICES Journal of Marine Science, 59:897–908.-- Vilhjálmsson, H., 1994.The Icelandic capelin stock. Capelin Mallotus villosus (Muller) in the Iceland, Greenland, Jan Mayen area. Rit Fiskideildar, 13(1), 281p.
  80. Nakashima, B.S. and J.P.Wheeler, 2002. Capelin (Mallotus villosus) spawning behaviour in Newfoundland waters – the interaction between beach and demersal spawning. ICES Journal of Marine Science, 59:909–916.
  81. Davenport, J., 1989.The effects of salinity and low temperature on eggs of the Icelandic capelin Mallotus villosus. Journal of the Marine Biological Association of the United Kingdom, 69:1–9.-- Davenport, J. and A. Stene, 1986. Freezing resistance, temperature and salinity tolerance in eggs, larvae and adults of the capelin, Mallotus villosus, from Balsfjord. Journal of the Marine Biological Association of the United Kingdom, 66:145–157.
  82. Frank, K.T. and W.C. Leggett, 1981. Prediction of egg development and mortality rates in capelin (Mallotus villosus) from meteorological, hydrographic, and biological factors. Canadian Journal of Fisheries and Aquatic Sciences, 38:1327–1338.
  83. Nakashima, B.S. and J.P.Wheeler, 2002. Capelin (Mallotus villosus) spawning behaviour in Newfoundland waters – the interaction between beach and demersal spawning. ICES Journal of Marine Science, 59:909–916.
  84. Carscadden, J.E., B.S. Nakashima and K.T. Frank, 1997. Effects of fish length and temperature on the timing of peak spawning in capelin (Mallotus villosus). Canadian Journal of Fisheries and Aquatic Sciences, 54:781–787.
  85. summarized by: Frank, K.T., J.E. Carscadden and J.E. Simon, 1996. Recent excursions of capelin (Mallotus villosus) to the Scotian Shelf and Flemish Cap during anomalous hydrographic conditions. Canadian Journal of Fisheries and Aquatic Sciences, 53:1473–1486.
  86. Frank, K.T., J.E. Carscadden and J.E. Simon, 1996. Recent excursions of capelin (Mallotus villosus) to the Scotian Shelf and Flemish Cap during anomalous hydrographic conditions. Canadian Journal of Fisheries and Aquatic Sciences, 53:1473–1486.
  87. Shackell, N.L., J.E. Carscadden and D.S. Miller, 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.
  88. Mowbray, F.K., 2002. Changes in the vertical distribution of capelin (Mallotus villosus) off Newfoundland. ICES Journal of Marine Science, 59:942–949.-- Shackell, N.L., J.E. Carscadden and D.S. Miller, 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.
  89. Mowbray, F.K., 2002, Op. cit.
  90. Carscadden, J.E., K.T. Frank and W.C. Leggett, 2001. Ecosystem changes and the effects on capelin (Mallotus villosus), a major forage species. Canadian Journal of Fisheries and Aquatic Sciences, 58:73–85.
  91. Winters, G.H. and J.P.Wheeler, 1987. Recruitment dynamics of springspawning herring in the Northwest Atlantic. Canadian Journal of Fisheries and Aquatic Sciences, 44:882–900.
  92. Lilly, G.R. and M. Simpson, 2000. Distribution and Biomass of Capelin, Arctic Cod and Sand Lance on the Northeast Newfoundland Shelf and Grand Bank as Deduced from Bottomtrawl Surveys. Department of Fisheries and Oceans, Canadian Stock Assessment Secretariat Res. Doc. 2000/091, 40p.-- Lilly, G.R., H. Hop, D.E. Stansbury and C.A. Bishop, 1994. Distribution and abundance of polar cod (Boreogadus saida) off southern Labrador and eastern Newfoundland. ICES C.M.1994/O:6, 21p.
  93. DFO, 2002b. Northern shrimp (Pandalus borealis) - Div. 0B to 3K. Department of Fisheries and Oceans, Canada, Science Stock Status Report C2-05.
  94. Bundy, A., 2001. Fishing on ecosystems: the interplay of fishing and predation in Newfoundland-Labrador. Canadian Journal of Fisheries and Aquatic Sciences, 58:1153–1167.-- Lilly, G.R., D.G. Parsons and D.W. Kulka, 2000.Was the increase in shrimp biomass on the Northeast Newfoundland Shelf a consequence of a release in predation pressure from cod? Journal of Northwest Atlantic Fishery Science, 27:45–61.-- Worm, B. and R.A. Myers, 2003. Meta-analysis of cod-shrimp interactions reveals top-down control in oceanic food webs. Ecology, 84:162–173.
