You are on page 1of 36

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/228650264

PARADiGMs iN FishERiEs ocEANoGRAPhy

Article · January 2008

CITATIONS READS
53 459

7 authors, including:

Francis In William Leggett


Monash University (Australia) Queen's University
100 PUBLICATIONS   4,216 CITATIONS    212 PUBLICATIONS   12,076 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Fish Recruitment View project

Ecological Stability,Compensation and Core Areas View project

All content following this page was uploaded by William Leggett on 07 February 2016.

The user has requested enhancement of the downloaded file.


Oceanography and Marine Biology: An Annual Review, 2008, 46, 331-363
© R. N. Gibson, R. J. A. Atkinson, and J. D. M. Gordon, Editors
Taylor & Francis

Paradigms in fisheries oceanography


William C. Leggett1 & Kenneth T. Frank2
1Department of Biology, Queen’s University, Kingston, ON K7L 3N6, Canada

E-mail: wleggett@queensu.ca
2Department of Fisheries and Oceans, Bedford Institute

of Oceanography, Dartmouth NS B2Y 4A2, Canada


E-mail: FrankK@mar.dfo-mpo.gc.ca

Abstract  The development of the field of fisheries oceanography over the past century has
been heavily influenced by a relatively small number of paradigms that have shaped thinking,
influenced lines of enquiry and occasionally stalled progress in the field. This review provides
an overview of what are considered to be the most influential paradigms in the discipline. Each
begins with a brief discussion of its origins. Next their respective (and often overlapping) impact
on the development of the discipline is discussed and then the evolution of these paradigms
as shaped by new advances in approaches and technologies and by direct challenges to their
underlying assumptions is reviewed. For each, the endpoint is an overview of the current state
of knowledge and thinking and the probable future direction of research in the area. The review
concludes with an overview of the probable future directions of research in the discipline as a
whole.

Introduction
The discipline of modern fisheries oceanography has its origins in the work of Johan Hjort (1914,
1928), who was the first to formally hypothesize a link between the dynamics of fish populations
and the dynamics of their environment. In his ‘critical period’ paradigm Hjort argued that variabil-
ity in food availability during the transition from endogenous to exogenous feeding in larval fishes,
typically a very narrow time window, was central to the survival of individual larval cohorts. Hjort
hypothesized that when food was abundant, survival (and recruitment) would be high, and when
food was scarce survival and recruitment would be low. This hypothesis was subsequently general-
ized by Cushing (1975) who reasoned, in his ‘match mismatch’ hypothesis, that food availability
was linked to the interaction between interannual differences in the timing of spawning/hatching
and the timing and magnitude of the primary and secondary production cycles in the ocean. These
hypotheses dramatically influenced the direction of research in the field by (1) focusing research into
the causes of interannual variability in recruitment on the egg and larval stages of fishes, (2) provid-
ing a simplifying construct within which to explore the causes of temporal changes in the abun-
dance of commercially important marine fishes and (3) by linking fluctuations in the abundance of
fishes directly to the dynamics of other components of the ocean ecosystem. In hindsight, Hjort’s
major contribution appears to have been to awaken thinking and research into the nature of this
dynamic interaction between fish and their environment. Prolonged adherence to Hjort’s ideas, and
the lure of Cushing’s hypothesis, combined to dominate thinking and research in the field for most
of the twentieth century, some would say negatively (Leggett & DeBlois 1994). As the discipline has
advanced several new paradigms have evolved. While originally offered as simplifying constructs,

331
WILLIAM C. LEGGETT & KENNETH T. FRANK

Table 1  Paradigms in fisheries oceanography


1. Spawning stock biomass (SSB) is a suitable proxy for the reproductive potential of a stock.
2. Marine fish eggs and larvae are generally designed for dispersion and potential colonization (panmixia).
3. In marine temperate systems, fish spawn in springtime so that peak larval abundance coincides with maximum prey
availability (Cushing’s match/mismatch hypothesis).
4. Environmentally based recruitment models, when updated with new data, invariably fail; recruitment prediction is an
intractable problem, particularly when it is based on processes associated with the growth and mortality of the early
life-history stages.
5. Populations cannot irreversibly collapse/collapsed populations will recover in the absence of fishing.
6. Fish stocks can be managed in isolation from their total environment/habitat.
7. Population recovery is synonymous with rebuilding.

and as the foundation for new studies of important processes governing the dynamics of fish stocks,
some of these new paradigms now approach the status of dogma once accorded Hjort’s hypothesis
and have the potential, if pursued uncritically, to once again delay progress in the field. This review
examines the major paradigms that have shaped thinking and research in fisheries oceanography
over the past century and explore how recent research has affected understanding of their validity
and usefulness now, and going forward. Table 1 provides an overview of these paradigms ordered
hierarchically from individuals to ecosystems.
The need for a more holistic and effective approach to the science of fisheries oceanography,
and of the management strategies applied to marine fishes, is evident from the breadth of species
now undergoing serious decline, not only worldwide, but also within more restricted ecosystems
(Dower et al. 2000, Figure 1).
The advent of new methods (ageing techniques for larval fishes, the application of biochemical
and molecular techniques, the development of instrumentation and computing capacity that has
allowed physical oceanographers to measure and model the highly dynamic and variable environ-
ment of continental shelf ecosystems, satellite imagery, etc.) have created new opportunities to
explore the dynamics of marine fishes in relation to the dynamics of their physical environment and
the ecosystems they inhabit.

Paradigm 1: Spawning stock biomass is a suitable


proxy for the reproductive potential of a stock
The Hjort and Cushing models focused primarily on the factors determining interannual variability
in mortality rates experienced by larvae (the life stage widely believed to be the most vulnerable to
environmental variability (reviewed in Leggett & DeBlois 1994). Another highly influential para-
digm holds that it is the product of the number of offspring generated by a spawning stock and the
rate of mortality experienced by those offspring until they recruit to the reproducing population
that govern the dynamics of fish populations. This thinking is inherent in stock recruitment models
(Ricker 1954, Beverton & Holt 1957) that have formed the cornerstone of studies of the recruitment
dynamics of fishes, and of management based on these studies, for most of the latter half of the
twentieth century. These models assume strong density dependence (Figure 2).
Fundamental to the practical application of these models has been the assumption that the total
spawning stock biomass (SSB) is an acceptable proxy for the reproductive potential of the stock of
interest. This assumption is employed because SSB can be readily derived from fisheries survey
data, whereas total egg production (TEP), a more reliable and realistic indicator of reproductive
output is more labour intensive to obtain and, for this reason, does not exist for most stocks.
The use of SSB as a proxy for TEP implies that (1) spawner biomass is proportional to the TEP
and (2) in lay terms, “an egg is an egg is an egg,” that is, all eggs are equal in their potential to

332
PARADIGMS IN FISHERIES OCEANOGRAPHy

herring 4VWX-SSB
yellowtail flounder 5Z-SSB
turbot 4RST-SSB
little skate 4VWX-SSB
winter skate 4X-SSB
smooth skate 4VWX-SSB
smooth skate 4T-SSB
cod 4T-SSB
cod 4X-SSB
silver hake 4VWX-SSB
cod 5Z-SSB
winter skate 4VsW-SSB
haddock 4VW-SSB
cod 4Vn-SSB
mackerel-SSB
winter skate 4T-SSB
haddock 4X-SSB
plaice 3Ps-SSB
haddock 5Z-SSB
thorny skate 4T-SSB
thorny skate 4VWX-SSB
cod 4VsW-SSB
pollack 4VWX-SSB
white hake 4T-SSB
cod 2J3KL-Offhore Survey-SSB
cod 3NO spring-SSB
plaice 2J3K-SSB
plaice 4VT-SSB
1970 1975 1980 1985 1990 1995 2000
Years

Figure 1  (See also Colour Figure 1 in the insert following p. 250.) Ordination of the time of series of spawn-
ing stock biomass of various species from scientific surveys conducted throughout the north-west Atlantic,
illustrating that the majority of the stocks are at biomass levels well below (red) the long-term average.
Intensity of colours is proportional to the magnitude of the standardized anomaly in standard deviation units.
Alphanumeric labelling refers to the Northwest Atlantic Fisheries Organization management unit.

produce a recruit to the population. However, the variation in SSB is generally insufficient to account
for the related variation observed in recruitment in marine fish populations (Shepherd et al. 1984,
Wooster & Bailey 1989, Rijnsdorp et al. 1991, Marshall et al. 1998). Notwithstanding this reality,
the paradigm that SSB is a suitable proxy for total reproductive effort has persisted and has, in turn,
been an important contributor to the persistence of the Hjort and Cushing hypotheses and the focus
of research into the causes of recruitment variation in marine fishes. This includes the search for
ecological and/or environmental factors affecting egg and larval mortality — an approach that has
proved challenging and, in only limited cases, successful (see Paradigm 4). This has impeded the
development of a coherent understanding of the recruitment dynamics of marine fishes in particular
and the science of fisheries management in general (Rothschild 1986, Hilborn & Walters 1992).

333
WILLIAM C. LEGGETT & KENNETH T. FRANK

North Sea herring

40

Recruitment
20

0
1 2 3
SSB

Figure 2  Ricker stock recruitment model fit to data for North Sea herring. SSB, spawning stock biomass.

In the 1990s these assumptions began to come under increasing scrutiny (Lowerre-Barbieri
et al. 1998, Marshall et al. 1998, Marteinsdottir & Steinarsson 1998, Marshall & Frank 1999,
Marshall et al. 1999). Marshall et al. (1998) were the first to demonstrate in situ the potential fallacy
of the assumption inherent in the use of SSB as a surrogate for TEP. Their work with Barents Sea
cod (Gadus morhua) showed that recruitment was positively correlated with the quantity of lipid
energy stored in the liver of mature females or, in other words, with female condition. Boyd et al.
(1998) observed a similar relationship between parental fat levels and recruitment in Cape anchovy
(Engraulis capensis). In cod, total liver energy was proportional to TEP and, like TEP, varied with
capelin (Mallotus villosus) abundance (Yaragina & Marshall 2000) such that by either measure the
reproductive potential of a fixed number of mature females was significantly higher when the quan-
tity of food available to maturing fish was abundant and correspondingly lower when it was scarce.
In contrast, cod SSB was found to be not statistically different at high and low prey levels. In short,
their study confirmed prior suspicions regarding the paradigm that SSB is an inadequate surrogate
for TEP and demonstrated that replacing SSB with more accurate measures of reproductive poten-
tial is fundamental to a fuller understanding of the dynamics of recruitment in marine fishes.
The work of Marshall et al. (1998, 1999) has been an important factor in redirecting the focus
of studies into the causes of interannual variation of recruitment in marine fish populations. Their
findings demonstrated that a more successful pursuit of process-oriented models of the recruitment
dynamics of marine fish will require integration of the processes affecting reproductive output
(growth, production and condition of spawners) with mortality processes affecting egg, larval and
juvenile survival (Ulltang 1996, Marshall et al. 2000).
Research by Yaragina & Marshall (2000) has demonstrated that the search for a comprehensive
understanding of the factors regulating variability in reproductive potential, independent of SSB,
may be as complex, and as inextricably tied to environmental and ecological factors, as are the
causes of variation in egg and larval survival. For example, their study of temporal variation in the
liver condition index (LCI) — an important determinant of reproductive potential (Marshall et al.
1999) — of five length classes of north-east Arctic cod (Gadus morhua) showed that while varia-
tions in the abundance of capelin, a major food source, was the proximate determinant of variation
in LCI, an indirect, but perhaps not ultimate, determinant of variation in the index appears to have
been the abundance of herring (Clupea harengus), which influence cod LCI indirectly via their
predation on capelin, which in turn are the main prey of cod.
Interest in the role and importance of maternal effects as regulators of the recruitment dynamics
of marine fishes has increased since the publication of these findings. Scott et al. (2006) modelled
the daily reproductive output of a range of simulated age/size-structured populations of Atlantic

