Top Predators in Marine Ecosystems PDF
Top Predators in Marine Ecosystems PDF
The sustainable exploitation of the marine environment depends upon our capacity to
develop systems of management with predictable outcomes. Unfortunately, marine
ecosystems are highly dynamic and this property could conflict with the objective of
sustainable exploitation. This book investigates the theory that the population and
behavioural dynamics of predators at the upper end of marine food chains can be used
to assist with management. Since these species integrate the dynamics of marine
ecosystems across a wide range of spatial and temporal scales, they offer new sources
of information that can be formally used in setting management objectives. This book
examines the current advances in the understanding of the ecology of marine
predators and will investigate how information from these species could be used in
management.
Edited by
i . l . b o y d , s . w a n l e s s a n d c . j . ca m p h u y s e n
Cambridge, New York, Melbourne, Madrid, Cape Town, Singapore, São Paulo
Cambridge University Press has no responsibility for the persistence or accuracy of s
for external or third-party internet websites referred to in this publication, and does not
guarantee that any content on such websites is, or will remain, accurate or appropriate.
Contents
1 Introduction [1]
i . l . b o y d , s . w a n l e s s a n d c . j . ca m p h u y s e n
7 Spatial and temporal variation in the diets of polar bears across the
Canadian Arctic: indicators of changes in prey populations and
environment [98]
s. j. iverson, i. stirling and s. l. c. lang
14 How many fish should we leave in the sea for seabirds and marine
mammals? [211]
r . w. f u r n e s s
19 The method of multiple hypotheses and the decline of Steller sea lions
in western Alaska [275]
n. wolf, j. melbourne and m. mangel
Index [370]
Contributors
C. Asseburg C. J. Camphuysen
Centre for Conservation Science Royal Netherlands Institute for Sea
University of St Andrews Research
St Andrews PO Box 59
Fife KY16 9LZ, UK 1790 AB Den Burg, Texel, the Netherlands
D. Austin V. Christensen
Dalhousie University Fisheries Centre
Halifax, Nova Scotia University of British Columbia
Canada B3H 4J1 Vancouver, British Columbia
Canada V6T 1Z4
C. A. Beck
A. J. Constable
Alaska Department of Fish and Game
Australian Antarctic Division
Division of Wildlife Conservation
Australian Department of Environment and
Marine Mammals Section
Heritage
Anchorage
203 Channel Highway, Kingston
Alaska 99518
Tasmania 7050, Australia
W. D. Bowen J. P. Croxall
Marine Fish Division British Antarctic Survey, Natural
Bedford Institute of Oceanography Environment Research Council
Department of Fisheries and Oceans High Cross, Madingley Road
Dartmouth, Nova Scotia Cambridge CB3 0ET, UK
Canada B2Y 1A
F. Daunt
Dalhousie University NERC Centre for Ecology and Hydrology
Halifax, Nova Scotia Banchory Research Station
Canada B3H 4J1 Hill of Brathens
I. L. Boyd Banchory AB31 4BW, UK
Sea Mammal Research Unit G. K. Davoren
Gatty Marine Laboratory Zoology Department
University of St Andrews University of Manitoba, Winnipeg
St Andrews Manitoba
Fife KY16 8LB, UK Canada R3T 2N2
S. Benvenuti M. R. Enstipp
Dipartimento di Etologia Centre d’Ecologie et Physiologie
Ecologia ed Evoluzione Energétiques
Università di Pisa, Via Volta 6 Centre National de la Recherche
I-56126 Pisa, Italy Scientifique
Contributors ix
St Andrews A. W. Trites
Fife KY16 8LB, UK Fisheries Centre
University of British Columbia
I. J. Staniland
Vancouver, British Columbia
British Antarctic Survey, Natural
Canada V6T 1Z4
Environment Research Council
High Cross, Madingley Road S. Wanless
Cambridge CB3 0ET, UK NERC Centre for Ecology and Hydrology
I. Stirling Banchory Research Station
Canadian Wildlife Service Hill of Brathens
Edmonton, Alberta Banchory AB31 4BW, UK
Canada T6H 3S5
N. Wolf
P. M. Thompson MRAG Americas
Lighthouse Field Station 110 South Hoover Boulevard, Suite 212
School of Biological Sciences Tampa, Florida 33609, USA
University of Aberdeen Center for Stock Assessment Research
Cromarty IV11 8YJ, UK University of California Santa Cruz
S. E. Thorpe 1156 High Street
British Antarctic Survey, Natural California 95064, USA
Environment Research Council
J. M. Yearsley
High Cross, Madingley Road
Culterty Field Station
Cambridge CB3 0ET, UK
The School of Biological Sciences
P. N. Trathan The College of Life Sciences and
British Antarctic Survey, Natural Medicine
Environment Research Council University of Aberdeen
High Cross, Madingley Road Newburgh
Cambridge CB3 0ET, UK Ellon AB41 6AA, UK
Preface
This book began its evolution in 1999 when the British Antarctic Survey,
where I worked at the time, began a new research programme on the man-
agement of marine ecosystems. This programme concentrated upon the
krill-based ecosystem at South Georgia which has been the subject of almost
continuous study since the Discovery Expeditions in the 1920s. Latterly,
international efforts to understand the dynamics of this ecosystem and the
wider Southern Ocean have been coordinated by the Commission for the
Conservation of Antarctic Marine Living Resources (CCAMLR). The daunt-
ing task of describing ecosystem dynamics over such a large oceanic area
with relatively limited resources led to the establishment of the CCAMLR
Ecosystem Monitoring Programme, an internationally coordinated effort at
data collection. Among other things, this contained a major component of
monitoring the seal and seabird populations in the region. The logic for
their inclusion was that they foraged over most of the regions of interest
but returned to breed at very well defined locations. By undertaking a series
of measurements of these predators at these locations, it was then argued
that aspects of the ecosystem dynamics should be reflected by variability in
the measurements of the predators. It was hoped that appropriate choices
of the predators and measurement variables would provide indicators of the
dynamics of their prey at different spatial and temporal scales.
The same concept has been developed in parallel within other ecosys-
tems during the past 20 years. The North Sea, California Current, northwest
Atlantic, Bering Sea, Gulf of Alaska and Barents Sea are regions in which
long-term monitoring studies of seabirds and seals are recognized as pro-
viding insights into ecosystem processes that can then be fed into the pro-
cess of management. Even though the implementation and use of mea-
surements has differed between regions, there has been a strong recogni-
tion that the interpretation of data about predator dynamics in the context
of ecosystem dynamics can only be achieved on the back of basic research
into the ecology of the species concerned. This book is, therefore, an effort
xiv Preface
I. L. Boyd
1
Introduction
I. L. BOYD, S. WANLESS AND C. J. CAMPHUYSEN
& Butterworth 2004). The move towards the partial-system (or ‘minimum
realistic’, e.g. Punt & Butterworth 1995) approach leads to a necessity to
define a ‘horizon of relevance’, meaning that components of the ecosystem
that lie beyond this horizon are deemed to be of sufficiently low relevance
to the focus of management that they will not have an important influence
on the outcome of the scenarios being modelled (Schweder (Chapter 21 in
this volume)). However, these partial-system models are challenged by the
problems of diffuse effects (Yodzis 2000) which mean that the horizon of
relevance often lies well beyond our data resources (Plaganyi et al. 2001).
The problems that dog the whole-system approach to modelling marine
ecosystems therefore also dog the partial-system approach.
Like the ‘event horizon’ in cosmology, we contend that the horizon
of relevance in ecosystem modelling is an insurmountable boundary that
severely limits the extent to which we will ever be able to model rationally
constrained management scenarios for biological resources in the oceans
(and perhaps in all complex ecosystems). This is a fairly gloomy outlook
but there may be some hope for the future. This hope comes from two
directions: one involves the potential/possibility that ecosystem dynamics
could be constrained to a narrow set of rules similar to those involved in,
or associated with, the allometry of individual organisms (Garlaschelli et al.