  95. Lilly, G.R., D.G. Parsons and D.W. Kulka, 2000.Was the increase in shrimp biomass on the Northeast Newfoundland Shelf a consequence of a release in predation pressure from cod? Journal of Northwest Atlantic Fishery Science, 27:45–61.
  96. Parsons, D.G. and E.B. Colbourne, 2000. Forecasting fishery performance for northern shrimp (Pandalus borealis) on the Labrador Shelf (NAFO Divisions 2HJ). Journal of Northwest Atlantic Fishery Science, 27:11–20.
  97. Bundy, A., 2001. Op. cit.
  98. DFO, 2002a. Newfoundland and Labrador snow crab. Department of Fisheries and Oceans, Canada, Science Stock Status Report C2-01.
  99. Stenson, G.B., M.O. Hammill, M.C.S. Kingsley, B. Sjare,W.G.Warren and R.A. Myers, 2002. Is there evidence of increased pup production in northwest Atlantic harp seals, Pagophilus groenlandicus? ICES Journal of Marine Science, 59:81–92.
  100. Hammill, M.O. and G.B. Stenson, 2000. Estimated prey consumption by harp seals (Phoca groenlandica), hooded seals (Cystophora cristata), grey seals (Halichoerus grypus) and harbour seals (Phoca vitulina) in Atlantic Canada. Journal of Northwest Atlantic Fishery Science, 26:1–23.
  101. DFO, 2003a. Northern (2J+3KL) Cod. Department of Fisheries and Oceans, Canada, Stock Status Report 2003/018.
  102. Ibid.
  103. Frank, K.T., R.I. Perry and K.F. Drinkwater, 1990. Predicted response of Northwest Atlantic invertebrate and fish stocks to CO2-induced climate change.Transactions of the American Fisheries Society, 119:353–365.-- Shuter et al., 1999
  104. Frank, K.T., R.I. Perry and K.F. Drinkwater, 1990. Predicted response of Northwest Atlantic invertebrate and fish stocks to CO2-induced climate change.Transactions of the American Fisheries Society, 119:353–365.
  105. Ibid.
  106. Shuter, B.J., C.K. Minns, H.A. Regier and J.D. Reist, 1999. Canada country study: Climate impacts and adaptations: fishery sector.Vol. VII, National Sectoral Volume, pp 258–317.
  107. Furevik,T., H. Drange and A. Sorteberg, 2002. Anticipated changes in the Nordic Seas marine climate: scenarios for 2020, 2050 and 2080. Institute of Marine Research, Bergen. Fisken og Havet No. 4.
  108. Ibid.
  109. 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. Rit Fiskideildar, 15(1):7–29.
  110. Shaw, J., R.B.Taylor, D.L. Forbes, M.-H. Ruz and S. Solomon, 1998. Sensitivity of the Coasts of Canada to Sea-level Rise. Geological Survey of Canada Bulletin 505, 79p.
  111. Nakashima, B.S. and J.P.Wheeler, 2002. Capelin (Mallotus villosus) spawning behaviour in Newfoundland waters – the interaction between beach and demersal spawning. ICES Journal of Marine Science, 59:909–916.
  112. G. Stenson, Fisheries and Oceans Canada, St. John’s, pers. comm., 2003
  113. Bundy, A., G.R. Lilly and P.A. Shelton, 2000. A Mass Balance Model of the Newfoundland-Labrador Shelf. Canadian Technical Report of Fisheries and Aquatic Sciences 2310, xiv + 157p.-- Hammill, M.O. and G.B. Stenson, 2000. Estimated prey consumption by harp seals (Phoca groenlandica), hooded seals (Cystophora cristata), grey seals (Halichoerus grypus) and harbour seals (Phoca vitulina) in Atlantic Canada. Journal of Northwest Atlantic Fishery Science, 26:1–23.
  114. DFO, 2003a. Northern (2J+3KL) Cod. Department of Fisheries and Oceans, Canada, Stock Status Report 2003/018
  115. Crawford, R.E., 1992. Life history of the Davis Strait Greenland halibut, with reference to the Cumberland Sound fishery. Canadian Manuscript Report of Fisheries and Aquatic Sciences 2130.-- Pike, D.G., 1994.The Fishery for Greenland Halibut (Reinhardtius hippoglossoides) in Cumberland Sound, Baffin Island, 1987–1992. Canadian Technical Report of Fisheries and Aquatic Sciences 1924.
  116. M.A.Treble, Fisheries and Oceans Canada,Winnipeg, pers. comm., 2003
  117. For a more extensive discussion see: Schrank,W.E., N. Roy, R. Ommer and B. Skoda, 1992. An inshore fishery: A commercially viable industry or an employer of last resort. Ocean Development and International Law, 23:335–367.