334
PARADIGMS IN FISHERIES OCEANOGRAPHy

cod, created under contrasting stock recruitment scenarios, over an entire spawning season. The
objective was to determine the effects of individual female condition and egg quality on stock repro-
ductive potential (SRP). Their findings suggest that, in two populations having equal SSB, the effect
of low levels of individual condition in one can lead to almost total reproductive failure for that
population. Moreover, the positive effect of increased individual condition was found to depend on
the particular population structure designated in the model, a variable that, in situ, can be strongly
influenced by both fishing intensity and environment. Indeed, they reported that differences in
population size and age structure produced by differences in fishing mortalities ranging from 0 to
1.0 could reduce the SRP by 48–74% depending on the related assumptions regarding female con-
dition and quality. Any factor, anthropogenic or environmental, inducing changes in mortality of
equivalent magnitude would be expected to produce a similarly scaled result.
Vallin & Nissling (2000) studied the effect of female age on the survival of cod eggs and lar-
vae in the Baltic, where successful spawning is restricted to the deep basins at salinities varying
between 11 and 20 ppm. Due to the oxygen depletion commonly prevailing in these areas, neutral
egg buoyancies above oxygen-critical depths are important to egg survival. They found that large
females produce larger eggs that exhibit neutral egg buoyancy at lower salinities, thereby ensuring
egg development in more favourable oxygen conditions. The number of recruited cod (age 2), in
the Baltic Sea over two eras characterized by different conditions (1967–1980 and 1981–1994), was
found to be positively related to the fraction of eggs produced by older females (5+ yr), implying a
strong maternal effect on recruitment. Berkeley et al. (2004) report a similar positive relationship
between female age and probability of egg survival in longnose grenadier (Coelorhynchus carmina-
tus). Marteinsdottir & Steinarsson (1998) report similar findings for the egg and early larval stages
of cod.
The findings of Vallin & Nissling (2000) also illustrate the potential for negative effects of
fishery-induced changes in population age structure on population size and persistence. Scott et al.
(1999) modelled this fishing effect and found that the effects of the loss of more fecund older/larger
individuals in the population could lead to overestimation of the number of potential recruits to
populations experiencing higher levels of fishing mortality by as much as 60%.
Environmental variables can also influence the condition of females and the demographics of
populations in ways that can dramatically alter reproductive output independent of SSB. Scott et al.
(1999) note that when size- (growth-) related maternal effects on egg viability were incorporated
into their fishing effect model, the number of potential recruits to heavily exploited populations
could be reduced by a further 10%.
The work of Choi et al. (2004) illustrates just how profound these environmental influences
can be. In their paper on the devolution of the Scotian Shelf ecosystem off Nova Scotia, Canada,
they document declines in growth rates, size at age and age at maturation of the entire benthic fish
community that mirror the changes modelled and documented above. These changes occurred pro-
gressively over a 45-yr time period (1960–2005), the most dramatic changes occurring in the early
1990s. They resulted in average sizes of mature (age 5) cod, haddock (Melanogrammus aeglefinus),
pollock (Pollachius virens) and silver hake (Merluccius bilinearis) that were 70–80% of the 1970s
levels and remain depressed in spite of a moratorium on fishing that began in 1993 (Figure  3).
Drinkwater (2002) estimates that approximately 30–50% of the decline in SSB of cod that occurred
in the 1980s and early 1990s was linked to these reductions in deep-water temperatures and their
depressing effect on growth rates and corresponding sizes at age. Given the strong positive relation-
ship between body size and fecundity in these species (Pinhorn 1984, Waiwood & Buzeta 1989), a
corresponding decline in per individual reproductive potential, independent of losses due to coinci-
dent reductions in the total number of spawners, clearly occurred.

335
WILLIAM C. LEGGETT & KENNETH T. FRANK

70
Cod
Haddock
Pollock
60
Length at age 5, cm Silver hake

50

40

30

20
1970 1975 1980 1985 1990 1995 2000

Figure 3  (See also Colour Figure 3 in the insert.) Changes with time in lengths at age 5 for four groundfish
species from the eastern Scotian Shelf.

Their analysis also revealed that these dramatic changes in reproductive output were not simply
the consequence of reductions in the size of reproducing adults (a product of the negative effects
that quotas based on biomass and the behaviour of fishermen focusing their efforts on the largest
fish available; e.g., see Drinkwater 2002). For example, during the 1960s and 1970s the physiologi-
cal condition of these fishes exceeded the 45-yr norm. However, throughout the 1980s and 1990s
physiological condition declined sharply and by 2000 the majority of the fish were in below-average
condition. Given the documented impact of female condition on egg quality and recruitment in
cod (Marshall et al. 1999), and the evidence of maternal effects on survival linked to female size
and condition, it is likely that in this species, and perhaps others, a further reduction in effective
reproductive output, independent of SSB, was experienced as a consequence of combined effects of
declining size and condition, both of which had an environmental component.
There is also evidence that the total metabolic rate of the entire demersal ecosystem was
depressed as a consequence of changes in ocean climate. These ocean climate changes included
a progressive decline in bottom temperatures, an increase in the volume of the cold intermediate
layer, and movement of the Gulf Stream Front to a more offshore position. These changes are judged
to have been caused by an increased along-shelf advection from the Gulf of St. Lawrence and south-
ern Newfoundland augmented by local, atmospherically induced cooling (Drinkwater et al. 2003).
One of the outcomes of this change was a dramatic reduction in the biomass of the once-dominant
euphausiid (Euphausia superba) and a corresponding decline in their contribution to the overall diet
of the demersal fish assemblage. For example, the contribution of this euphausid species to the diet
of pollock and cod declined from >65% by weight in the 1980s to <10% in the 1990s (Hanson &
Chouinard 2002, Carruthers et al. 2005).
The resulting hysteresis in the structure of the Scotian Shelf ecosystem in the late 1980s pro-
duced an ecosystem dominated by smaller pelagic fishes and benthic invertebrates (mainly shrimp
and crab) in which the once-dominant large-bodied demersal fishes now play a relatively minor role.
The fact that these once-dominant species have not recovered even in the absence of exploitation,

336
PARADIGMS IN FISHERIES OCEANOGRAPHy

as traditional stock recruitment models would predict, suggests that a fundamental change in their
fitness has occurred. The disjunction between SSB and effective reproductive output appears to be a
significant factor in this change in fitness and may have contributed in a major way to the dramatic
collapse of these once-dominant demersal species in the early 1990s.

Paradigm 2: Marine fish eggs and larvae are generally


designed for dispersion and potential colonization (panmixia)
Reproduction in commercially exploited marine species generally involves the spawning of billions
of tiny eggs that float in the near-surface waters during the spring of each year. The generality and
magnitude of these events were widely considered to be an adaptation to the very high mortality of
early life stages caused by the vagaries of the environment (Rothschild 1986, Rothschild & DiNardo
1987). One variable identified as important in this context was ocean currents, often driven by wind,
which caused larvae to be transported to regions where food and other factors were less favourable.
Several correlation studies reported significant relationships between larval survival, measured by
the abundance of the recruiting year-class, and dispersive events
For example, it has been hypothesized that warm core rings (eddies representing instabilities in
the Gulf Stream that can entrain large volumes of water from the continental shelf) can transport
sufficient numbers of eggs and larvae off the shelf to have a significant negative impact on recruit-
ment. Myers & Drinkwater (1989), who examined this hypothesis using estimates of entrainment
from 14 yr of satellite imagery and recruitment estimates for 25 fish and shellfish stocks ranging
from the Mid-Atlantic Bight to the southern Grand Banks, found that increased warm core ring
activity was associated with low recruitment levels in 17 of 18 groundfish stocks examined. The sole
exception was cod on Georges Bank. While the level of significance related to each stock was low,
the collective result was consistent with the original hypothesis. For other examples of the possible
link between dispersive effects and recruitment see Carruthers et al. (1951) and Bailey (1981). The
findings of these and other studies led to the view that either excessive dispersion of larvae or dis-
placement to areas distant from the nursery ground was a primary cause of the reduced recruitment
that resulted when wind strengths and direction were unfavourable. However, in most cases updated
analyses of these models have failed to support the relationships originally observed (Myers 1998,
see Paradigm 4).
The seminal work of Sinclair (1988) showed that many continental shelf spawning locations of
marine fishes were located in areas characterized by retention features or semipermanent gyres, the
current characteristics of which act to minimize the advection of the egg and larval stages and con-
sequently their vulnerability to dispersive processes. Many subsequent studies have expanded the
suite of species that utilize relatively non-dispersive oceanographic settings for spawning and early
larval development (Hinrichsen et al. 2001, North & Houde 2001, Lett et al. 2007). In addition,
the advent of technologies allowing high-resolution, discrete-depth sampling of the water column
revealed that ontogenetic shifts in depth distributions and vertical migratory behaviour of larvae
also serve to minimize dispersion from the spawning areas. These findings have moderated the
belief that the early life stages of marine species were simply passive drifters and that survival was
largely dependent on the “mercy” of the currents.
One of the most striking examples of this non-dispersive reality involves haddock that spawn
on the offshore banks of the Scotian Shelf. Because of the recirculation that occurs around each
of these discrete spawning banks there is a strong tendency for the egg and larval distributions of
haddock that spawn there to be discrete. And, when larvae metamorphose to the juvenile stage, a
development process that takes about 90 days, settlement to the bottom occurs in the same zones in
which they were spawned, which now become prime feeding areas for the juvenile and adult stages
(Frank et al. 2000).

337
WILLIAM C. LEGGETT & KENNETH T. FRANK

These findings are consistent with the results of modelling studies of the scales of connectivity
in Caribbean reef fishes (Cowen et al. 2006). They found that larval dispersal scales for most species
were of the order of 50–100 km, much smaller than has generally been assumed. Their analyses also
indicate that passive dispersal/retention is insufficient for population replenishment, recruitment
levels generated from the use of passive dispersal assumptions in the model being one to two orders
of magnitude lower than that required for successful population replenishment.
These studies call for a general rethinking of the adaptive character of the spawning behaviour
of many marine species from spawning as a vehicle for colonization or bet hedging to one in which
the entire early life-history period is viewed as an elegant suite of adaptations to an environment that
is highly variable at interannual and smaller timescales but is more predictable in the longer term.
However, while this longer-term predictability is reflected in adaptations involving both the
timing and location of spawning, interannual variability in the ocean environment can result in
important disruptions to their efficacy that lead to recruitment variability and even recruitment fail-
ure. Well-known examples include the relationship between capelin recruitment and onshore wind
discovered by Leggett et al. (1984), the destabilizing effects of warm core rings on the retention
dynamics of Scotian Shelf banks (Myers & Drinkwater 1989) and the influence of upwelling events
on recruitment processes along the North African coast (Cury & Roy 1989). Such disruptive events
can also lead to the establishment of new subpopulations through colonization of new habitats when
environmental conditions permit. Detection of such colonization episodes is most evident when they
occur outside the normal range of distribution. The larval drift associated with anomalous oceano-
graphic conditions involving the displacement of cod larvae from Iceland to West Greenland — a
colonization event covering a distance of almost 1000 km (Frank 1992) — is a striking case in point.
A similar displacement, on a much smaller geographic scale, was shown for haddock larvae originat-
ing from the eastern Scotian Shelf. Larvae of this species episodically drift downstream to Browns
Bank on the western Scotian Shelf. In the case of both cod and haddock, the dispersed larval stages
appear to persist and to contribute to the recruiting year-class in the colonized area from which they
then make a subsequent return to their spawning area during the maturation phase (Frank 1992,
Brickman 2003). Similar events appear to be a characteristic of many Caribbean fishes (Cowen et al.
2006). Discovery of this phenomenon has resulted in the need for significant upward revisions in the
estimates of year-class strength at the natal site and a corresponding bonus to the local fishery. Its
implication for the assessment and management of adjacent stocks is also profound.