2003); the other direction, which is the one that is explored in this book, is
to reject the reductionist approaches to ecosystem modelling by establish-
ing ecosystem boundaries and only examining ecosystem dynamics at these
boundaries. This is like attempting to understand the crustal dynamics of
the Earth by only looking at surface features. It may be possible to mea-
sure some of the critical outputs of the ecosystem in a way that provides an
insight into the internal dynamics and that could lead to some broad pre-
dictions about the behaviour of the ecosystem, especially when correlated
with known inputs. In biogeochemical terms the inputs and outputs of an
ecosystem involve primary production and the products of respiration plus
the sequestration of organic carbon, in this case as sediment on the seabed.
However, in ecological terms, the outputs could be seen as the terminal
links in food chains, sometimes also known as the top of the food chain.
Moreover, it may also be possible to understand the outputs from the ter-
minal links in the food chains without the necessity of understanding the
intermediate linkages between them and the physical-forcing processes that
are the inputs driving the food-chain dynamics. Many who like to model the
internal dynamics of these systems will consider this to be a leap of faith but,
where the intermediate dynamics have complex properties, there may be no
choice.
Introduction 3
In practical terms, this means using the species at the top of marine food
chains as our indicators of ecosystem status and performance. We refer to
these species as ‘top predators’ but this is synonymous with ‘upper-trophic-
level predators’. For most purposes here we refer to top predators as pin-
nipeds (true seals, sea lions, fur seals and walrus), seabirds, cetaceans and
some large predatory fish. In general, they are species beyond the level of
secondary consumers. This approach has advantages and disadvantages as
outlined below.
Advantages
(1) By definition, top predators are downstream, in terms of energy flow,
of changes within an ecosystem. This means that changes in
ecosystem structure that also affect the energy flows through the
system are likely to be reflected in changes at the top of food chains.
(2) Top predators often exploit marine resources at similar spatial and
temporal scales to those used by man, thus increasing the potential for
competition. It is a truism of marine-ecosystem management that it is
only possible to manage the activities of man; however, the data we
collect about the marine ecosystem – data that come from these
activities – are collected at similar spatial and temporal scales to those
that are relevant to understanding how resource variability is likely to
affect other predators that also forage at the same scales.
(3) Many predators are accessible during important parts of their life
histories mainly because they have terrestrial breeding seasons. This
also constrains their foraging ranges because of their need to return
regularly to the breeding site. Not only does this make it relatively easy
to provide consistent indices of population sizes, it also allows
estimation of regional productivity from the productivity of the
predators themselves. This advantage applies only to seabirds and
pinnipeds, and has the effect of narrowing the focus of interest in
using top predators as measures of ecosystem outputs to these groups.
This bias is reflected in many of the following chapters.
(4) Most of the species used for measuring the outputs from ecosystems
command a high level of public interest and studies of them are likely
to attract support over the long time periods needed to measure these
ecosystem outputs.
Disadvantages
(1) Measuring the changes in top-predator populations or in the behaviour,
performance or productivity of predators does not necessarily titrate
the effects of different management interventions within ecosystems.
4 I. L. Boyd et al.
This book sets out to explore the hypothesis that top predators can be
used in a whole-system approach to managing marine ecosystems. In some
circumstances these predators may also provide information relevant to the
management of specific resources. The emphasis on this hypothesis does
not preclude other approaches or imply that measuring predator responses
will always be informative. However, such an approach could potentially be
part of the set of measures, insights and interpretations used within sophis-
ticated management systems. Such an integrated approach is particularly
useful where there is a need to balance the competing demands for adequate
precaution in setting resource exploitation levels against the economic and
social demands to increase these levels of exploitation still further. It takes
the focus of attention away from the resource being managed and places it
onto the ecosystem in a way that is comprehensible to most components
of the decision-making hierarchy of the management structure and to the
public.
The book represents a collection of case studies and reviews of top preda-
tors as indicators of marine-ecosystem dynamics. Many of these studies are
Introduction 5
We hope that this book will illustrate that we are part of the way to achieving
these objectives.
REFERENCES
Agnew, D. J. (1997). The CCAMLR ecosystem monitoring programme. Antarct.
Sci., 9, 235–42.
Boyd, I. L. (2002). Integrated environment–prey–predator interactions off South
Georgia: implications for management of fisheries. Aquat. Conserv., 12, 119–26.
Chavez, F. P., Ryan, J., Lluch-Cota, S. E., & Niquen, M. (2003). From anchovies to
sardines and back: multidecadal change in the Pacific Ocean. Science, 299,
217–21.
Fowler, C. W. & MacMahon, J. A. (1982). Selective extinction and speciation: their
influence on the structure and functioning of communities and ecosystems.
Am. Nat., 119, 480–98.
Garlaschelli, D., Calderelli, G. & Pietronero, L. (2003). Universal scaling relations
in food webs. Nature, 423, 165–8.
Mori, Y. & Boyd, I. L. (2004). The behavioral basis for nonlinear functional
responses and optimal foraging in Antarctic fur seals. Ecology, 85, 398–410.
Pauly, D., Christensen, V., Dalsgaard, J., Froese, R. & Torres, F., Jr (1998). Fishing
down marine food webs. Science, 279, 860–3.
Pauly, D., Christensen, V., Guénette, S. et al. (2002). Towards sustainability in
world fisheries. Nature, 418, 689–95.
Plaganyi, E. E. & Butterworth, D. S. (2004). A critical look at the potential of
Ecopath with Ecosim to assist in practical fisheries management. Afr. J. Mar.
Sci., 26, 261–87.
Plaganyi, E. E., Butterworth, D. S. & Brandao, A. (2001). Towards assessing the
South African abalone Haliotis midae stock using an age-structured production
model. J. Shellfish Res., 20, 813–27.
Punt, A. & Butterworth, D. S. (1995). The effects of future consumption by the
Cape fur seal on catches and catch rates of the Cape hakes. 4. Modelling the
10 I. L. Boyd et al.
biological interaction between Cape fur seals Arctocephalus pusillus pusillus and
the Cape hake Merluccius capensis and M. paradoxus. S. Afr. J. Mar. Sci., 16,
255–85.
Sydeman, W. J., Hester, M. M., Thayer, J. A. et al. (2001). Climate change,
reproductive performance and diet composition of marine birds in the
southern California Current system, 1969–1997. Prog. Oceanogr., 49, 309–29.
Yodzis, P. (1998). Local trophodynamics and the interaction of marine mammals
and fisheries in the Benguela ecosystem. J. Anim. Ecol., 67, 635–58.
(2000). Diffuse effects in food webs. Ecology, 81, 261–6.
2
Effects of fisheries on ecosystems: just another
top predator?
A . W . T R I T E S , V . C H R I S T E N S E N A N D D . P AU L Y
Humans have long had an association with the coastal regions of the world
and can trace the expansion of civilization to the ready supply of fish, inver-
tebrates and mammals that could be easily gathered, caught or hunted.
Until relatively recently, human activities were restricted to the nearshore
and surface waters. But all of this has changed over the past century as fossil
fuels and the application of new technologies have allowed fisheries to move
further from shore, to fish in deeper waters and to become more effective at
finding and capturing species from all levels of the food chain (Botsford et al.
1997, Merrett & Haedrich 1997, Hutchings 2000, Pauly et al. 2002, 2003).
The geographic and technical expansion of fisheries around the world
through the early to late 1900s was mirrored by steady increases in world
Effects of fisheries on ecosystems 13
90
80
Global catch (× 106 t)
70
60
50
0
1970 1980 1990 2000
Year
Fig. 2.1 Global reported catch with (dotted line) and without (solid line) the
Peruvian anchovy. Total world catches have fluctuated around 82 million tonnes
since the late 1980s due to high catches of Peruvian anchovy which compensated
for the global decline in landings of all other species combined. Adapted from
Watson and Pauly (2001).
catches. However, global landings began stagnating in the early 1980s, and
have decreased since the late 1980s (Watson & Pauly 2001). This decreasing
trend is particularly apparent when the widely fluctuating catch of Peruvian
anchovy is discounted (Fig. 2.1).