  118. Government of Newfoundland and Labrador, 2001. Securing Our FutureTogether: Final Report on the Renewal Strategy for Jobs and Growth. Government of Newfoundland and Labrador, St. John’s, Newfoundland.
  119. Overton, J., 2001. Official acts of vandalism: privatizing Newfoundland’s Provincial Parks in the 1990s. In: D. McGrath (ed.). From Red Ochre to Black Gold. Flanker Press, Newfoundland.
  120. Schrank,W.E. and B. Skoda, 2003.The cost of marine fishery management in eastern Canada: Newfoundland, 1989/1990 to 1999/2000. In:W.E. Schrank, R. Arnason and R. Hannesson (eds.). The Cost of Fisheries Management, pp, 91–128. Ashgate Publishers.
  121. Corbett, F., 2002. Fax message from F. Corbett, Department of Fisheries and Oceans to A. St. Croix, May 27, 2002.
  122. Pushie, G.F. (Chairman), 1967. Report of the Royal Commission on the Economic Prospects of Newfoundland and Labrador. Office of the Queen’s Printer, St. John’s, Newfoundland.
  123. Schrank,W.E., 1995. Extended fisheries jurisdiction: origins of the current crisis in Atlantic Canada’s fisheries. Marine Policy, 19:285–299.
  124. Schrank,W.E., N. Roy and E. Tsoa, 1986. Employment prospects in a commercially viable Newfoundland fishery: An application of ‘An Econometric Model of the Newfoundland Groundfishery’. Marine Resource Economics, 3:237–263.
  125. Government of Canada, 1976a. Policy for Canada’s Commercial Fisheries. Fisheries and Marine Service of the Department of the Environment, Ottawa.-- Government of Canada, 1976b. Community and Employment Implications of Restructuring the Atlantic Fisheries. Analysis and Liaison Branch of the Department of Regional Economic Expansion, Ottawa.
  126. Government of Newfoundland and Labrador, 1978. Setting a Course: A Regional Strategy for Development of the Newfoundland Fishing Industry to 1985. Department of Fisheries, St. John’s, Newfoundland.
  127. Schrank,W.E., 1995. Extended fisheries jurisdiction: origins of the current crisis in Atlantic Canada’s fisheries. Marine Policy, 19:285–299.
  128. LeBlanc, R. and P. de Bané, 1981. A Proposed Memorandum of Understanding between DFO and DREE Respecting Joint Action for Development of the Fisheries of Eastern Canada. Signed by Roméo LeBlanc and Pierre De Bané, March 20, 1981. National Archives of Canada, RG124, Accession 87–88/35, File #4560-F-0 Part 1.
  129. Kirby, M.J.L. (Chairman), 1982. Navigating Troubled Waters – A New Policy for the Atlantic Fisheries. Department of Supply and Services of the Government of Canada, Ottawa.
  130. Schrank,W.E., 1995. Extended fisheries jurisdiction: origins of the current crisis in Atlantic Canada’s fisheries. Marine Policy, 19:285–299.
  131. Alverson, D.L., 1987. A Study of Trends of Cod Stocks off Newfoundland and Factors Influencing their Abundance and Availability to the Inshore Fishery: A Report to the Honorable Tom Siddon, Minister of Fisheries of Canada.Task Group on Newfoundland Inshore Fisheries.
  132. Harris, L. (Chairman), 1989. Independent Review of the State of the Northern Cod Stock. Department of Fisheries and Oceans, Ottawa.-- Harris, I. (Chairman), 1990. Independent Review of the State of the Northern Cod Stock: Final Report of the Northern Cod Review Panel. Department of Supply and Services, Ottawa.
  133. see: Schrank,W.E., 1997.The Newfoundland fishery: past, present and future. In S. Burns (ed.). Subsidies and Depletion of World Fisheries: Case Studies.World Wildlife Fund,Washington, D.C.
  134. Ibid.
  135. Ibid.
  136. Statistics Canada, 2002a. www12.statcan.ca/english/census01/products/standard/popdwell/Table-PR.cfr (April 18, 2005)
  137. Statistics Canada, 2002b. 2001 Community Profiles. ww12.statcan.ca/english/profil01/PlaceSearchForm1.cfm (March 18, 2003).
  138. Ibid.
  139. Schrank,W.E., B. Skoda, P. Parsons and N. Roy, 1995.The cost to government of maintaining a commercially unviable fishery: the case of Newfoundland 1981/82 to 1990/91. Ocean Development and International Law, 26:357–390.
  140. CBC, 2002. Item broadcast on The Morning Show in St. John’s, Newfoundland on Canadian Broadcasting Company Radio 1 in the half hour starting at 7:00 a.m.Wednesday, April 24, 2002.