Paradigm 3: In marine temperate systems, fish


spawn in springtime so that peak larval abundance
coincides with maximum prey availability
(Cushing’s match/mismatch hypothesis)
The Hjort and Cushing hypotheses, which were so instrumental in directing the early research into
the population dynamics of marine fishes, were founded on the dynamic interaction between the
temporal and spatial distributions of larval fishes and their prey. Leggett & DeBlois (1994) system-
atically reviewed the scientific evidence for and against these hypotheses and found Hjort’s hypoth-
esis to be wanting. Support for Cushing’s hypothesis was judged equivocal, mainly because of the
difficulty of adequately operationalising the hypothesis and the technical challenges of assembling
data on appropriate time and space scales, realities acknowledged by Cushing himself in his update
of the hypothesis (Cushing 1990).
An important feature of the Cushing hypothesis was its removal of the restriction inherent
in Hjort’s thesis that food-mediated mortality would be restricted to a brief ‘critical’ stage in lar-
val development (the transition from endogenous to exogenous feeding). Under the Cushing model

338
PARADIGMS IN FISHERIES OCEANOGRAPHy

recruitment success became associated (at least in theory) with the abundance of food during the
entire period of larval development. One of the explicit corollaries of this thinking was the expecta-
tion that spawning in north temperate marine fishes would be adaptively linked, temporally, to the
annual spring (and in some cases fall) production cycle(s) in the sea. This corollary also determined
and directed much of the research into the link between ocean physics, primary and secondary
production, and the recruitment dynamics of fish populations.
Solid evidence of the importance of this link to recruitment has been elusive (Cushing 1990,
Leggett & DeBlois 1994). Two papers highlight the contribution of advances in the understanding
of physical/biological linkages in the ocean and the availability of more sophisticated sampling
technologies, in particular satellite remote sensing, to an understanding of the potential importance
of Cushing’s ideas.
Mueter et al. (2006) utilized established relationships between the dynamics of the mixed layer
and the onset of primary production to develop a 35+-yr time series for the timing of the spring
bloom in the Bering Sea in an attempt to explain recruitment variation in walleye pollock (Theragra
chalcogramma). This indirect measure of the onset of spring primary production conformed closely
to the onset as determined from 6 yr of fluorescence data. The production series so developed was
then related to indices of walleye pollock survival over the same time interval. The indices were
expressed as the residuals from a Ricker stock recruitment model. Walleye pollock survival rates
were strongly and inversely related to the timing of the spring bloom dates, a relationship inter-
preted by the authors as evidence of the important role of the timing of the spring bloom in relation
to the timing of spawning as a determinant of larval survival and recruitment in this species. While
possibly a case of overinterpretation due to the use of indirect approximations of the independent
and dependent variables, the findings are nonetheless supportive of the Cushing hypothesis.
In a more direct and convincing study, Platt et al. (2003) employed satellite remote sensing to
determine the timing of the spring phytoplankton peak on the eastern Scotian Shelf. This they related
to peaks in the production of larval haddock. While their data series was short (8 yr) and was divided
into two distinct periods (1979–1981 and 1997–2001) their analyses revealed a strong relationship
(r2 = 0.89) between variation in recruitment and variation in the timing of the spring bloom.
Unresolved by both studies is the extent to which other components of the ecosystem, perhaps
themselves linked to and affected by physical attributes of the ecosystem that are co-related to the
timing of the spring bloom (predator abundance and diversity, temperature effects, etc.), might
influence the environment occupied by eggs and larvae (either positively or negatively) and the
extent to which this result can be generalized to other areas and other species.
In contrast to these findings, Mousseau et al. (1998), who examined the annual cycles of abun-
dance of fish larvae and their zooplankton prey in relation to the biomass and production of phyto-
plankton on the Scotian Shelf, discovered that the production of copepod nauplii and copepodites
was sustained throughout the year and that fish larvae specializing on copepod prey also occurred
year-round. This occurred notwithstanding the fact that the spring bloom of large phytoplankton,
normally assumed to form the basis of the food web for larval fishes, was restricted to February–
April. This year-round food availability was produced by a non-peak food web structured on the
large-microphage shunt of the microbial food web (small phytoplankton → appendicularians/­
pteropods → fish larvae). Mousseau et al. (1998) concluded that the year-round presence of fish lar-
vae and the fact that several of their major prey items exploit the microbial food web challenges the
long-standing belief that the feeding of marine fish larvae depends primarily on the reproduction of
herbivorous calanoid copepods grazing the spring and autumn blooms of large phytoplankton.
De Figueiredo et al. (2005) provide further support for this link between larval feeding and the
large microphage shunt of the microbial food web. They found that Protozoa and appropriately sized
metazoan prey, previously largely ignored as potential food items, can contribute significantly to
the dietary and energy requirements of larval fish. They conclude that the inclusion of these dietary

339
WILLIAM C. LEGGETT & KENNETH T. FRANK

A B C

Larval mortality

Body size Stage duration Growth rate

Figure 4  Graphical representation of the ‘bigger is better’ (A), ‘stage duration’ (B) and ‘growth-selective’ (C)
hypotheses.

components into measures of the food resource available to larval fishes diminishes the validity
of the key assumption underlying the Cushing hypotheses — that is, when, at times other than the
spring bloom, food levels in the sea are limiting to larval growth and survival.
There is mounting evidence, as well, that at larger scales, water temperature may play a greater
role in mediating larval growth and survival than food availability (see Meekan et al. 2003, Takasuka
& Aoki 2006). This, too, raises the questions, what other components of the ecosystem (predator densi-
ties, types, sizes, feeding rates, etc.) are influenced by physical factors linked to the timing of the spring
bloom, and how do these ecosystem level interactions influence larval survival and recruitment?
A second corollary of the Cushing hypothesis is that larvae that experience superior feeding
conditions, by virtue of their association with areas/times of enhanced food abundance, should
exhibit faster growth and higher survival, leading ultimately to improved recruitment. This corol-
lary, the so-called ‘growth-mortality’ hypothesis (Anderson 1988, Hare & Cowen 1997) is, in fact,
comprised of three non-exclusive functional hypotheses: the ‘bigger is better’ hypothesis, the ‘stage
duration’ hypothesis and the ‘growth-selective’ hypothesis (Figure 4).
Underlying all three hypotheses is the general assumption that following the transition to exog-
enous feeding the risk of death to an individual will be inversely related to the quantity and quality
of food available and, by extension, to rates of growth and development. The positive effect of high
growth and development rates on survival is generally believed to result from the rapid increase in
length and/or changes in behaviour related to development during this period that, in turn, differen-
tially influences the susceptibility of individual larvae to predation (reviewed in Litvak & Leggett
1992, Leggett & DeBlois 1994).

‘Bigger is better’ hypothesis


The ‘bigger is better’ hypothesis evolved, in large part, from the results of general models that
aggregate data at the species level (an approach criticized by Pepin & Miller (1993) for its distorting
effects when generalized to intraspecific studies) and, in part, from the results of laboratory studies
that showed larger larvae were less vulnerable to predation. However, as demonstrated by Litvak &
Leggett (1992) the experimental designs typically used in these studies to assess the relationship
between size and vulnerability commonly compounded the effects of age and size by using larvae
of different ages to obtain the desired range of sizes investigated. Furthermore, most studies of the
effects of size and/or age failed to provide the predator with a choice of prey sizes/ages and there-
fore examined only the capture component of the predation act, whereas predation involves three
multiplicative probabilities: encounter, attack and capture (reviewed in Litvak & Leggett 1992).
Laboratory and in situ mesocosm experiments in which larvae of identical ages but different sizes
and of identical sizes but different ages were subjected to predation by both visual and non-visual
predators clearly illustrated this bias (Litvak & Leggett 1992). Contrary to the predictions of the big-
ger is better model, in predation trials involving experimental cohorts of larvae of identical age, but

340
PARADIGMS IN FISHERIES OCEANOGRAPHy

possessed of a distribution of sizes, both non-visual (jellyfish) and visual (fish) predators consumed
disproportionately more larger larvae. And, when these predators were offered cohorts of larvae of
identical sizes but possessing a distribution of ages, older larvae were selectively consumed. This
also is contrary to expectation. Pepin et al. (1992) report similar findings. Bertram & Leggett (1994)
found no difference in predation risk when metamorphosing winter flounder (Pseudopleuronectes
americanus) of like age but different sizes and of like size but different ages were subjected to
predation by Crangon sp. The principal difference in all these experiments, relative to those from
which the bigger is better model was developed, is that the predators were offered a choice of larvae
of different sizes or ages upon which to prey, as would occur in nature.
As noted by Leggett & DeBlois (1994), the findings reported by Litvak & Leggett (1992) and
supported by Pepin et al. (1992) and Bertram & Leggett (1994) suggest that, contrary to predictions
of the bigger is better model, for individual members of a larval cohort being smaller at a given age
could, under some conditions, and at certain stages of development, confer a survival advantage.
This difference in perspective derives from an explicit recognition of the reality that, in nature,
individual larvae exist, and are preyed upon, as members of a population characterized by evolving
distributions of ages, sizes and sizes at age. Supporting field evidence is provided by the work of
Fortier & Quinonez-Velazquez (1998), who studied the hatch date frequency distributions and daily
growth rates of pollock and haddock larvae on the Scotian Shelf in 1992 and 1993. They found
that, in haddock, growth varied little over the hatching season and that there was no significant
relationship between growth and survival. In pollock, slow growth invariably resulted in low sur-
vival, but fast growth resulted in both low and high survival of individual daily cohorts. Fortier &
Quinonez-Velazquez (1998) concluded that fast growth was a contributing but insufficient condition
for enhanced survival and that for both species investigated, increased predation late in the hatch-
ing season could decouple growth and survival. It appears, therefore, that differences in individual
vulnerability to predation, and the population effects of this variability on recruitment, will be a
function of the evolving structure of the distributions of larval size, age, and size at age throughout
the larval period and of the predators’ response to them. Clearly, many of the inferences initially
drawn from the bigger is better model require rethinking.

‘Stage duration’ hypothesis


The ‘stage duration’ hypothesis was advanced simultaneously by Houde (1987) and Chambers &
Leggett (1987). This model is founded on the observations that (1) variance in sizes at metamorpho-
sis in larval fishes is lower than variance in sizes at intermediate ages between hatching and meta-
morphosis, thereby implying a ‘target size’ for metamorphosis and (2) the timing of metamorphosis
is more strongly related to the achievement of a target size rather than a target age (Chambers et al.
1988). Therefore, larvae that experience more favourable feeding conditions and grow more quickly
should develop more rapidly and achieve metamorphosis at earlier ages. The stage duration hypoth-
esis predicts that, as a consequence, faster-growing larvae should experience lower cumulative mor-
tality during the larval stage, during which mortality rates are known to be very high. Simulation
studies of the potential impact of the reduced time to metamorphosis resulting from accelerated
growth indicate that resulting survival could be increased up to 100-fold by rapid growth (Chambers
& Leggett 1987, Houde 1989).
However, when Bertram (1993) incorporated the size-/age-specific mortality relationships iden-
tified by Litvak & Leggett (1992) and Pepin et al. (1992) into the stage-specific survival model in a
manner that recognized the combined probabilities of detection, attack and capture that character-
ize vulnerability to predation, he observed that the hypothesized positive effect of fast growth on
cumulative survival inherent in the stage-specific model was negated. Indeed the revised model
demonstrated that, in some cases, survival advantage went to slower-developing individuals. This

341
WILLIAM C. LEGGETT & KENNETH T. FRANK

outcome is consistent with the experimental results reported by Litvak & Leggett (1992) and Pepin
et al. (1992), with the finding by Pepin et al. (2003) that wild, fast-growing larvae studied via high-
resolution in situ sampling experienced higher rates of mortality than did slower-growing larvae and
with the conclusion by Fortier & Quinonez-Velazquez (1998) that rapid growth and development is
a necessary, but not sufficient, condition for improved survival.

‘Growth-selective’ hypothesis
The ‘growth-selective’ hypothesis derives principally from the work of Takasuka et al. (2003), who
found that slower-growing larvae experienced higher rates of predatory losses independent of their
individual size at capture (hence the name ‘growth-selective’ hypothesis). Inherent in this hypoth-
esis is the assumption that the nutritional/physiological state of individual larvae may influence
their probability of capture (perhaps because of a reduced capacity to detect and/or avoid preda-
tory attacks) independent of size (‘bigger is better’) or developmental state (‘stage duration’). Their
in situ analysis was conducted by comparing the size and growth rates of individual larvae recov-
ered from the guts of a suite of predators (as assessed through otolith analysis) with the distribution
of sizes in the larval prey field exploited by these predators. Significantly, they found evidence
of both negative and positive size-selective predation depending on the particular predator spe-
cies investigated. However, the combined effect of all predators was neutral. This finding helps to
explain the conflicting results of individual studies in which one or a small number of predators was
employed or studied.
A follow-up study (Takasuka et al. 2004) provided further evidence of predator-specific growth-
selective mortality. Larval anchovy (Engraulis japonicus) having slower recent growth rates than
those in the population from which they were consumed (independent of size at capture), exhibited
higher predation losses when preyed on by juvenile conspecifics. However, predation by skipjack
tuna (Katsuwonus pelamis) on the same population of anchovy larvae produced no such growth-
selective mortality.
Fuiman et al. (2006) examined the presumed basis for growth-selective predation in a suite of
experiments involving larval red drum (Sciaenops ocellatus). These experiments were conducted
both in the laboratory and in in situ enclosures. They found that populations of larval red drum
consisting of 15 fast- and 15 slow-growing larvae of comparable size, when exposed to predation
and maintained in in situ enclosures, showed no evidence of growth-selective predation. This result
could, however, be a product of the particular predator used and is consistent with the findings of
Takasuka et al. (2004). Laboratory experiments involving larval red drum designed specifically to
evaluate the behavioural dynamics presumed to underlie growth-selective predation revealed no
evidence of a growth rate effect on performance in 11 survival skills. This outcome raises important
questions about the fundamental assumption of the growth-selective hypothesis.
The strong focus of research on the importance of size, growth rate and development rate as
determinants of survival reflected in these three hypotheses, and in the creative field and laboratory
experiments designed to assess them, has meaningfully advanced knowledge of the behaviour, ecol-
ogy and dynamics of the larval stage of fishes. To date, however, it has contributed only modestly
to the ultimate objective of a fuller understanding of recruitment processes in fishes. Like the Hjort
and Cushing hypotheses that preceded them, there may be a danger that this focus, if sustained at
the expense of other research directions could stall advances in the field by precluding a more holis-
tic view of the factors regulating larval survival and recruitment. The hypothesis that larger larvae
should experience a survival advantage when subjected to predation is, for example, counter to the
predator-based hypothesis of optimal foraging that predicts predators will actively select larger prey
items from a prey field consisting of individuals of different sizes (Werner & Hall 1974, Charnov
1976, Litvak & Leggett 1992). Moreover, the underlying premise that the processes inherent in

342
PARADIGMS IN FISHERIES OCEANOGRAPHy

these growth-related hypotheses are vital to recruitment continues to be built, in the main, on a
shaky foundation of correlation studies in which cause and effect are often conflated and/or con-
fused. For example Jenkins & King (2006) conclude, on the basis of correlation studies, that larval
growth in seagrass-associated fishes was important to post-larval abundance and interannual varia-
tion in recruitment. However, larval growth was also strongly correlated to water temperature, and
water temperature was, in turn, strongly correlated with recruitment. An equally credible conclu-
sion might be that temperature had an influence, either directly or indirectly (see Frank & Leggett
1982, 1983), on survival independent of growth rates. In fairness to Jenkins & King (2006) some
consideration was given to other processes, including variable physical transport processes that
affected larvae differentially from year to year (of which variable sea-surface temperature could be
an indicator). However in this example, as in others (see, for example, Meekan et al. 2003, Takasuka
& Aoki 2006), the link between increased growth rates and enhanced survival and recruitment has
become almost axiomatic.