Christensen et al. (2003b) reconstructed the biomass of commercially
important fishes that were present in the North Atlantic around 1900.
They considered the abundances of high-trophic-level species such as tuna,
sharks, mackerel, cod, flatfish and salmon – and relied on 23 spatialized
ecosystem models and multiple regressions that considered environmen-
tal and biological factors to predict abundances over spatial resolutions of
1
2
× 21 degree latitudes and longitudes. Plotting the densities of fishes in the
North Atlantic showed the relatively high productivity of the shelf regions
of Europe and eastern North America (Fig. 2.2). However, the data also
showed a major decline between 1900 and 1999 in the range and biomass
of the top predatory fishes that are typically found on dinner tables. Biomass
in the North Atlantic fell by 90% during the twentieth century, leading to
declines of catches throughout the North Atlantic, notably in eastern Canada
(Fig. 2.2). Similar downward trends in the biomasses of high-trophic-
level fishes have been seen elsewhere in the world where they have been
14 A. W. Trites et al.
Biomass Catch
> 11
1900 < 11
> 10
<9
> 2.0
< 2.0
< 1.0
< .75
<8
< .5
<7
< .4
<6
< .3
<5
< .2
<4
< .15
<3
< .1
<2
< .05
<1
< .01
1999
Fig. 2.2 Predicted biomass distributions and estimated fishing intensity for
high-trophic-level fishes (trophic level (TL) ≥ 3.75) in the North Atlantic in 1900
and 1999. Legend indicates biomass in tonnes per square kilometre. Adapted
from Christensen et al. (2003b).
5 Sperm
Beaked whales
whales
Toothed
whales Steller
sea lions
Pisc. Large Deepwater
4 Seals Other
demersal
bird flatfish fish
Baleen octopus
whales Walrus and
bearded seal
Adult pollock
Pelagics Small
Jellyfish Juvenile flatfish
pollock
3 Benthic
feeders
Epifauna
Large
zooplankton
Herbivorous Infauna
2 plankton
1 Phytoplankton Detritus
Trophic levels are the layers that make up food webs, wherein animals
are ranked according to how many steps they are above the primary pro-
ducers at the base of the food web (e.g. Fig. 2.3). Microscopic plants at the
bottom are assigned a trophic level of 1, while the herbivores and detriti-
vores that feed on the plants and detritus make up trophic level 2. Higher-
order carnivores, such as most marine mammals, are assigned trophic lev-
els ranging from 3 to 5 (Pauly et al. 1998b, Trites 2001). Animals that feed
from more than one trophic level have non-integer trophic levels. Thus,
knowing what an animal eats is all that is needed to calculate its trophic
level.
Pauly et al. (1998a) calculated the mean trophic level of reported catches
and found that it had declined over the years. The sharpest declines were
noted in the northwest Atlantic where the mean trophic level dropped from
a peak of 3.7 in 1965 to 2.8 in 1997. Smaller declines were noted in the north-
east Atlantic and the Mediterranean Sea, and have been reported elsewhere
as well (e.g. Arias-Gonzalez et al. 2004, Sanchez & Olaso 2004). Overall,
there has been an average worldwide trophic decline of 0.1 per decade in
16 A. W. Trites et al.
110
100
Maximum length (cm)
90 North Atlantic
80
70
Global coastal
60
50
40
1950 1960 1970 1980 1990 2000
Year
Fig. 2.4 Maximum length attained by species landed from 1950 to 2000.
Adapted from R. Watson, unpublished data.
the mean trophic level of species caught since the 1970s. The inevitable
consequence of such a decline is that species from the lowest trophic levels
may eventually become the mainstay of commercial fisheries.
In addition to progressively catching increasing numbers of species
from lower trophic levels, there are also clear signs that the fish being caught
are no longer as big as they once were. Globally, there has been about a
25% reduction in the mean maximum lengths of landed fish from coastal
waters (Fig. 2.4). These trends reflect the increasing importance to fisheries
of catching smaller species, as well as the selective forces of fisheries which
are resulting in individual fish maturing at smaller sizes.
Reductions in size-at-age and age-at-maturation of commercially
exploited fish have been reported in a number of ecosystems (Trippel 1995,
Rochet 1998, Law 2000). On the Scotian Shelf in eastern Canada, for exam-
ple, average weights of individual demersal fish have decreased by 41% to
51% since the 1970s (Zwanenburg 2000). Declines have also been reported
in the sizes of large demersal fishes in other high-latitude regions, and
have been detected in some – but not all – tropical regions (Pauly 1980,
Bianchi et al. 2000). Such changes in body sizes are particularly troubling,
given that survival and reproduction are functions of body size (Reiss 1989).
Small fish generally incur higher mortality rates and produce fewer eggs
than larger fish. Fisheries thus appear to have the potential to cause evolu-
tionary changes based on the apparent phenotypic and genetic response of
exploited species (Heino & Godo 2002, Hutchings 2004, Olsen et al. 2004).
Effects of fisheries on ecosystems 17
Whether or not such effects of fisheries are long lasting or reversible is not
known. However there are indications that genetic drift may compromise a
population’s recovery after severe depletion (Hutchinson et al. 2003).
The most obvious direct effect of fisheries on marine ecosystems has
been reductions in the abundances of targeted species, such as cod on the
east coast of Canada (Walters & Maguire 1996). Less obvious has been
declines in the spatial ranges and degrees of overlap of depleted targeted
species with others in the ecosystem (e.g. Garrison & Link 2000), as well
as the potential losses of biodiversity (Agardy 2000). Losses of individuals
through bycatch and ghost fishing are poorly documented, but are known
to have directly affected many populations of the larger species such as
seabirds, turtles and marine mammals. Entanglement in fishing gear or
being caught and drowned by baited hooks have threatened a number of
species (e.g. Tasker et al. 2000). Tuna and shrimp fisheries have also been
responsible for large levels of bycatch of unwanted species, but the ecosys-
tem effects of bycatch are less well documented or understood. Finally, there
are the direct and immediate effects of bottom trawling that can modify
benthic habitat and community structure (e.g. Koslow et al. 2000, Jennings
et al. 2001), as well as severely decreasing benthic mega-fauna production
(Hermsen et al. 2003).
Over the past 80 years, food-web research has sought to identify recur-
ring patterns and underlying mechanisms and constraints on ecosystem
structure (e.g. Elton 1927, Lindeman 1942, Pimm 1982, Cohen et al. 1990,
Trites 2003). Many of the conclusions that stem from such studies make
intuitive sense, such as the fact that most food chains are size-structured:
where most predators are larger than their prey and trophic level typi-
cally increases with increasing body size (Pope et al. 1994). Overall, system
biomass appears to be proportional to primary production (Pimm 1982),
and the proportion of species occupying top, intermediate and basal trophic
levels appears to be constant across food webs (Cohen 1978). There also
appears to be a relatively constant ratio of two to three species of prey
for every predator in an ecosystem (Martinez 1991), although numbers of
species of prey consumed by each species of predator tends to increase as
the size of the food web increases.