Paradigm 4: Environmentally based recruitment


models, when updated with new data, invariably fail
There is a long history of attempts to explain variation in recruitment based on the relationship
between some direct or indirect measure of year-class strength and environmental variables
(Cushing 1982). The most commonly used environmental variables have been temperature, salin-
ity and wind. These have been incorporated into exploratory correlation analyses that typically
use indices of year-class strength based on landings, catch per unit of effort, fishery-independent
scientific surveys, or model outputs based on reconstructed population sizes such as virtual popula-
tion analysis (VPA), as measures of recruitment. Unfortunately, estimation of year-class strengths
using these approaches is not a trivial problem and the resulting estimates can be misleading. Often
researchers have been forced to use commercial landings data or landings per unit of effort (e.g.,
Sutcliffe et al. 1977), data sources that typically fail to conform to the key assumption of time-
invariant fishing effort. Such data series have been distorted by technological developments that
change the relationship between a unit of effort and the resultant catch. More reliable data, derived
from fishery-independent surveys, are, unfortunately, scarce.
Temperature, because it regulates many physiological processes, has long been considered an
important explanatory variable of recruitment and has recently received growing attention in the
context of global warming (Cardinale & Hjelm 2006). One general pattern that has emerged is
that above-average temperatures are associated with better survival among stocks occupying the
northern limit of their distribution (Ottersen & Stenseth 2001). Salinity has frequently been used
as an indirect measure of nutrient flux, with higher-than-average salinity corresponding to elevated
nutrients. Sutcliffe et al. (1983) found a positive correlation between cod recruitment and salinity
on the Newfoundland/Labrador shelves, which they assumed represented increased vertical mixing
that supported increased plankton production. The physical process by which wind may influence
recruitment is thought to be primarily through distributional effects on the egg and larval stages
(but wind-induced upwelling could also increase vertical mixing and plankton production). For
example, the larval stages of Pacific hake (Merluccius productus) appear to be negatively affected
by along-shelf winds that induce offshore Ekman transport to areas where survival is poor (Bailey
1981). Conversely, Atlantic menhaden (Brevoortia tyrannus) appear to benefit from both alongshore
and longshore wind-induced transport to coastal nursery grounds (Stegmann et al. 1999). While
there is sufficient biological justification to consider these physical variables as important contribu-
tors to recruitment variability it should be noted that they are often selected due to their ease of
measurement and the long record of measurement — decades to centuries in some cases.

343
WILLIAM C. LEGGETT & KENNETH T. FRANK

Unfortunately, this rationale makes it unlikely that such relationships will withstand the test
of time. Indeed, when updated with new data such relationships are prone to failure. In a compre-
hensive reexamination of over 60 published environment-recruitment correlation studies of marine
fish and invertebrates, Myers (1998) tested the performance of each using data accumulated sub-
sequent to the publication of the original relationship. In many cases the analysis applied to these
new data duplicated exactly the original. The overall result was disappointing. Most previously
published relationships failed when updated with new data. Interestingly, those correlations that
remained significant when retested generally occurred in populations close to the limits of the spe-
cies range. In northern waters this included Barents Sea cod, Baltic Sea cod, and Pacific herring
(Clupea pallasii). Recruitment in these species exhibited a consistently positive relationship with
temperature. At the southern limit of the range limits recruitment was typically negatively corre-
lated with temperature.
Insights into the possible causes of such latitudinally linked effects are provided by the work
of Fogarty et al. (2001), who examined the relative variability in recruitment of cod and haddock
stocks distributed throughout the north Atlantic. Recruitment variability was measured as the stan-
dard deviation of the residuals from a Ricker stock and recruitment relationship. Despite their close
taxonomic relationships, overlapping spawning times, and similar egg sizes, haddock stocks gen-
erally exhibited higher recruitment variability than cod in each of nine geographic areas where
they co-occurred. Both stocks also exhibited higher variability at the extremes of their geographic
ranges (Barents Sea in the north and Georges Bank to the south; Figure 5).
The authors explain these species-specific differences on the basis of contrasting life-history
traits, physiological tolerances and population dynamic processes. For example, cod, which spawn
over a more extended time period than haddock, may thereby dampen the risk of extreme survival
outcomes (boom-bust) due to environmental variability. Laboratory studies (Laurence 1977) show
that cod larvae can withstand a much broader range of temperature and salinities than haddock,
implying that equivalent environmental forcing can be expected to have a lesser impact on growth
and survival of larvae of cod relative to haddock. The reduced temporal variability in recruitment
of cod relative to haddock also led Fogarty et al. (2001) to speculate that density-dependent mecha-
nisms, possibly cannibalism, may be a more important process in cod. These species differences

1.2 Cod
Haddock
Recruitment variability

0.8

0.4

0
nd
ea

es

S ea

es
n

n
lan

tia

ia

org
o
sS

tla

cot
Far

Sco
rth

Ice
ent

Sco

Ge
E. S
No

W.
B ar

W.

Figure 5  Variability in the long time series of recruitment data for cod and haddock showing that in every
region where the two species co-occur, haddock exhibit higher interannual variability in recruitment, sugges-
tive of stronger environmental effects on haddock than on cod.

344
PARADIGMS IN FISHERIES OCEANOGRAPHy

80 80

60 60
Biomass, '000 mt

Exploitation, %
40 40

20 20

0 0
1970 1975 1980 1985 1990 1995 2000 2005

Figure 6 Spawning stock biomass (dark line), fishable biomass (dashed line) and exploitation rates (grey line)
for eastern Scotian Shelf haddock.

have important implications for the management and potential recovery timescales of the two spe-
cies. For example, the formation of large year-classes at low levels of spawning stock may be more
common in haddock, thus serving to initiate the process of stock recovery. Evidence in support of
this possibility may be drawn from the study by Platt et al. (2003) of haddock larval survival in rela-
tion to spring bloom timing and the apparent recovery of the eastern Scotian Shelf haddock stock
following its collapse in the early 1990s (Figure 6).
Myers (1998) offers several statistical and analytical suggestions for improving the success of
future attempts to link environmental factors to recruitment variability. These include testing of
general hypotheses through the examination of several populations at once to detect general pat-
terns; employing data-splitting procedures — one portion for the exploratory analysis and the other
for testing; correcting for variation in spawner abundance (trends in recruitment may result from
changes in SSB).
Notwithstanding these inherent difficulties, such correlative approaches remain important
tools for identifying relationships worthy of more in-depth investigation through a combination
of laboratory and in situ studies and simulation modelling. Ultimately, however, the development
of recruitment models with ‘staying power’ will likely depend upon the acquisition of a greater
understanding of the biology of the species and of the behaviour of key elements of the physical
and biological environment occupied in support of the development and rigorous testing of a priori
hypotheses. This hypothesis-testing approach has been more common in freshwater than marine
systems (Magnuson 1988, Frank & Leggett 1994), in part because of the many logistic advantages
of working at smaller scales and in what are often less complex systems.
One of the few examples from the marine environment of a long-lasting environment/recruit-
ment model involves the demonstrated link between the frequency of onshore winds during the
period of residence of the larvae of capelin in the beach gravel where they are spawned and their
survival and recruitment (Leggett et al. 1984). This model rests on the results of an extensive series
of field and laboratory investigations focused on the early life history of the species in eastern
Newfoundland waters. In the case of capelin, onshore winds produce a shoreward movement of
warmer, prey-rich waters that displace the cold, upwelled, predator-laden waters that typically dom-
inate the near shore during the capelin spawning season (Frank & Leggett 1982). A combination of
passive (wave-induced) and active (apparently temperature-induced) emergence from the spawning
gravel follows hatching. These cause the larvae to become entrained in a water mass rich in prey and
relatively poor in predators (Frank & Leggett 1982, 1985, Figure 7), thereby increasing the survival
probabilities of individual larvae (Frank & Leggett 1983). The frequency of these onshore wind
events and of the resulting emergence events is also key because survival time of larvae in the gravel
is limited to 3–4 days post-hatch — a time limit set by the rate of consumption of yolk reserves

345
WILLIAM C. LEGGETT & KENNETH T. FRANK

10 3

Interval between occurrences of onshore wind (days)


9

Year-class strength
6

5 2

0 1
1970 1975 1980 1985 1990
Quiescent Capelin Emerging Capelin
Offshore Wind Onshore Wind

Figure 7  Dynamics of the nearshore environment influencing the larval stages of capelin and the resulting
impact on recruitment. The histogram shows onshore wind interval.

following hatching. In years when the frequency of onshore winds is high, survival is favoured.
When onshore winds are infrequent, survival is depressed.
During the period of larval hatching, the combined effects of water mass replacement creat-
ing favourable conditions for feeding, growth and survival coupled with short post-hatching beach
resident times result in increased larval survival and recruitment. The original model documenting
this relationship was based on capelin year-class data from 1966 to 1987 and from wind field data
that reflect variability in the atmospheric pressure systems that generate onshore winds over large
areas of Newfoundland’s east coast. It explained nearly 60% of the interannual variation in year-
class strength in the eastern Newfoundland capelin stock (Leggett et al. 1984). Higher-than-average
water temperatures during the 6-month period of post-emergent pelagic drift also had a positive
effect on year-class strength. There is no documented biological basis for this added contribution
to explained variance. When retested with an additional 16 yr of data the model was found to be
robust (Carscadden et al. 2000). However, the added contribution to explained variance contrib-
uted by the incorporation of temperature during the post-emergent drift interval into the model no
longer prevailed — another example of a failed correlation based on purely exploratory correlative
assessments. This body of research provides an important counter-example to the paradigm that

346
PARADIGMS IN FISHERIES OCEANOGRAPHy

environmentally based recruitment models invariably fail when updated with new data. It also illus-
trates the importance of a detailed understanding of the causative factors shaping the relationships
to the success of the approach.
Clearly variation in recruitment in marine fishes has an important environmental component.
However, understanding the nature of this component and the development of models that reliably
predict recruitment based on this knowledge remains in its infancy. Adjusting for the spawning
stock component in such models and accounting for other factors such as the size or age compo-
sition of the stock, physiological condition of the spawners, distributional changes, and adverse
effects associated with low population levels, so-called Allee effects (Liermann & Hilborn 2001)
are important next steps, as is the gaining of a deeper understanding of the underlying biological
and ecological drivers of the relationships that allow them to be fine-tuned in time and space.
The work of Ottersen et al. (2006) illustrates how the changing composition of the Barents Sea
cod spawning stock has actually strengthened the effect of the environment on recruitment. They
found that the magnitude of the correlation between cod year-class strength and depth-averaged water
temperatures during the period 1946–1999 increased from being relatively weak to strongly posi-
tive over time. The apparent mechanism behind this evolving relationship was the reduction in the
age and size composition of the stock, a change associated with an intensive, size-selective fishery.
It was speculated that this, in turn, may have resulted in differences in the timing, duration or loca-
tion of spawning, all of which have been demonstrated to be age or size dependent in other species.
This suggests that while a diverse age structure may effectively dampen the impact of environmental
variability on recruitment at both annual and interannual timescales, erosion of this age structure
may cause environmental effects to predominate. Such changes may help to explain, in part, the low
success rate of environmentally based models to reliably predict recruitment through time.

Paradigm 5: Populations cannot irreversibly collapse/


collapsed populations will recover in the absence of fishing
Until recently, thinking in the discipline and resulting management strategies were focused on the
idea that the compensatory dynamics inherent in stock recruitment models that form much of the
foundation for research into the population dynamics of marine fishes (Ricker, 1954, Beverton &
Holt 1957, Myers et al. 1995) would preclude irreversible population collapse, even under extreme
exploitation or highly variable recruitment. Both the Ricker and Beverton & Holt models incorpo-
rate a compensatory response that assumes increasing per capita reproductive success (recruits per
adult) as SSB declines to low levels. This leads naturally to the assumption that when fishing mor-
tality is reduced or eliminated, overfished stocks should recover to a stable equilibrium at relatively
high levels of SSB (Frank & Brickman 2000). This expectation was reinforced by the dramatic
resurgence of heavily exploited stocks during the virtual cessation of fishing in the north Atlantic,
North Sea and Baltic during WWII (Pope & Macer 1996). More recently, support for this paradigm
was provided by the rapid positive response of north-west Atlantic groundfish stocks following the
implementation of extended jurisdiction to 200 miles in Canadian Atlantic waters in 1978 and the
resulting reduction of fishing effort (Figure 8).
However, the lack of generality of this paradigm was brought into sharp focus in the early 1990s
when one of the largest cod stocks in the world collapsed (SSB declined by +99% relative to the
1960s) (Hutchings & Reynolds 2004) and then failed to respond to a moratorium on exploitation
that continues to this day (Figure 8). This phenomenon was repeated over the same time interval for
several other fish stocks within the north-west Atlantic (Choi et al. 2004, 2005).
While controversy regarding the primary cause of the collapse continues (Hutchings 1996, Rose
et al. 2000, Lilly et al. 2001) it is becoming clear that it was not the simple product of overfishing.
The popular analogy of the straw that broke the camel’s back appears to apply in this case — the

347
400
1200 1 2 400 3
300
1000 300

200
800 200

400 100 100

Thousands of metric tonnes


0 0 0
200
120 4 5 6
200
150
80
100
100
40
50

348
Thousands of metric tonnes
0 0 0
30 80 100
7 8 9
60 80
20
60
40
40
10
WILLIAM C. LEGGETT & KENNETH T. FRANK

20
20

Thousands of metric tonnes


0 0 0
1950 1960 1970 1980 1990 2000 1950 1960 1970 1980 1990 2000 1950 1960 1970 1980 1990 2000

Figure 8  Spawning stock biomass (solid line) and landings (dashed line) of cod in nine areas in the north-west Atlantic. 1, northern (Northwest
Atlantic Fisheries Organization Division 2J3KL); 2, northern Gulf of St. Lawrence; 3, southern Gulf of St. Lawrence; 4, St. Pierre Bank; 5,
southern Grand Banks; 6, eastern Scotian Shelf; 7, Gulf of Maine; 8, western Scotian Shelf; 9, Georges Bank.
PARADIGMS IN FISHERIES OCEANOGRAPHy

1800

1600 62

1400 64
63
65
Population biomass ('000 mt)

1200
68
66
Spinal integrity

1000 67 69

800
70
71
600 72
73 81
74 82
400
75 90

200 76 91
76 92
93
0
0.3 0.4 0.5 0.6
Cumulative depletion
(% of biomass removed)

Straw load

Figure 9  The last straw: time course of change in the biomass of northern cod (Northwest Atlantic Fisheries
Organization Division 2J3KL) in relation to cumulative exploitation.

argument is, in reality, about which of the many interacting factors constituted the proverbial straw
(Figure 9).
Clearly, overfishing was an issue. These populations were subjected to many decades of sus-
tained heavy exploitation, commonly two to four times the level of natural mortality, that resulted
in reductions in spawning biomass from 77% to 99% of the historical maximum (Myers et al. 1997).
However this reduction occurred without the collapse of the populations affected even though all
stocks exhibited growth overfishing (exploitation of individuals prior to age of maturity) and sub-
sequent recruitment overfishing (recruitment inadequate to replenish the SSB). Notwithstanding
repeated scientific advice to reduce fishing mortality, exploitation remained high until the collapse
occurred (Lilly et al. 2001). The collapse was immediately followed by an initial 2-yr moratorium
on fishing with the expectation, consistent with the prevailing paradigm, that recovery would occur.
To date no such recovery has occurred notwithstanding repeated extensions of the fishing morato-
rium (Shelton & Healey 1999).
Changes in ocean climate clearly played a role in the collapse. The ocean climate on the
Labrador and Newfoundland shelves is determined by local winter atmospheric conditions that
form part of a larger-scale atmospheric pressure system known as the NAO or North Atlantic
Oscillation (Drinkwater 2002, 2006). The NAO is defined as the winter-time sea-surface pressure
difference between Iceland and the Azores. The relative strength of this difference varies annually.
Positive NAO values (associated with years when the Icelandic low deepens and the Azores high

349
WILLIAM C. LEGGETT & KENNETH T. FRANK

increases) produce both an intensified latitudinal pressure gradient and westerly winds. Negative
NAO values (reflecting a weakening of these pressure systems) produce decreased westerly winds.
The 40+-yr time series of the NAO index exhibits strong decadal variability. Positive NAO years
are characterized by southward penetration of cold Arctic air masses, which produce greater ice
formation along the Newfoundland coast, within the Gulf of St. Lawrence, and along the eastern
Scotian Shelf (Figure 10). These cold, windy conditions also increase convective heat loss, produc-
ing below-normal water temperatures and increases in the volume of the cold intermediate layer
(Figure 10). Positive NAO years dominated the early 1970s and early 1980s and were particularly
pronounced during the late 1980s to mid-1990s, when cod stocks were collapsing throughout the
northern waters of the Canadian Atlantic region (Drinkwater 2002).
These intensified cold periods are believed to have contributed in fundamental ways to changes
in the growth rates, age and size at maturity and even the distribution of many species inhabiting
the cold intermediate layer. Recent examples of the effect on distribution include the southward
extension of Arctic cod from the Labrador Shelf to the southern Grand Banks off Newfoundland
and of capelin from the southern Grand Banks to the Scotian Shelf and beyond (Figure 11).
In general, growth rates, size at maturity and age at maturity in the areas affected by the NAO
declined (Choi et al. 2004, Olsen et al. 2005, Figure 3). A corresponding decrease in individual and
population fecundity (reproductive potential) presumably resulted. This coupled with a significant
reduction in age structure would render the populations more vulnerable to collapse.
In the case of cod the reasons for their failure to recover in the absence of fishing are now com-
ing to light (Frank et al. 2005). This failure is now known to be linked to the dominant position
that cod once occupied as top predator in the benthic food chain of the historical ecosystem and to
fundamental changes in the structure and energy flux in this ecosystem subsequent to their collapse
(Frank et al. 2005). The collapse of cod and the resulting reduction in the predation pressure they
exercised resulted in a massive increase in the biomass of small pelagic fish species and of benthic
macroinvertebrates, species that historically constituted their principal prey (Figure 12). These spe-
cies then became, in turn, important predatory regulators of the early life stages of their former gad-
oid predators (Frank et al. 2005, 2006). This negative feedback is believed to have exacerbated the
recruitment failure being experienced by cod (Swain & Sinclair 2000) and, together with the loss
of reproductive potential related to low water temperatures, reduced growth and reduced individual
fecundity at maturity (Choi et al. 2004), may have constituted the final straw in the demise of cod.
Several studies conducted since the collapse have now demonstrated that extensive delays in,
and/or failure of, the recovery of collapsed stocks are common. Hutchings & Reynolds (2004)
examined the evidence of the ability of 90 marine fish populations, representing 38 species and
11 families, to recover after collapse. They found recovery to be negatively associated with collapse
(r = −0.46, p < .0001). Moreover, 41% of the 90 populations continued to decline 5 yr after collapse.
The magnitude of these population collapses was also negatively associated with recovery 10 and
15 yr after the declines. Only 12% of marine stocks (all of them clupeids) had exhibited full recov-
ery, while 40% (primarily gadoids, but some clupeids) had experienced no or very limited recovery
15 yr after collapse. Moreover, among 36 stocks in which exploitation rate declined after collapse,
and for which estimates of fishing mortality were available, population recovery was still highly
and negatively associated with the magnitude of the population decline (all populations, r = −0.63,
p < .0001; clupeids excluded, r = −0.79, p < .0001) over periods of 5 or 15 yr. Beverton (1990), in
his review of the collapse of 10 major fisheries for small pelagic marine species, notes that only
one stock (Icelandic summer spawning herring (Clupea harengus harengus)) had fully regained its
original size and that the Icelandic spring-spawning herring stock had failed to recover 20 yr since
it collapsed.
A major assumption, inherent in the formulation of the compensatory stock/recruitment mod-
els discussed, is that the curve representing the relationship between stock and recruitment passes

350
PARADIGMS IN FISHERIES OCEANOGRAPHy

1960 1970 1980 1990 2000 2010


–20
NAO winter index –15
–10
–5
0
5
10 2

Air temperature (°C)


1

0
Cartwright
St. John’s –1
–6
–4
Nfld ice extent

–2
105 km2

0
2
4

Area* duration (105 km2 d)


–8
–6
–4

GSL Ice
–2
0
2
4
–20
Area* duration (105 km2 d)

–10

0
SS Ice

10
–20

CIL Area (km2)


–10

10
Sta 27 0-175 Temperature

0.6
0.4
0.2
0.0
–0.2
–0.4
Density 50m-0m (kg m–3)

–0.6 –0.3
–0.2
–0.1
0.0
0.1
0.2
1960 1970 1980 1990 2000 2010

Figure 10  Physical environmental changes from the Newfoundland Shelf and areas to the south illustrating
the backdrop against which many of the stocks were subjected during the past four decades. Note that all axes
except air temperature have been reversed to highlight the coherence of changes in the physical variables in
response to NAO forcing. (GSL = Gulf of St. Lawrence; SS = Scotian Shelf; CIL = Cold Intermediate Layers;
Stn. 27 = 30k east of St. John’s, Newfoundland.)

351
WILLIAM C. LEGGETT & KENNETH T. FRANK

Pre-1985 Post-1985
60 60

55 55

50 50

45 45

40 40
–75 –70 –65 –60 –55 –50 –45 –75 –70 –65 –60 –55 –50 –45

Figure 11  Arctic cod range expansion based on scientific surveys conducted between 1970 and 1994. The
period post-1985 was a time of intense cooling and a resultant southward increase in the distribution of Arctic
cod. Dots show locations of positive catches (i.e., presence).

400
Cod
Shrimp 1
Shrimp 2
Shrimp 3
300 Shrimp 4
Snow crab
Cod SSB ('000 mt)

200

100

0
1970 1975 1980 1985 1990 1995 2000

Figure 12 Changes in cod biomass in the northern Gulf of St. Lawrence and associated changes in their
prey, namely, shrimp from four different areas and snow crab.

through the origin. This assumption underlies the expectation that populations should recover
from the relaxation of fishing, even when SSB has been driven to very low levels. The logical
necessity that the relationship between parent stock size and recruitment pass through the origin
(i.e., without parents there can be no offspring) has probably done more harm than good in terms
of the management of commercial fisheries. This implies that the parent stock size can be driven
to very low levels without any permanent consequences, yet it is well known from other animal
populations that thresholds exist below which population recovery is impossible. These thresholds,

352
PARADIGMS IN FISHERIES OCEANOGRAPHy

Non-zero intercept S-R model

20

Recruitment
10

0
0 20 60 100 140
SSB

Figure 13 Schematic representation of the non-zero intercept stock recruitment model illustrating the basis
for the Allee effect. (SSB = spawning stock biomass.)

termed ‘Allee effects’ (Liermann & Hilborn 2001, Figure 13), result from a variety of behavioural
and physiological changes that occur as population densities decline.
Several recent examples demonstrate the existence of Allee effects in commercially exploited
marine species for which recruitment is inhibited below a parent population size well before the
zero intercept is reached (Shepherd & Partington 1995). In the case of southern Australian abalone
(Haliotis australis) which exhibits this effect, the underlying mechanism for this relatively seden-
tary species is believed to be density-dependent mating success, the probability of finding a mate
being the limiting factor when population size is low. Coho salmon (Oncorhynchus kisutch) also
exhibit an Allee effect (Chen et al. 2002). Another line of evidence in support of the existence of
Allee effects is given by a non-zero intercept model fitted to southern Grand Banks yellow­tail floun-
der (Limanda ferruginea) stock and recruitment data for the years 1968–1982. This model provided
a superior fit to the data compared to that obtained with a standard Ricker model (Figure 14A,B).
For yellowtail flounder a threshold SSB of 18 kilotonnes was indicated. Recruitment failure
would be expected to occur if biomass levels fall below this threshold. In this analysis the residual
pattern diverged between the two model fits, the magnitude of the residual variation being different
in nearly every year and in some years of opposite sign. Such differences have a profound impact on
any conclusions drawn from attempts to explain the residual variation based on the availability of
environmental data, a common practice as detailed above.
Increased awareness of the vulnerability of marine fish populations to irreversible collapses,
triggered by the collapse of north-west Atlantic cod, sparked an awareness of, and interest in, the
implications of the operation of Allee effects in the stock/recruitment dynamics of marine fishes
in general. A 2006 Institute for Scientific Information (ISI) search of all records linking the Allee
effect to fish revealed only 38 records, of which only 6 were published prior to 2000, the earliest
publication date being 1993, 3 yr after the collapse of cod. Because of the difficulty in detecting
these effects with the data currently available (Myers et al. 1995, Shelton & Healey 1999) most
attempts to understand its potential consequences have involved modelling (but see Hutchings 1996,
Marshall et al. 1999).
Frank & Brickman (2000) showed that biological reference points relative to stock recruitment
relationships for marine fishes that are developed using aggregated data from multiple stocks, a
common practice in conventional fisheries management and stock assessments, are likely to be inac-
curate and to obscure Allee effects. This finding and their observation that the extent and timescale
of recovery of collapsed stocks may depend on recolonization from adjacent substocks have, in turn,
fostered interest in the importance of stock structure and metapopulations in population stability,
collapse and recovery.