The lengths of most chains that form food webs are typically short (two
to five links) with maximum reported lengths of eight in tropical shelf seas,
20 A. W. Trites et al.
A number of the parallels and differences that exist between fisheries and
apex marine predators offer additional insights into the effects that com-
mercial fisheries can have on ecosystems. As with fisheries, there is a long-
held belief that marine mammals significantly affect prey populations (Gul-
land 1987, Butterworth 1992, Tamura 2003). This is best demonstrated by
the declines of crabs, abalone and urchin numbers that have followed the
re-introductions and range expansions of sea otters (Estes 1996). Similarly,
declines of sea otters and other marine mammals in the Aleutian Islands
and Gulf of Alaska may be the result of ‘over-hunting’ by killer whales (Estes
et al. 1998, Doroff et al. 2003, Springer et al. 2003). As for other obser-
vations of the direct effects of marine mammals on ecosystems, examples
appear to be lacking. This is probably due to an absence of data on the abun-
dance of prey species prior to the recovery of exploited marine mammal
populations – or it might reflect a lack of suitable experimental controls or
monitoring of forage species to make proper comparisons and conclusions.
As with trawl fisheries, some apex predators such as walrus and grey
whales can influence the turnover of nutrients by feeding on species that
live in bottom sediments. Marine mammals can also influence growth rates
and the sizes at maturity of their prey – as demonstrated in lakes in Quebec,
Canada which are home to land-locked harbour seals that feed on trout.
Trout in these lakes are younger and spawn at younger ages than in adjacent
lakes without harbour seals (Power & Gregoire 1978). The trout also grow
faster and attain smaller sizes in the lakes inhabited by harbour seals.
While there are undoubtedly parallels between the effects of fisheries
and those of marine mammals on food chains, there are at least three impor-
tant differences. One is that mammals and all other species that make up
food webs are generally limited by the size of prey they can consume. They
Effects of fisheries on ecosystems 21
also tend to be specialized feeders and hence draw their energy from a very
limited range of trophic levels (e.g. Fig. 2.3); in contrast, humans can feed on
any size of organism at any trophic level. A second major difference between
fisheries and apex predators is that predator populations in naturally occur-
ring systems are regulated through density-dependent processes that limit
reproduction and survival as prey populations decline; however, there is lit-
tle to regulate the rate of fishery catches apart from economic incentives,
which normally increase (rather than decline) as the species becomes rarer.
The third, and perhaps most significant, difference between the two is that
stable marine food webs are the result of a long period of natural selection
and co-evolution between predators and prey, whereas fisheries represent
an abrupt, knife-edged selective force that has potentially destabilizing con-
sequences.
To capture their prey, apex predators have evolved special sensory abil-
ities (e.g. vision and hearing), morphologies (e.g. dentition) and physiolo-
gies (e.g. diving and breath-holding abilities) (Trites 2002). They have also
evolved specialized feeding behaviours to capture prey that move diurnally
up and down the water column, or to capture prey that move seasonally
across broad geographic ranges. In response, fish and other cold-blooded
species of prey have evolved a number of strategies to increase their chances
of survival. One is cryptic coloration, such as flatfish that blend in with the
bottom when viewed from above but have white undersides to avoid detec-
tion when seen from below against a bright sea surface. Many species of
fish, invertebrates and zooplankton take refuge from predators in the deep,
dark waters during the day and move towards the surface to feed under
the cover of night. Another strategy that some species evoke is predator
swamping, such as the large aggregations of spawning salmon and her-
ring which reduce the numerical effect of predators on their prey popula-
tions. Schooling is another anti-predator behaviour that creates confusion
through the sheer volume of stimuli from a fleeing school, making it dif-
ficult for apex predators to actively select and maintain pursuit of single
individuals.
In addition to directly affecting the behaviours and life histories of other
species, some apex predators may also have indirectly influenced the evolu-
tion of non-targeted species in their ecosystems by consuming the predators
of these species. One such example of this is the relative lack of chemical
defences of kelp and other marine algae against urchin predation in systems
that also contained sea otters (Estes 1996). Another is the consumption by
cod of the potential predators and competitors of their young that has effect-
ively resulted in cod ‘cultivating’ their young (Walters & Kitchell 2001).
22 A. W. Trites et al.
CONCLUSIONS
ACKNOWLEDGEMENTS
REFERENCES
Gulland, J. A. (1987). The impact of seals on fisheries. Mar. Policy, 11, 196–204.
Heino, M. & Godo, O. R. (2002). Fisheries-induced selection pressures in the
context of sustainable fisheries. Bull. Mar. Sci., 70, 639–56.
Hermsen, J. M., Collie, J. S. & Valentine, P. C. (2003). Mobile fishing gear reduces
benthic megafaunal production on Georges Bank. Mar. Ecol. Prog. Ser., 260,
97–108.
Hutchings, J. A. (2000). Collapse and recovery of marine fishes. Nature, 406,
882–5.
(2004). The cod that got away. Nature, 428, 899–900.
Hutchinson, W. F., Oosterhout, C. van, Rogers, S. I. & Carvalho, G. R. (2003).
Temporal analysis of archived samples indicates marked genetic changes in
declining North Sea cod (Gadus morhua). Proc. R. Soc. Lond. B, 270, 2125–32.
Jackson, J. B. C., Kirby, M. X., Berger, W. H. et al. (2001) Historical overfishing and
the recent collapse of coastal ecosystems. Science, 293, 629–38.
Jennings, S. & Warr, K. J. (2003). Smaller predator–prey body size ratios in longer
food chains. Proc. R. Soc. Lond. B, 270, 1413–17.
Jennings, S., Dinmore, T. A., Duplisea, D. E., Warr, K. J. & Lancaster, J. E. (2001).
Trawling disturbance can modify benthic production processes. J. Anim. Ecol.,
70, 459–75.
Katona, S. & Whitehead, H. (1988). Are cetacea ecologically important? Oceanogr.
Mar. Biol. Annu. Rev., 26, 553–68.
Koslow, J. A., Boehlert, G. W., Gordon, J. D. M. et al. (2000). Continental slope and
deep-sea fisheries: implications for a fragile ecosystem. ICES J. Mar. Sci., 57,
548–57.
Law, R. (2000). Fishing, selection, and phenotypic evolution. ICES J. Mar. Sci., 57,
659–68.
Lindeman, R. L. (1942). The trophic-dynamic aspect of ecology. Ecology, 23,
399–418.
Martinez, N. D. (1991). Artifacts or attributes? Effects of resolution on the Little
Rock Lake food web. Ecol. Monogr., 61, 367–92.
Merrett, N. R. & Haedrich, R. L. (1997). Deep Sea Demersal Fish and Fisheries.
London: Chapman and Hall.
Myers, R. A. & Worm, B. (2003). Rapid worldwide depletion of predatory fish
communities. Nature, 423, 280–3.
Olsen, E. M., Heino, M., Lilly, G. R. et al. (2004). Maturation trends indicative of
rapid evolution preceded the collapse of northern cod. Nature, 428, 932–5.
Pauly, D. (1980). A new methodology for rapidly acquiring basic information on
tropical fish stocks: growth, mortality and stock-recruitment relationships. In
Stock Assessment for Tropical Small-Scale Workshop, 19–21 September, 1979,
University of Rhode Island, eds. S. Saila & P. Roedel. Kingston, RI: International
Center for Marine Resources Development, pp. 154–72.
Pauly, D., Christensen, V., Dalsgaard, J., Froese, R. & Torres, F., Jr (1998a). Fishing
down marine food webs. Science, 279, 860–3.
Pauly, D., Trites, A. W., Capuli, E. & Christensen, V. (1998b). Diet composition
and trophic levels of marine mammals. J. Mar. Sci., 55, 467–81.
Pauly, D., Christensen, V., Guénette, S. et al. (2002). Towards sustainability in
world fisheries. Nature, 418, 689–95.
26 A. W. Trites et al.
Pauly, D., Alder, J., Bennett, E. et al. (2003). The future for fisheries. Science, 302,
1359–61.
Pew Oceans Commission (2003). America’s Living Oceans: Charting a Course for Sea
Change. A Report to the Nation. May 2003. Arlington, VA: Pew Oceans
Commission.
Pimm, S. L. (1982). Food Webs. London: Chapman and Hall.