353
WILLIAM C. LEGGETT & KENNETH T. FRANK

200
A

160
Recruitment (millions)

120

80

40

0
0 20 40 60
SSB ('000 ml)

120 B 0, 0 non-zero
Observed-predicted recruitment

80

40

–40

–80
1968 1970 1972 1974 1976 1978 1980 1982

Figure 14  Fitting of a non-zero intercept model versus the standard Ricker model (A) for yellowtail floun-
der on the southern Grand Banks (Northwest Atlantic Fisheries Organization Division 3NO). The result is a
better fit, estimation of a minimum spawning stock biomass (SSB) level, and a very different residual pattern
compared to the standard model (B).

Gruntfest et al. (1997) defined a metapopulation as a set of local populations connected by


migration in which the crucial feature is a low, often-random and density-dependent rate of move-
ment between patches. Thus the dynamics of local populations are largely, but importantly, not
completely independent. Dispersal between local populations has been shown to reduce the likeli-
hood of an Allee effect (Gascoigne & Lipcius 2004) and even a small amount of interconnectivity
may be sufficient to reduce the risk of extinction/failure to recover for each of the subpopulations
involved (Hill et al. 2002). An important corollary is that habitat fragmentation that isolates local
subpopulations, such as those previously discussed on the various banks of the Scotian Shelf, or dra-
matic reductions in abundance, and thus the interconnectivity of local populations, can create Allee
effects (Cantrell et al. 2001). In metapopulations or in spatial populations consisting of demes, the
high costs of dispersal can induce the Allee effect, particularly when suitable habitats are separated
by uninhabitable areas (Hui & Li 2004).
Zhou & Wang (2004) developed a model that incorporated both a local population and a meta-
population. Their work indicates that a threshold exists below which the metapopulation goes extinct
as a direct consequence of the existence of Allee effects in the corresponding local populations.
They also observed that the threshold fraction of occupied patches needed by the metapopulation

354
PARADIGMS IN FISHERIES OCEANOGRAPHy

to persist increases as the population size on occupied patches declines, and that as demographic
stochasticity in local populations increases in the presence of an Allee effect the risk of extinction
of the metapopulation increases, no matter how many patches are occupied initially.
Brassil (2001) reported that the magnitude of the Allee effect required to reduce the mean time
to extinction of a metapopulation by one half was on the order of 50 times lower than that required
to effect a similar reduction in the mean time to extinction of a single population. Moreover, as
rates of migration between subunits decreased, as might occur as local population density declines
(Frank 1992), the metapopulation became increasingly more sensitive to small changes in the mag-
nitude of the Allee effect.
The overall reduction in the size of local stocks and the progressive loss of spawning compo-
nents from north-west Atlantic cod demonstrates the unplanned, negative consequences of such an
aggregated management scale (Smedbol & Stephenson 2001). The rapid and unexpected collapse
of what might be termed the ‘cod megapopulation’ of the north-west Atlantic is also consistent with
the results and predictions of inquiries into the dynamics of Allee effects at the local and meta­
population levels.

Paradigm 6: Fish stocks can be managed in isolation


from their total environment
Typically fish stock assessments have ignored the complexities detailed above and have focused
principally on the intrinsic aspects of fish population dynamics and fishing as if the stock existed in
a simple vacuum. Such simplifications, once deemed necessary to tackle a very complex problem,
are no longer required given the evolving understanding of the broader ecosystem dynamics within
which individual populations operate.
For example, the consequences of the collapse of gadoid species in the north-west Atlantic have
now been shown to reach far beyond the interplay between the historically dominant predators and
their prey. Frank et al. (2005) documented a trophic cascade resulting from the collapse of cod that
encompassed all trophic levels and nutrients (Figure 15).
The correlation between the groundfish community time series and that of each subsequent
trophic level was negative for small pelagics (−0.61, n = 33), changed to positive for zooplankton
(0.63, n = 23), and then to negative for phytoplankton (−0.73, n = 23). The ecosystem restructuring
associated with this cascade has totally, and perhaps permanently, altered the dynamics of the area
thereby fulfilling the prophecy of Knowlton (2004) that “humans are well on their way to impact-
ing oceans as if they were merely a very large lake, with implications for alternate stable states”.
Traditionally only simple, small-scale, low-diversity ecosystems were considered susceptible to
such profound ecosystem restructuring. The dynamic ecosystem response to the collapse of cod is
the first demonstration of a trophic cascade in a large marine ecosystem.
This ecosystem restructuring to an alternate state may account for the failure of cod populations
to comply with the historical paradigm that predicts the recovery of depressed populations in the
absence of exploitation. The alternate state that now exists resulted from a complex suite of interact-
ing variables and led, in addition to increased predatory mortality on the early life stages of cod and
other demersal species, to a suite of phenotypic changes that included smaller average body sizes,
reduced condition and reduced reproductive output in the gadoid species, all of which reduced the
potential for recovery along lines predicted by stock recruitment theory.
Capelin, too, have recently shown evidence of significant phenotypic changes, related to changes
in the ecosystem they occupy, that have altered the historical relationship between numerical abun-
dance and biomass. Beginning in the early 1990s, coincident with the collapse of cod, size at age
in capelin began to decline. In addition, acoustic surveys designed to assess the abundance of the
species became ineffective, in many cases failing to detect these typically pelagic species. Cod were

355
WILLIAM C. LEGGETT & KENNETH T. FRANK

500 20 80 400
A landings B small fish
biomass snow crab
400 seals shrimp

Snow crab and shrimp


95% C 15 60 300
95% C

Small fish, kt
Groundfish

300

Seals
10 40 200
200

5 20 100
100

0 0 0 0

200 400 2 800


C large D colour nutrients 95% C
small
150 95% C 300 1.5 600

Small zooplankton

Phytoplankton
Zooplankton

Nutrients
100 200 1 400

50 100 0.5 200

0 0 0 0
Pre-collapse Post-collapse

Figure 15  Trophic cascade on the eastern Scotian Shelf. Data the same as in Frank et al. (2005) but expressed
as means during the pre- and post-collapse periods of groundfish on the eastern Scotian Shelf. CI, confidence
interval.

major predators of capelin. Following the collapse of cod, seals and seabirds became the dominant
predator of capelin. This shift in dominance from demersal to surface predators may have caused
capelin to move deeper in the water column (where they are more difficult to detect acoustically),
thereby displacing them from their zooplankton prey and resulting in reduced feeding and growth
rates. There is a growing interest in such non-lethal trophic effects (termed ‘trait-mediated indirect
interactions’) in which the presence of a predator in the environment influences the interaction
between two other species (prey and their resource) by altering a behavioural trait of the prey spe-
cies (Trussell et al. 2003). The slower seasonal growth rates experienced by capelin also led to later
and later spawning times (peaks in August as opposed to historical peaks in June; Carscadden et al.
1997). This shift in the timing of spawning may have negatively affected egg and larval survival
and, in turn, led to reduced recruitment success.

Paradigm 7: Recovery is synonymous with rebuilding


In those cases where recovery from depressed population numbers does occur, it is commonly
believed that the structure of the population, including substock structure, production attributes,
behavioural adaptations, and phenotypic makeup, will resemble the historical condition. Recent
evidence suggests that this may also be incorrect. For example, haddock populations in the north-
west Atlantic collapsed synchronously with cod but, unlike cod, have shown modest signs of recov-
ery during the fishing moratorium. This partial recovery was fuelled by the survivorship of strong
year-classes produced in 1998 and 1999, leading to strong numerical abundances of this species.
However, biomass levels remain well below the historical at these levels of numerical abundance
(Figure 16).

356
PARADIGMS IN FISHERIES OCEANOGRAPHy

1.2
Total mortality

0.8

0.4

0
Weight at age, kg

0
Biomass, '000 mt Weight at 45 cm 50% maturity

40
Length at

30

900
800

60 60

Exploitation, %
40 40

20 20

0
1970 1975 1980 1985 1990 1995 2000 2005

Figure 16  Total mortality, weight at age, length and 50% maturity and index of condition from two surveys
for eastern Scotian Shelf haddock. (Bottom panel is the same as Figure 6.)

This reduced biomass results from major changes in the phenotypic characteristics of the recov-
ered stocks, some of which may have a genetic basis. Haddock now mature at smaller sizes and
younger ages, exhibit lower weights at age (implying reduced growth) and experience increased
natural mortality that is now approximately double (M = 0.2–0.4) the precollapse levels Thus, in the
traditional sense of the word, while these stocks have recovered numerically they have not rebuilt.

Conclusion
This review has attempted to provide the reader with a critical but objective insight into the evolu-
tion and current status of seven paradigms that have had a major influence on the development of
the science of fisheries oceanography over the past century. Many remain central to the field but are,
as should be the case, continuing to evolve as new information becomes available and new insights
are drawn from this information.

357
WILLIAM C. LEGGETT & KENNETH T. FRANK

One of the central themes of the review is that nothing is ever as simple as it seems, and that too
rigid an adherence to a hypothesis that is intuitively appealing can, if not continuously challenged,
dampen the creativity of thought and action that is necessary to advance the discipline. The main-
tenance of an open, critical and creative dialogue, unpopular and uncomfortable as this can be both
for proponents of a theory and for those who challenge, is vital to the advancement of the field.
A second theme is that the dynamics of the systems and phenomena we seek to understand
are invariably more complex than initially meets the eye. It is becoming more and more evident
that even what appear to be simple systems and system responses are shaped in complex ways by
the ecosystem of which they are a part. This requires that we be more holistic in our thinking and
more comprehensive in our studies. Ecosystems rather than individuals or populations are rapidly
becoming the object of study as the search for solutions to traditional problems such as recruitment
dynamics are pursued.
Fortuitously, advances in the technologies and methodologies necessary to approach the disci-
pline from this perspective are occurring at a rapid pace, as is the ability of the current and upcom-
ing generation of physical, chemical and biological oceanographers and marine ecologists to speak
one another’s language with sufficient fluency to create the synergy that is clearly required to be
effective at this level of inquiry. This is, perhaps, one of the greatest advances in the discipline in
recent decades. No aspirant to the profession should ignore this reality.
Finally, this need to think in ecosystem-level terms and to conduct effective research at this
scale bring into focus, once again, the value of large-scale monitoring programmes that provide the
high-quality decadal and longer timescale databases essential to advances in the discipline. Satellite
imagery and in situ remote sensing are now contributing remarkable new insights at the ecosystem
level and will become increasingly important in the future. And the assembly and ease of acces-
sibility of large-scale databases deriving from monitoring and process-oriented studies around the
world are proving to be powerful in advancing the discipline. However, the requirement for ships
at sea in support of skillfully designed sampling programmes executed to provide both decadal
and longer-scale insights into the dynamics of key aspects of large-scale ecosystems and to support
process-based investigations of the operational basis of these dynamics has never been greater. One
need only look to the contributions of initiatives such as the Continuous Plankton Recorder pro-
gramme administered by the Sir Alister Hardy Foundation for the Ocean Sciences, the California
Cooperative Oceanic Fisheries Investigation, annual bottom trawl surveys such those administered
by the Department of Fisheries and Oceans in Canada, the National Marine Fisheries Service in the
United States, and by International Council for the Exploration of the Sea (ICES) member nations in
Europe to recent advances in the discipline to understand this reality. This may not be the glamor-
ous science that garners competitive grant monies, but without it the contribution of those monies to
the understanding and solution of problems that have taken on enormous worldwide importance in
this time of global climate change will be increasingly at risk.

Acknowledgements
This work was supported by Discovery grants from the Natural Sciences and Engineering Research
Council of Canada to W.C. Leggett and to K.T. Frank and by Fisheries and Oceans Canada, Ocean
Sciences Division. We thank Liam Petrie for his assistance in producing the figures.

References
Anderson, J.T. 1988. A review of size dependent survival during pre-recruit stages of fishes in relation to
recruitment. Journal of the Northwest Atlantic Fisheries Organization 8, 55–66.
Bailey, K.M. 1981. Larval transport and recruitment of Pacific hake, Merluccius productus. Marine Ecology
Progress Series 6, 1–9.

358
PARADIGMS IN FISHERIES OCEANOGRAPHy

Bertram, D.F. 1993. Growth, development and mortality in metazoan early life histories with particular refer-
ence to marine flatfish. Ph.D. thesis, McGill University, Toronto, Canada.
Bertram, D.F. & Leggett, W.C.1994. Predation risk during the early life history periods of fishes: separating the
effects of size and age. Marine Ecology Progress Series 109, 105–114.
Beverton, R.J H. 1990. Small marine pelagic fish and the threat of fishing; are they endangered? Journal of Fish
Biology 37(Supplement A), 5–16.
Beverton, R.J.H. & Holt, S.J. 1957. On the Dynamics of Exploited Fish Populations, Series II, Volume 19.
London, Great Britain: Fish Investigations Ministry of Agriculture Fishes and Food.
Berkeley, S.A., Chapman, C. & Sogard, S.M. 2004. Maternal age as a determinant of larval growth and survival
in a marine fish, Sebastes melanops. Ecology 85, 1258–1264.
Boyd, A.J., Shannon, L.J., Schulien, F.N. & Taunton-Clark, J. 1998. Food, transport, and anchovy recruit-
ment in the southern Benguela upwelling system of South Africa. In Global Versus Local Changes in
Upwelling Systems, M.H. Durand et al. (eds). Paris: Editions de l’Orstom, 195–209.
Brassil, C.E. 2001. Mean time to extinction of a metapopulation with an Allee effect. Ecological Modelling
143, 9–16.
Brickman, D. 2003. Controls on the distribution of Browns Bank juvenile haddock. Marine Ecology Progress
Series 263, 235–246.
Cantrell, R.S., Cosner, C. & Fagan, W.F. 2001. How predator incursions affect critical patch size: the role of the
functional response. The American Naturalist 158, 368–375.
Cardinale, M. & Hjelm, J. 2006. Marine fish recruitment variability and climate indices. Marine Ecology
Progress Series 309, 307–309.
Carruthers, E.H., Neilson, J.D., Waters, C. & Perley, P. 2005. Long-term changes in the feeding of Pollachius
virens on the Scotian Shelf: responses to a dynamic ecosystem. Journal of Fish Biology 66, 327–347.
Carruthers, J.N., Lawford, A.L., Veley, V.C.F. & Parrish, B.B. 1951. Variations in brood strength in the North
Sea haddock, in the light of relevant wind conditions. Nature 168, 317–319.
Carscadden, J.E., Frank, K.T. & Leggett, W.C. 2000. Evaluation of an environment-recruitment model for
capelin (Mallotus villosus). ICES Journal of Marine Science 57, 412–418.
Carscadden, J.E., Nakashima, B. & Frank, K.T. 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.
Chambers, R.C. & Leggett, W.C. 1987. Size and age at metamorphosis in marine fishes: an analysis of labora-
tory reared winter flounder (Pseudopleuronectes americanus) with a review of variation in other species.
Canadian Journal of Fisheries and Aquatic Sciences 44, 1936–1947.
Chambers, R.C., Leggett, W.C. & Brown, J.A. 1988. Variation in and among early life history traits of labo-
ratory-reared winter flounder (Pseudopleuronectes americanus). Marine Ecology Progress Series 47,
1–15.
Charnov, E.L. 1976. Optimal foraging and the marginal value theorem. Theoretical Population Biology 9,
129–136.
Chen, D.G., Irvine, J.R. & Cass, A.J. 2002. Incorporating Allee effects in fish stock-recruitment models and
applications for determining reference points. Canadian Journal of Fisheries and Aquatic Sciences 59,
242–249.
Choi, J.S., Frank, K.T., Leggett, W.C. & Drinkwater, K. 2004. Transition to an alternate state in a continental
shelf ecosystem. Canadian Journal of Fisheries and Aquatic Sciences 61, 505–510.
Choi, J.S., Frank, K.T., Petrie, B. & Leggett, W.C. 2005. Integrated assessment of a large marine ecosystem:
a case study of the devolution of the Eastern Scotian Shelf, Canada. Oceanography and Marine Biology
An Annual Review 43, 47–67.
Cowen, R.K., Paris, C.B. & Srinivasan, A. 2006. Scaling of connectivity in marine populations. Science 311,
522–527.
Cury, P. & Roy, C.1989. Optimal environmental window and pelagic fish recruitment success in upwelling
areas. Canadian Journal of Fisheries and Aquatic Sciences 46, 670–680.
Cushing, D.H. 1975. Marine Ecology and Fisheries. Cambridge, U.K.: Cambridge University Press.
Cushing, D.H. 1982. Climate and Fisheries. London: Academic Press.
Cushing, D.H. 1990. Plankton production and year-class strength in fish populations — an update of the match
mismatch hypothesis. Advances in Marine Biology 26, 249–293.

359
WILLIAM C. LEGGETT & KENNETH T. FRANK

de Figueiredo, G.H., Nash, R.D.M. & Montagnes, D.J.S. 2005. The role of the generally unrecognised micro-
prey source as food for larval fish in the Irish Sea. Marine Biology 148, 395–404.
Dower, J.F., Leggett, W.C. & Frank, K.T. 2000. Improving fisheries oceanography in the future. In Fisheries
Oceanography: An Integrative Approach to Fisheries Ecology and Management, P.J. Harrison & T.R.
Parsons (eds). Cambridge, U.K.: Blackwell Science, 263–281.
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.
Drinkwater, K.F. 2006. The regime shift of the 1920s and 1930s in the North Atlantic. Progress in Oceanography
68, 134–151.
Drinkwater, K.F., Petrie, B. & Smith, P.C. 2003. Hydrographic variability on the Scotian Shelf during the
1990s. ICES Marine Science Symposium 219, 40–49.
Fogarty, M.J., Myers, R.A. & Bowen, K.G. 2001. Recruitment of cod and haddock in the North Atlantic: a
comparative analysis. ICES Journal of Marine Science 58, 952–961.
Fortier, L. & Quinonez-Velazquez, C. 1998.Dependence of survival on growth in larval pollock Pollachius
virens and haddock Melanogrammus aeglefinus: a field study based on individual hatchdates. Marine
Ecology Progress Series 174, 1–12.
Frank, K.T. 1992. Demographic consequences of age-specific dispersal in marine fish populations. Canadian
Journal of Fisheries and Aquatic Sciences 49, 2222–2231.
Frank K.T. & Brickman, D. 2000. Allee effects and compensatory population dynamics within a stock com-
plex. Canadian Journal of Fisheries and Aquatic Sciences 57, 513–517.
Frank, K.T. & Leggett, W.C. 1982. Coastal water mass replacement: its effect on zooplankton dynamics and
the predator-prey complex associated with larval capelin. Canadian Journal of Fisheries and Aquatic
Sciences 39, 991–1003.
Frank, K.T. & Leggett, W.C. 1983. Multi-species larval fish associations: accident or adaptation? Canadian
Journal of Fisheries and Aquatic Sciences 40, 754–762.
Frank, K.T. & Leggett, W.C. 1985. Reciprocal oscillations in densities of larval fish and potential preda-
tors: a reflection of present or past predation? Canadian Journal of Fisheries and Aquatic Sciences 42,
1844–1849.
Frank, K.T. & Leggett, W.C. 1994. Fisheries ecology in the context of ecological and evolutionary theory.
Annual Review of Ecology and Systematics 25, 401–422.
Frank, K.T., Petrie, B., Choi, J.S. & Leggett, W.C. 2005. Trophic cascades in a formerly cod-dominated eco-
system. Science 308, 1621–1623.
Frank, K.T., Petrie, B., Shackell, N.L. & Choi, J.S. 2006. Reconciling differences in trophic control in mid-
latitude continental shelf ecosystems. Ecology Letters 9, 1096–1105.
Frank, K.T., Shackell, N.L. & Simon, J.E. 2000. An evaluation of the Emerald/Western Bank juvenile haddock
closed area. ICES Journal of Marine Science 57, 1023–1034.
Fuiman L.A., Cowan, J.H., Jr., Smith, M.E. & O’Neal, J.P. 2006. Behavior and recruitment success in fish lar-
vae: variation with growth rate and the batch effect. Canadian Journal of Fisheries and Aquatic Sciences
62, 1337–1349.
Gascoigne, J. & Lipcius, R.N. 2004. Allee effects in marine systems. Marine Ecology Progress Series 269,
49–59.
Gruntfest, Y., Arditi, N. & Dombronsky, Y.1997. A fragmented population in a varying environment. Journal of
Theoretical Biology 185, 539–547.
Hanson, J.M. & Chouinard, G.A. 2002. Diet of Atlantic cod in the southern Gulf of St. Lawrence as an index
of ecosystem change, 1959–2000. Journal of Fish Biology 60, 902–922.
Hare, J.A. & Cowen, R.K. 1997. Size growth, development and survival of the planktonic larvae of Pomatomus
saltatrix (Pisces: Pomatomidie). Ecology 78, 2415–2431.
Hilborn, R. & Walters, C.J. 1992. Quantitative Fisheries Stock Assessment: Choice, Dynamics and Uncertainty.
New York: Chapman & Hall..
Hill, M.F., Hastings, A. & Botsford, L.W. 2002. The effects of small dispersal rates on extinction times in struc-
tured metapopulation models. The American Naturalist 160, 389–402.
Hinrichsen, H.–H., St. John, M., Aro, E., Gronkjaer, P. &. Voss, R. 2001. Testing the larval drift hypothesis in
the Baltic Sea: retention versus dispersion caused by wind-driven circulation. ICES Journal of Marine
Science 58, 973–984.

360
PARADIGMS IN FISHERIES OCEANOGRAPHy

Houde, E.D. 1987. Fish early life dynamics and recruitment variability. American Fisheries Society Symposium
2, 17–29.
Houde, E.D. 1989. Subtleties and episodes in the early life of fishes. Journal of Fish Biology 35(Supplement
A), 29–38.
Hjort, J. 1914. Fluctuations in the great fisheries of northern Europe viewed in the light of biological research.
Rapport Process-Verbaux Reunions Conseil International pour l’Exploration de la Mer 20, 1–228.
Hjort, J. 1928. Fluctuations in the year classes of important food fishes. Journal du Conseil, Conseil Permenant
International pour l’Exploration de la Mer 1, 5–38.
Hui, C. & Li, Z. 2004. Distribution patterns of a metapopulation determined by Allee effects. Population
Ecology 46, 55–63.
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. & Reynolds, J.D. 2004. Marine fish population collapses: Consequences for recovery and
extinction risk. Bioscience 54, 297–309.
Jenkins, G.P. & King, D. 2006. Variation in larval growth can predict the recruitment of a temperate, seagrass-
associated fish. Oecologia 147, 641–649.
Knowlton, N. 2004. Multiple “stable” states and the conservation of marine ecosystems. Progress in Oceanog­
raphy 60, 387–396.
Laurence, G.C. 1977. Comparative growth, respiration, and delayed feeding ability of larval cod and haddock
as influenced by temperature during laboratory studies. ICES CM 1977/L, 31. Copenhagen: International
Council for the Exploration of the Sea.
Leggett, W.C. & DeBlois, E. 1994. Recruitment in marine fishes: is it regulated by starvation and predation in
the egg and larval stages? Netherlands Journal of Sea Research 32, 119–134.
Leggett, W.C., Frank, K.T. & Carscadden, J.E. 1984. Meteorological and hydrographic regulation of year-
class strength in capelin (Mallotus villosus). Canadian Journal of Fisheries and Aquatic Sciences 41,
1193–1201.
Lett, C., Penven, P., Ayón, P. & Fréon, P. 2007. Enrichment, concentration and retention processes in relation to
anchovy (Engraulis ringens) eggs and larvae distributions in the northern Humboldt upwelling ecosys-
tem. Journal of Marine Systems 64, 189–200.
Liermann, M. & Hilborn, R. 2001. Depensation: evidence, models and implications. Fish and Fisheries 2,
33–58.
Lilly, G.R., Shelton, P.A., Brattey, J., Cadigan, N.G., Healey, B.P., Murphy E.F. & Stansbury, D.E. 2001. An
assessment of the cod stock in NAFO Divisions 2J+3KL. NAFO SCR Document 01/33. Dartmouth N.S.
Canada: North Atlantic Fisheries Organization.
Litvak M.K. & Leggett, W.C. .1992 Age and size-selective predation on larval fishes: the bigger is better
hypothesis revisited. Marine Ecology Progress Series 81, 13–24.
Lowerre-Barbieri, S.K., Lowerre, J.M. & Barbieri, L.R. 1998. Multiple spawning and the dynamics of fish
populations: inferences from an individual-based simulation model. Canadian Journal of Fisheries and
Aquatic Sciences 55, 2244–2254.
Magnuson, J.J. 1988. Two worlds for fish recruitment: lakes and oceans. American Fisheries Society Symposium
5, 1–6.
Marshall, C.T. & Frank, K.T. 1999. The effect of interannual variation in growth and condition on haddock
recruitment. Canadian Journal of Fisheries and Aquatic Sciences 56, 347–355.
Marshall C.T., Kjesbu, O.S. & Yaragina, N.A. 1998. Is spawner biomass a sensitive measure of the reproductive
and recruitment potential of Northeast Arctic cod? Canadian Journal of Fisheries and Aquatic Sciences
55, 1766–1783.
Marshall, C.T., Yaragina, N.A. & Adlandsvik, B. 2000. Reconstructing the stock-recruit relationship for
Northeast Arctic cod using a bioenergetic index of reproductive potential. Canadian Journal of Fisheries
and Aquatic Sciences 57, 2433–2442.
Marshall, C.T., Yaragina, N.A., Lambert, Y. & Kjesbu, O.S. 1999. Total lipid energy as a proxy for total egg
production by fish stocks. Nature 402, 288–290.
Marteinsdottir, G. & Steinarsson, A. 1998. Maternal influence on the size and viability of Iceland cod (Gadus
morhua) eggs and larvae. Journal of Fish Biology 52, 1241–1258.