(1991). The Balance of Nature? Chicago, IL: University of Chicago Press.
Pope, J. G., Shepherd, J. G. & Webb, J. (1994). Successful surf-riding on size
spectra: the secret of survival in the sea. Phil. Trans. R. Soc., 343, 41–9.
Power, G. & Gregoire, J. (1978). Predation by freshwater seals on the fish
community of Lower Seal Lake, Quebec. J. Fish. Res. Board Can., 35, 844–50.
Reiss, M. J. (1989). The Allometry of Growth and Reproduction. Cambridge, UK:
Cambridge University Press.
Rochet, M. J. (1998). Short-term effects of fishing on life-history traits of fishes.
ICES J. Mar. Sci., 55, 371–91.
Sanchez, F. & Olaso, I. (2004). Effects of fisheries on the Cantabrian Sea shelf
ecosystem. Ecol. Model., 172, 151–74.
Springer, A. M., Estes, J. A., Vliet, G. B, van et al. (2003). Sequential megafaunal
collapse in the North Pacific Ocean: an ongoing legacy of industrial whaling?
Proc. Natl Acad. Sci. U. S. A., 100, 12 223–8.
Tamura, T. (2003). Regional assessments of prey consumption and competition by
marine cetaceans in the world. In Responsible Fisheries in the Marine Ecosystem,
eds. M. Sinclair & G. Valdimarsson. Rome, Italy: Food and Agriculture
Organization; Wallingford, UK: CABI Publishing, pp. 143–70.
Tasker, M. L., Camphuysen, C. J., Cooper, J. et al. (2000). The impacts of fishing
on marine birds. ICES J. Mar. Sci., 57, 531–47.
Trippel, E. A. (1995). Age at maturity as a stress indicator in fisheries. Bioscience,
45, 759–71.
Trites, A. W. (1997). The role of pinnipeds in the ecosystem. In Pinniped
Populations, Eastern North Pacific: Status, Trends and Issues, eds. G. Stone,
G. Goebel & S. Webster. Boston, MA: New England Aquarium, Conservation
Department, pp. 31–9.
(2001). Marine mammal trophic levels and interactions. In Encyclopedia of Ocean
Sciences, eds. J. Steele, S. Thorpe & K. Turekian. London: Academic Press,
pp. 1628–33.
(2002). Predator–prey relationships. In Encyclopedia of Marine Mammals, eds,
W. F. Perrin, B. Wursig & H. G. M. Thewissen. San Diego, CA: Academic
Press, pp. 994–7.
(2003). Food webs in the ocean: who eats who and how much? In Responsible
Fisheries in the Marine Ecosystem, eds. M. Sinclair & G. Valdimarsson. Rome,
Italy: Food and Agriculture Organization, Rome; Wallingford, UK: CABI
Publishing, pp. 125–43.
Trites, A. W., Christensen, V. & Pauly, D. (1997). Competition between fisheries
and marine mammals for prey and primary production in the Pacific Ocean.
J. Northw. Atl. Fish. Sci., 22, 173–87.
Trites, A. W., Livingstone, P., Mackinson, S. et al. (1999). Ecosystem Change and the
Decline of Marine Mammals in the Eastern Bering Sea: Testing the Ecosystem Shift
Effects of fisheries on ecosystems 27
oceans (Murphy et al. 1995, White & Peterson 1996, Connolley 2002,
Trathan & Murphy 2002). Anomalies in sea-ice extent are now thought to
precess eastwards at speeds consistent with the ACC (Murphy et al. 1995).
Similarly, anomalies in other physical factors are also thought to propagate
at speeds consistent with the flow (Jacobs & Mitchell 1996, White & Peter-
son 1996) – including anomalies in atmospheric pressure, wind stress, sea-
surface temperature and sea height. This precession of anomalies has been
described as the Antarctic Circumpolar Wave (ACW) (White & Peterson
1996).
Processes such as that exemplified by the ACW potentially provide an
increased ability to forecast variability over large spatial scales (cf. White &
Cherry 1999). As physical anomalies propagate eastwards with the ACC,
environmental variability should be predictable with lag periods commen-
surate with the flow. For areas where physical variability is a dominant fea-
ture of the ecosystem, the ability to predict change is of ecological and/or
commercial importance (Trathan &Murphy 2002).
(a) (b)
1.0 1.0
Autocorrelation function
Autocorrelation function
0.5 0.5
0.0 0.0
~3 to 4 years ~3 to 4 years
−0.5 −0.5
0 20 40 60 0 20 40 60
Lag (months) Lag (months)
(c) (d)
~0 years
0.4 0.4
Crosscorrelation function
Crosscorrelation function
0.2 0.2
0.0 0.0
−0.2 −0.2
SG1 leads SG1 leads EI Niño leads
Sea-ice leads
~3 years
−0.4 −0.4
Fig. 3.2 (a) Lagged autocorrelation function for sea-ice anomalies (northerly
√
extent) for the Scotia Sea with 95% confidence limits ( ± 1.96/ n where n =
256). (b) Lagged autocorrelation function for SST anomalies at SG1 with 95%
√
confidence limits ( ± 1.96/ n where n = 256). (c) Lagged crosscorrelation
function between SST anomalies at SG1 and Scotia Sea sea-ice anomalies
√
(northerly extent) with 95% confidence limits ( ± 1.96/ n where n = 256).
(d) Lagged crosscorrelation function between SST anomalies in the Pacific
El Niño and SST anomalies at the SG1 region with 95% confidence limits
√
( ± 1.96/ n where n = 256).
where x was the sample mean and α r were the estimated autoregres-
√
sive coefficients of order r. The 95% confidence limits ( ± 1.96/ n
where n = 256) were used to test whether crosscorrelations were
significantly different from zero.
Physical forcing in the southwest Atlantic 35
Spatial and temporal structure also exist in SST at South Georgia, with peri-
odicity equating to an approximate 3- to 4-year cycle (Box 3.1). Previous anal-
yses of SST in this area (Venegas et al. 1997, Trathan & Murphy 2002) have
shown that the correlation at 4-years dominates and is consistent with the 4-
year period reported for the ACW by White and Peterson (1996). However,
Trathan and Murphy (2002) have shown that the period and strength of this
relationship varies through time, being stronger during the late 1980s and
early 1990s when compared with the late 1990s. Our analyses confirm that
this relationship continues to vary and is currently weaker than in previous
years.
Given that there is strong periodicity evident in both sea-ice anoma-
lies and SST anomalies, it is not surprising that strong correlations exist
between the two data series (Box 3.1). This strong relationship between
regional sea-ice and SST is not unexpected, as both datasets reflect the link-
age between the ocean and the atmosphere.
Strong correlations also exist between the SSTs observed at South Georgia
and those found in the Pacific (Box 3.1); the strongest correlations occurred
with the Pacific leading South Georgia. This suggests that SSTs at South
Georgia are related to the Pacific with a lag period of approximately 2 to
3 years, with warmer (colder) water occurring at South Georgia following
warm (cold) events in the Pacific.
36 P. N. Trathan et al.
(a)
Nests
Years of extremely low performance
6000 Chicks 2.0
Success
Total number counted
1.5
Breeding success
4000
1.0
2000
0.5
0 0.0
1989
1990
1991
1992
1993
1994
1995
1996
1997
1998
1999
2000
2001
2002
2003
Year
1989 = 1988–9
(b)
Born
Years of extremely low performance 1000
1000 Died
Survivors
800 800
Total number counted
400 400
200 200
0 0
1989
1990
1991
1992
1993
1994
1995
1996
1997
1998
1999
2000
2001
2002
2003
Year
1989 = 1988–9
Fig. 3.3 (a) The reproductive performance of gentoo penguins showing the
number of nests, the number of chicks and the breeding success (average
number of chicks fledged per nest per year) of the study population over the
period 1988/9 to 2002/4. (b) The reproductive performance of Antarctic
fur seals showing the number of live births, the number of deaths and the
number of pups surviving in the study population over the period 1988/9
to 2002/3.