361
WILLIAM C. LEGGETT & KENNETH T. FRANK

Meekan, M.G., Carleton, J.H., McKinnon, A.D., Flynn, K. & Furnas, M. 2003. What determines the growth
of tropical reef fish larvae in the plankton: food or temperature? Marine Ecology Progress Series 256,
193–204.
Mousseau, L., Fortier, L. & Legendre, L.1998. Annual production of fish larvae and their prey in relation to
size-fractionated primary production (Scotian Shelf, NW Atlantic). ICES Journal of Marine Science 55,
44–57.
Mueter, F.J., Ladd, C., Palmer, M.C. & Norcross, B.L. 2006. Bottom-up and top-down controls of walleye
pollock (Theragra chalcogramma) on the eastern Bering Sea shelf. Progress in Oceanography 68,
152–183.
Myers, R.A. 1998. When do environment-recruitment correlations work? Reviews in Fish Biology and Fisheries
8, 285–305.
Myers, R.A., Barrowman, N.J. & Hutchings, J.A. 1995. Population dynamics of exploited fish stocks at low
population levels. Science 269, 1106–1108.
Myers, R.A. & Drinkwater, K.F. 1989. The influence of Gulf Stream warm core rings on recruitment of fish in
the northwest Atlantic. Journal of Marine Research 47, 635–656.
Myers, R.A., Hutchings, J.A. & Barrowman, N.J. 1997. Why do fish stocks collapse? The example of cod in
Atlantic Canada. Ecological Applications 7, 91–106.
North, E.W. & Houde, E.D. 2001. Retention of white perch and striped bass larvae: biological-physical interac-
tions in Chesapeake Bay estuarine turbidity maximum. Estuaries 24, 756–769.
Olsen, E.M., Lilly, G.R., Heino, K., Morgan, M.J., Brattey, J. & Dieckmann, U. 2005. Assessing changes in age
and size at maturation in collapsing populations of Atlantic cod (Gadus morhua). Canadian Journal of
Fisheries and Aquatic Sciences 62, 811–823.
Ottersen, G., Hjermann, D.O. & Stenseth, N.C. 2006. Changes in spawning stock structure strengthen the link
between climate and recruitment in a heavily fished cod (Gadus morhua) stock. Fisheries Oceanography
15, 230–243.
Ottersen, G. & Stenseth, N.C. 2001. Atlantic climate governs oceanographic and ecological variability in the
Barents Sea. Limnology and Oceanography 46, 1774–1780.
Pepin P., Dower, J.F. & Davidson, F.J.M. 2003. A spatially explicit study of prey-predator interactions in larval
fish: assessing the influence of food and predator abundance on larval growth and survival. Fisheries
Oceanography 12, 19–33.
Pepin, P. & Miller, T.J. 1993. Potential use and abuse of general empirical models of early life history processes
in fish. Canadian Journal of Fisheries and Aquatic Sciences 50, 1334–1345.
Pepin, P., Shears, T.H. & Delafontaine, Y. 1992. Significance of body size to the interaction between a larval
fish (Mallotus villosus) and a vertebrate predator (Gasterosteus aculeatus). Marine Ecology Progress
Series 81, 1–12.
Pinhorn, A.T. 1984. Temporal and spatial variation in fecundity of Atlantic cod (Gadus morhua) in Newfoundland
waters. Journal of the Northwest Atlantic Fisheries Organization 5, 161–170.
Platt, T., Fuentes-Yaco, C. & Frank, K.T. 2003. Spring algal bloom and larval fish survival. Nature 423,
398–399.
Pope, J.G. & Macer, C.T. 1996. An evaluation of the stock structure of North Sea cod, haddock, and whiting
since 1920, together with a consideration of the impacts of fisheries and predation effects on their bio-
mass and recruitment. ICES Journal of Marine Science 53, 1157–1169.
Ricker, W.E. 1954. Stock and recruitment. Canadian Journal of Fisheries and Aquatic Sciences 11, 559–621.
Rijnsdorp, A.D., Daan, N., van Beek, F.A. & Heesen, H.J.L. 1991. Reproductive variability in North Sea plaice,
sole and cod. Journal du Conseil. Conseil International pour l’Exploration de la Mer 47, 352–375.
Rose, G.A., deYoung, B., Kulka, D.W., Goddard, S.V. & Fletcher G.L. 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.
Rothschild, B.J. 1986. Dynamics of Marine Fish Populations. Cambridge, Massachusetts: Harvard University
Press.
Rothschild, B.J. & DiNardo, G.T. 1987. Comparison of recruitment variability and life history data among
marine and anadromous fishes. American Fisheries Society Symposium 1, 531–546.

362
PARADIGMS IN FISHERIES OCEANOGRAPHy

Scott, B., Marteinsdottir, G. & Wright, P. 1999. Potential effects of maternal factors on spawning stock-recruit-
ment relationships under varying fishing pressure. Canadian Journal of Fisheries and Aquatic Sciences
56, 1882–1890.
Scott, B.E., Marteinsdottir, G. & Begg, G.A. 2006. Effects of population size/age structure, condition and
temporal dynamics of spawning on reproductive output in Atlantic cod (Gadus morhua). Ecological
Modelling 191, 383–415.
Shelton, P.A. & Healey, B.P. 1999. Should depensation be dismissed as a possible explanation for the lack
of recovery of the northern cod (Gadus morhua) stock? Canadian Journal of Fisheries and Aquatic
Sciences 56, 1521–1524.
Shepherd, J.G., Pope, J.G. & Cousens, R.D. 1984. Variations in fish stocks and hypotheses concerning their
links with climate. Rapports et Proces-verbaux Reunions, Conseil International pour l’Exploration de la
Mer 185, 255–267.
Shepherd, S.A. & Partington, D. 1995. Studies on southern Australian abalone (Genus Haliotis). Recruitment,
habitat and stock relations. Marine and Freshwater Research 46, 669–680.
Sinclair, M. 1988. Marine Populations: An Essay on Population Regulation and Speciation. Seattle: University
of Washington Press.
Smedbol, R.K. & Stephenson, R. 2002. The importance of managing within-species diversity in cod and her-
ring fisheries of the north-western Atlantic. Journal of Fish Biology 59(Supplement A), 109–128.
Stegmann, P.M, Quinlan, J.A., Werner, F.E., Blanton, B.O. & Berrien, P. 1999. Atlantic menhaden recruitment
to a southern estuary: defining potential spawning regions. Fisheries Oceanography 8(Supplement 2),
111–123.
Sutcliffe, W.H., Drinkwater, K.F. & Muir, B.S. 1977. Correlations of fish catch and environmental factors in the
Gulf of Maine. Canadian Journal of Fisheries and Aquatic Sciences 34, 19–30.
Sutcliffe, W.H., Loucks, R.H., Drinkwater, K.F. & Coote, A.R. 1983. Nutrient flux onto the Labrador shelf from
Hudson Strait and its biological consequences. Canadian Journal of Fisheries and Aquatic Sciences 40,
1692–1701.
Swain, D.P. & Sinclair, A.F. 2000. Pelagic fishes and the cod recruitment dilemma in the Northwest Atlantic.
Canadian Journal of Fisheries and Aquatic Sciences 57, 1321–1325.
Takasuka, A. & Aoki, I. 2006. Environmental determinants of growth rates for larval Japanese anchovy
Engraulis japonicus in different waters. Fisheries Oceanography 15, 139–149.
Takasuka, A., Aoki, I. & Mitani, I. 2003. Evidence of growth-selective predation on larval Japanese anchovy
Engraulis japonicus in Sagami Bay. Marine Ecology Progress Series 252, 223–235.
Takasuka, A., Oozeki, Y., Kimura, R., Kubota, H. & Aoki, I. 2004. Growth-selective predation hypothesis revis-
ited for larval anchovy in offshore waters: cannibalism by juveniles versus predation by skipjack tunas.
Marine Ecology Progress Series 278, 297–302.
Trussell, G.C., Ewanchuk, P.J. & Bertness, M.D. 2003. Trait-mediated effects in rocky intertidal food chains:
predator risk cues alter prey feeding rates. Ecology 84, 629–640.
Ulltang, O. 1996. Stock assessment and biological knowledge: can prediction uncertainty be reduced? ICES
Journal of Marine Science 53, 659–675.
Vallin, L. & Nissling, A. 2000. Maternal effects on egg size and egg buoyancy of Baltic cod, Gadus morhua —
implications for stock structure effects on recruitment. Fisheries Research 49, 21–37.
Waiwood, K.G. & Buzeta, M.-I. 1989. Reproductive biology of southwestern Scotian Shelf haddock
(Melanogrammus aeglefinus). Canadian Journal of Fisheries and Aquatic Sciences 46 (Supplement 1),
153–170.
Werner, E. & Hall, D.J. 1974. Optimal foraging and size selection of prey by bluegill sunfish (Lepomis macro-
chirus). Ecology 55, 1042–1052.
Wooster, W.S. & Bailey, K.M. 1989. Recruitment of marine fishes revisited. In Effects of Ocean Variability on
Recruitment and an Evaluation of Parameters Used in Stock Assessment Models, R.J. Beamish & G.A.
McFarlane (eds). Canadian Special Publication of Fisheries and Aquatic Sciences, No. 108. Canada
Department of Fisheries and Oceans, Ottawa, Ontario, Canada.
Yaragina, N.A. & Marshall, C.T. 2000. Trophic influences on interannual and seasonal variation in the liver
condition index of Northeast Arctic cod (Gadus morhua). ICES Journal of Marine Science 57, 42–55.
Zhou, S.R. & Wang, G. 2004. Allee-like effects in metapopulation dynamics. Mathematical Biosciences 189,
103–113.

363
herring 4VWX-SSB
yellowtail flounder 5Z-SSB
turbot 4RST-SSB
little skate 4VWX-SSB

View publication stats


winter skate 4X-SSB
smooth skate 4VWX-SSB
70
smooth skate 4T-SSB cod
cod 4T-SSB
haddock
cod 4X-SSB
pollock
silver hake 4VWX-SSB 60
cod 5Z-SSB silver hake
winter skate 4VsW-SSB
haddock 4VW-SSB
cod 4Vn-SSB 50
mackerel-SSB
winter skate 4T-SSB
haddock 4X-SSB

Length at age 5, cm
plaice 3Ps-SSB 40

haddock 5Z-SSB
thorny skate 4T-SSB
thorny skate 4VWX-SSB
30
cod 4VsW-SSB
pollack 4VWX-SSB
white hake 4T-SSB
cod 2J3KL-Offhore Survey-SSB 20
cod 3NO spring-SSB
1970 1975 1980 1985 1990 1995 2000
plaice 2J3K-SSB
plaice 4VT-SSB Colour Figure 3 (Leggett & Frank)  Changes with time in lengths at age 5
1970 1975 1980 1985 1990 1995 2000 for four groundfish species from the eastern Scotian Shelf.
Years

Colour Figure 1 (Leggett & Frank)  Ordination of the time of series of spawning stock biomass
of various species from scientific surveys conducted throughout the north-west Atlantic, illustrating
that the majority of the stocks are at biomass levels well below (red) the long-term average. Intensity
of colours is proportional to the magnitude of the standardized anomaly in standard deviation units.
Alphanumeric labelling refers to the Northwest Atlantic Fisheries Organization management unit.

You might also like