Start of Dec
nesting Nov
Oct
Sep
Aug
Jul
Jun
May
Apr
Mar
Feb
Jan
Dec
Nov
Oct
(b)
5000
4000
Fledging success
3000
2000
1000
0
0.5 1.0 1.5 2.0
SST (°C)
Fig. 3.4 (a) Crosscorrelation coefficients between monthly SST at SG1 and
gentoo penguin reproductive output (number of chicks fledged). The dashed
lines indicate the 95% confidence intervals. (b) Averaged November-to-January
SST against the number of chicks fledged; the dashed line indicates the fitted
polynomial regression (F3, 11 = 12.24, R2 = 0.7696, p = 0.001; predicted chicks
fledged = 4871.54 – (2412.99 × SST) + (3242.86 × SST2 ) – (1434.23 × SST3 )).
(b)
600
Pup survival
500
400
300
Fig. 3.5 (a) Crosscorrelation coefficients between monthly SST at SG1 and
Antarctic fur seal pup survival. The dashed lines indicate the 95% confidence
intervals. (b) Averaged November-to-January SST against the number of pups
surviving; the dashed line indicates the fitted polynomial regression
(F3, 11 = 9.84, R2 = 0.7284, p = 0.002; predicted pup survival =
704.84 – (436.49 × SST) + (510.44 × SST2 ) – (177.98 × SST3 )).
accounted for the highest level of variance, with over 72.8% ex-
plained (Fig. 3.5b).
The indices of breeding performance of both the predators con-
sidered in this study showed strong relationships with SST in the
period preceding the breeding season. This suggests that the physi-
cal environment has significant impact on the population processes
of these species and that it affects the manner in which they achieve
breeding condition.
ACKNOWLEDGEMENTS
We thank all of the Zoological Field Assistants who have helped with the
Bird Island long-term monitoring studies. This work was carried out in sup-
port of the British Antarctic Survey DYNAMOE Core Science Programme
at South Georgia.
REFERENCES
Barlow, K. E., Boyd, I. L., Croxall, J. P. et al. (2002). Are penguins and seals in
competition for Antarctic krill at South Georgia? Mar. Biol., 140, 205–13.
Boyd, I. L., Croxall, J. P., Lunn, N. J., & Reid, K. (1995). Population demography of
Antarctic fur seals: the costs of reproduction and implications for life-histories.
J. Anim. Ecol., 64, 505–18.
Boyd, I. L., Staniland, I. J. & Martin, A. R. (2002). Distribution of foraging by
female Antarctic fur seals. Mar. Ecol. Prog. Ser., 242, 285–94.
Brierley, A. S. & Thomas, D. N. (2002). Ecology of Southern Ocean pack ice. Adv.
Mar. Biol., 43, 171–814.
Brierley, A. S., Demer, D. A., Watkins, J. L. & Hewitt, R. P. (1999). Concordance of
interannual fluctuations in acoustically estimated densities of Antarctic krill
around South Georgia and Elephant Island: biological evidence of same year
teleconnections across the Scotia Sea. Mar. Biol., 134, 675–81.
Brierley, A. S., Fernandes, P. G., Brandon, M. A. et al. (2002). Antarctic krill under
sea ice: elevated abundance in a narrow band just south of ice edge. Science,
295, 1890–2.
Brockwell, P. J. & Davis, R. A. (1991). Time Series: Theory and Methods. New York:
Springer-Verlag.
Chatfield, C. (2004). The Analysis of Time Series: An Introduction. London:
Chapman and Hall/CRC.
Christoph, M., Barnett, T. P. & Roeckner, E. (1998). The Antarctic Circumpolar
Wave in a coupled ocean–atmosphere GCM. J. Clim., 11, 1659–72.
Connolley, W. M. (2002). Long term variability of the Antarctic Circumpolar Wave.
J. Geophys., 108, article 8076.
Croxall, J. P. & Prince, P. A. (1979). Antarctic seabird and seal monitoring studies.
Polar Rec., 19, 573–95.
(1980). The food of Gentoo Penguins Pygoscelis papua and Macaroni Penguins
Eudyptes chrysolophus at South Georgia. Ibis, 122, 245–53.
Croxall, J. P., Prince, P. A. & Ricketts, C. (1985). Relationships between prey
life-cycles and the extent, nature and timing of seal and seabird predation in
the Scotia Sea. In Antarctic Nutrient Cycles and Food Webs, eds. W. R. Siegfried,
P. R. Condy & R. M. Laws. Berlin: Springer-Verlag, pp. 516–33.
Croxall, J. P., McCann, T. S., Prince, P. A. & Rothery, P. (1988). Reproductive
performance of seabirds and seals at South Georgia and Signy Island, South
Orkney Islands, 1976–1987: implications for Southern Ocean monitoring
studies. In Antarctic Ocean and Resources Variability, ed. D. Sahrhage. Berlin:
Springer-Verlag, pp. 261–85.
Physical forcing in the southwest Atlantic 43
Everson, I. (2000). Role of krill in marine food webs: the Southern Ocean. In Krill:
Biology, Ecology and Fisheries, ed. I. Everson. Oxford UK: Blackwell Science,
pp. 194–201.
Fedorov, V. & Philander, S. G. (2000). Is El Niño changing? Science, 288,
1997–2002.
Fedulov, P. P., Murphy, E. J. & Shulgovsky, K. E. (1996). Environment–krill
relations in the South Georgia marine ecosystem. CCAMLR Sci., 3, 13–30.
Harmer, S. F. (1931). Southern whaling. Proc. Linn. Soc. Lond., Session 142
(1929–30), 85–163.
Hofmann, E. E., Klinck, J. M., Locarnini, R. O., Fach, B. & Murphy, E. J. (1998).
Krill transport in the Scotia Sea and environs. Antarct. Sci., 10, 406–15.
Hunt, G. L., Stabeno, P., Walters, G. et al. (2002). Climate change and control of the
southeastern Bering Sea pelagic ecosystem. Deep-Sea Res. Part II, 49, 5821–53.
IPCC (Intergovernmental Panel on Climate Change) (2001). Climate Change 2001:
The Scientific Basis, eds. J. T. Houghton, Y. Ding, D. J. Griggs et al.
Contribution of Working Group I to the Third Assessment Report of the
Intergovernmental Panel on Climate Change. Cambridge, UK: Cambridge
University Press. Online at http://www.ipcc.ch/.
Jacka, T. H. & Budd, W. F. (1991). Detection of temperature and sea ice extent
changes in the Antarctic and Southern Ocean. In International Conference on
the Role of the Polar Oceans in Global Change, eds. G. Weller, C. L. Wilson &
B. A. B. Severin. Fairbanks, Alaska: Geophysical Institute, University of Alaska,
pp. 63–70.
Jacobs, G. A. & Mitchell, J. L. (1996)., Ocean circulation variations associated with
the Antarctic Circumpolar Wave. Geophys. Res. Lett., 23, 2947–50.
Kemp, S. & Bennet, A. G. (1932). On the distribution and movements of whales on
the South Georgia and South Shetland whaling grounds. Discovery Rep., 6,
165–90.
Lemke, P., Trinkl, E. W. & Hasselmann, K. (1980). Stochastic dynamic analysis of
polar sea ice variability. J. Phys. Oceanogr., 10, 2100–20.
Lunn, N. J. & Boyd, I. L. (1993). Effects of maternal age and condition on
parturition and the perinatal period of Antarctic fur seals. J. Zool. Lond., 229,
55–67.
Mackintosh, N. A. (1972). Life cycle of Antarctic krill in relation to ice and water
conditions. Discovery Rep., 36, 1–94.
Maslennikov, V. V. & Solyankin, E. V. (1988). Patterns of fluctuations in the
hydrological conditions of the Antarctic and their effect on the distribution of
Antarctic krill. In Antarctic Ocean and Resources Variability, ed. D. Sahrhage.
Berlin: Springer-Verlag, pp. 209–13.
McCafferty, D. J., Boyd, I. L., Walker, T. R. & Taylor, R. I. (1998). Foraging
responses of Antarctic fur seals to changes in the marine environment. Mar.
Ecol. Prog. Ser., 166, 285–99.
McGowan, J. A., Cayan, D. R. & Dorman, L. M. (1998). Climate–ocean variability
and ecosystem response in the northeast Pacific. Science, 281, 210–7.
Murphy, E. J. & Reid, K. (2001). Modelling Southern Ocean krill population
dynamics: biological processes generating fluctuations in the South Georgia
ecosystem. Mar. Ecol. Prog. Ser., 217, 175–89.
44 P. N. Trathan et al.
Murphy, E. J., Clarke, A., Symon, C. J. & Priddle, J. (1995). Temporal variation in
Antarctic sea-ice: analysis of a long-term fast-ice record from the South Orkney
Islands. Deep-Sea Res., Part I, 42, 1045–62.
Murphy, E. J., Watkins, J. L., Reid, K. et al. (1998). Interannual variability of the
South Georgia marine ecosystem: Biological and physical sources of variation
in the abundance of krill. Fish. Oceanogr., 7, 381–90.
Nicol, S., Pauly, T., Bindoff, N. L. et al. (2000). Ocean circulation off east
Antarctica affects ecosystem structure and sea-ice extent. Nature, 406, 504–7.
North, A. W. (2004). Mackerel icefish size and age at South Georgia and Shag
Rocks. CCAMLR Sci., 11.
North, A. W., White, M. G. & Trathan, P. N. (1998). Interannual variability in the
early growth rate and size of the Antarctic fish Gobionotothen gibberifrons
(Lonnberg). Antarct. Sci, 10, 416–22.
Peterson, R. G. & White, W. B. (1998). Slow oceanic teleconnections linking the
Antarctic Circumpolar Wave with the tropical El Niño-Southern Oscillation.
J. Geophys. Res., 103, 24 573–83.
Priddle, J., Croxall, J. P., Everson, I. et al. (1998). Large-scale fluctuations in
distribution and abundance of krill: a discussion of possible causes. In
Antarctic Ocean and Resources Variability, ed. D. Sahrhage. Berlin:
Springer-Verlag, pp. 169–82.
Reid, K., Watkins, J. L., Croxall, J. P. & Murphy, E. J. (1999). Krill population
dynamics at South Georgia 1991–1997, based on data from predators and nets.
Mar. Ecol. Prog. Ser., 177, 103–14.
Reynolds, R. W. & Smith, T. M. (1994). Improved global sea surface temperature
analyses using optimum interpolation. J. Clim. Res., 7, 929–48.
Reynolds, R. W., Rayner, N. A., Smith, T. M., Stokes, D.C. & Wang, W. Q. (2002).
An improved in situ and satellite SST analysis for climate. J. Clim., 15, 1609–25.
Staniland, I. J., Taylor, R. I. & Boyd, I. L. (2003). An enema method for obtaining
fecal material from known individual seals on land. Mar. Mamm. Sci., 19,
363–70.
Steele, J. H. (1998). Regime shifts in marine ecosystems. Ecological Applic., 8,
S33–6.
Stössel, A. & Kim, S.-J. (1998). An interannual Antarctic sea-ice–ocean model.
Geophys. Res. Lett., 25, 1007–10.
Tanton, J. L., Reid, K., Croxall, J. P. & Trathan, P. N. (2004). Winter distribution
and behaviour of gentoo penguins Pygoscelis papua at South Georgia. Polar
Biol., 27, 299–303.
Trathan, P. N. & Murphy, E. J. (2002). Sea surface temperature anomalies near
South Georgia: relationships with the Pacific El Niño regions. J. Geophys Res.,
108, article 8075.
Trathan, P. N., Brierley, A. S., Brandon, M. A. et al. (2003). Oceanographic
variability and changes in Antarctic krill (Euphausia superba) abundance at
South Georgia. Fish. Oceanogr., 12, 569–83.
Trenberth, K. E. & Hoar, T. L. (1996). The 1990–1995 El Niño Southern
Oscillation event: longest on record. Geophys. Res. Lett., 23, 57–60.
Tynan, C. T. (1998). Ecological importance of the Southern Boundary of the
Antarctic Circumpolar Current. Nature, 392, 708–10.
Physical forcing in the southwest Atlantic 45
Venegas, S. A., Mysak, L. A. & Straub, D. N. (1997). Evidence for interannual and
interdecadal climate variability in the South Atlantic. Geophys. Res. Lett., 23,
2673–6.
Waluda, C. M., Trathan, P. N. & Rodhouse, P. G. (1999). Influence of
oceanographic variability on recruitment in the Illex argentinus (Cephalopoda:
Ommastrephidae) fishery in the South Atlantic. Mar. Ecol. Prog. Ser., 183,
159–67.
(2004). Synchronicity in southern hemisphere squid stocks and the influence of
the Southern Oscillation and Trans Polar Index. Fish. Oceanogr., 13, 1–12.
White, W. B. & Cherry, N. J. (1999). The influence of the Antarctic circumpolar
wave upon New Zealand temperature and precipitation during
autumn–winter. J. Clim., 12, 960–76.
White, W. B. & Peterson, R. G. (1996). An Antarctic circumpolar wave in surface
pressure, wind, temperature and sea-ice extent. Nature, 380, 699–702.
Whitehouse, M. J., Priddle, J. & Symon, C. (1996). Seasonal and annual change in
seawater temperature, salinity, nutrient and chlorophyll a distributions around
South Georgia, South Atlantic. Deep-Sea Res., 43, 435–43.
Williams, T. D. (1991). Foraging ecology and diet of gentoo penguins Pygoscelis
papua at South Georgia during winter and an assessment of their winter prey
consumption. Ibis, 133, 3–13.
Zwally, H. J., Parkinson, C. L. & Comiso, J. C. (1983). Variability of Antarctic sea ice
and changes in carbon dioxide. Science, 220, 1005–12.
4
The use of biologically meaningful
oceanographic indices to separate the effects of
climate and fisheries on seabird breeding success
B. E. SCOTT, J. SHARPLES, S. WANLESS, O. N. ROSS,
M . F R E D E R I K S E N A N D F . D AU N T
The variability both in timing of the spring bloom and in the seasonal cycle
of primary production at a given location in the North Sea is driven by the
degree of mixing of the water column (Pingree et al. 1975, Simpson 1981,
Fèvre 1986). Therefore, as the depth and speed of tides at any location are
predictable (Pingree et al. 1978, Simpson & Bowers 1981), the variation in
mixing, and hence primary production, is due to the inter-annual differ-
ences in the amount of wind, radiation and freshwater input received at that
location. Thus, local meteorological forcing, such as daily wind speeds and
the amount of sunlight and rain, drive variation in the timing and amount
of production at the lowest trophic levels.
Almost 100 years ago, a hypothesis was formulated that the timing of
the spring bloom, and therefore the availability of appropriate food, would
greatly influence the survival of larval fishes (Hjort 1914). This idea was
expanded upon by Cushing (1975), who coined the match–mismatch theory
stating that high survival of fish larvae is expected in those years when the
timing of spawning and hatching is such that larvae overlap appropriately
with the timing of the spring bloom. Only recently has a study confirmed
that fish recruitment does indeed increase when such an overlap occurs
(Platt et al. 2003). However, the lack of support for the match–mismatch
theory does not stem from a scarcity of studies addressing this question.
Instead, it reflects the difficulties associated with sampling marine ecosys-
tems repeatedly over appropriate temporal and spatial scales required to
50 B. E. Scott et al.
simultaneously establish the timing of the spring bloom and estimate its
effect on fish survival and growth.
Although the environmental features that trigger spring blooms have
long been well understood in a general sense (Mann & Lazier 1996,
Miller 2004), it is only recently that physical oceanographic modelling
has advanced sufficiently to capture accurately the biological dynamics of
these events at the temporal and spatial scales appropriate to the feed-
ing behaviour of individual animals (Franks 1992, Sharples 1999, Waniek
2003). These types of models, in particular the one-dimensional physical–
biological coupled model of Sharples (1999), allow the monitoring, in one
location, of daily or hourly changes in vertical structure of the water col-
umn and the amount of primary production arising at any given depth at
that location. However, marine ecologists are still some way from under-
standing the impact of between-year variation in the seasonal production
cycle on higher trophic levels. This is because it is prohibitively expensive
to continuously and simultaneously sample phytoplankton, zooplankton,
and larval and adult fish. Therefore, a way to improve our understanding of
marine-ecosystem functioning is to combine the quantitative predictions of
these coupled physical–biological models with concurrent measurements
of the foraging behaviour and breeding success of highly visible top preda-
tors such as seabirds. If the top predators can be shown to be good inte-
grators of important signals being amplified as they move up the trophic
levels, then we will have more reliable and immediate indicators of the cur-
rent state of the ecosystem (Bertram et al. 2001, Gjerdrum et al. 2003).
Our study area (Firth of Forth, 55◦ 30 to 57◦ N, 3◦ W to 0◦ 30 E) contains
two of the hydrographic regions found within the North Sea (Otto et al.
1990): ‘Bank regions’ and ‘Shallow sea front regions’ (see Box 4.1). Both
these water types are important foraging areas for seabirds breeding on the
Isle of May, one of the main colonies in the area (see Daunt et al. (Chapter
12 in this volume) and Camphuysen et al. (Chapter 6 in this volume), for a
description of the foraging distributions of these seabirds).
We used a one-dimensional physical–biological coupled model
(Sharples 1999, see Box 4.2) to derive the inter-annual variability in
the seasonal primary production patterns within these two regions (see
Box 4.3). The one-dimensional model was parameterized using daily,
Biologically meaningful oceanographic indices 51
Moorings
In order to monitor at fine temporal and vertical resolution and
to collect the data needed to parameterize the one-dimensional
physical–biological model for the study area, moorings were placed
in two regions in which seasonal production cycles were expected
to differ: the bank region (depth 45 m, 56◦ 15 N, 02◦ 00 W) and
the shallow sea front region (depth 65 m, 56◦ 15 N, 01◦ 15 W).
The moorings provided information, at a 10-min resolution, on the
changes in vertical structure (at 5- to 10-m intervals), such that it
was possible to define the depth of the surface mixed layer and
52 B. E. Scott et al.
Primary production during the spring bloom provides food for zooplank-
ton populations which, in turn, are the main food source of sandeels
(Covill 1959, Monteleone & Peterson 1986). Therefore the timing, length
Biologically meaningful oceanographic indices 53
16 Surface temperature 14
Bottom temperature
14 12
10
12
o
6 2
4 0
1 31 61 91 121 151 181 211 241 271 301 331 361
Julian date
and intensity of the bloom are all potentially important factors in determin-
ing food availability and hence the time spent foraging by larval, juvenile
and adult sandeels. Sandeels in the North Sea spend the vast majority of
their life buried in the sand (Reay 1970, Winslade 1974a, 1974b, 1974c,
Pearson et al. 1984). In the Firth of Forth area they may only come out of the
sands to feed between April and September (Worsøe 1999). In addition, the
breeding component of the population emerges to spawn in late Decem-
ber and early January, and the eggs hatch by late February (Reay 1970,
Winslade 1974b). Once hatched, the distance over which the larvae may
be advected from spawning locations appears to be variable and dependent
on wind speeds, wind directions and how fast the larvae attain the ability to
make vertical migrations (Proctor et al. 1998, Munk et al. 2002, Jensen et al.
2003).
As soon as sandeels leave the protection of the sands and forage within
the water column, they are subject to predation by a wide range of preda-
tors – such as larger fish (Greenstreet (Chapter 15 in this volume)), seabirds
(Daunt et al. (Chapter 12 in this volume)) and marine mammals. Therefore,
for sandeels to leave the substrate, the gain from food intake must override
predation and starvation risks. In fact, it has been shown experimentally
that low food availability significantly increased the time sandeels remained
54 B. E. Scott et al.
The breeding season is the most energetically demanding part of the seabird
life cycle, and a successful outcome is critically dependent on the availabil-
ity of sufficient amounts of high-quality food. If the initiation of the annual
increase in primary production is the driving factor for the emergence of
adult sandeels, timing of the spring bloom may be very important for the
birds. The availability of adult sandeels at the right time is important in the
early stages of the breeding season (egg laying and incubation) and thus may
be a critical factor determining annual breeding success in seabirds (ICES
2004). During the chick-rearing period (typically in June), adult sandeels
seemingly become less available as they disappear out of the diets of both
kittiwakes and guillemots (Harris & Wanless 1985, Lewis et al. 2001). A
likely explanation is that adult sandeels spend more time in the sands as the
availability of their own food is declining; by this time of the year, primary
production is falling rapidly due to the lack of free nutrients for phytoplank-
ton growth (Miller 2004). Therefore, the birds must now depend more on
juvenile sandeels and other prey species (such as sprat Sprattus sprattus) to
feed their chicks and themselves. Spring conditions and their effect on the
timing of primary production will have influenced the growth and survival
of juvenile fish. It is therefore reasonable to assume that fledging success
is also influenced by the timing and location of spring blooms. In short,
spring-bloom timing is expected to influence all components of breeding
success.
Biologically meaningful oceanographic indices 55
(a)
Effect of spring-bloom date
Kittiwakes (F1,16 = 6.20, P = 0.02)
1.4
Breeding success (chicks per pair) Effect of fishery
(F1,16 = 44.92, P < 0.001)
1.2
No interaction
1.0 (F1,16 = 2.40, P = 0.14)
0.8
0.6
0.4
0.2
0
95 100 105 110 115 120 125
Spring bloom in bank region (Julian date)
(b)
0.90
Breeding success (chicks per pair)
Guillemots
0.85
0.80
0.75
0.70
0.65
0.60
95 100 105 110 115 120 125
Spring bloom in bank region (Julian date)
Fig. 4.2 Breeding success of (a) black-legged kittiwakes (1985–2003) and (b)
common guillemots (1982–2003) on the Isle of May in relation to the start date
of the spring bloom in the bank region, as estimated by the one-dimensional
physical–biological model. Years with no commercial fishery for sandeels are
represented by filled squares and years with a fishery with open squares.
was no evidence that the effect of date of the spring bloom on breeding
success was different in fishing and non-fishing years (interaction: F1,16 =
2.40, p = 0.14), but breeding success was 0.33 ± 0.05 (mean ± SE) chicks
higher in years without fishing than in years with fishing (F1,16 = 44.92,
p < 0.001). The final model containing both the effects of spring-bloom date
and the sandeel fishery explained 74% (p < 0.001) of the variance in breed-
ing success. Separating years with or without a fishery, the effect of climate
alone explained 56% of the variance in breeding success in years without a
fishery and 10% of the variance in years with a fishery. This suggests that
important climatic variables are more easily identified in the absence of the
confounding effects of a fishery.
A similar analysis carried out for guillemots, revealed that neither the
timing of the spring bloom, nor stratification in either oceanographic
region, had a significant effect on breeding success (Fig. 4.2).
DISCUSSION
Although we have found here that for some species breeding success
is linked to annual variation in the timing of spring blooms, seabird pop-
ulation growth is also affected by other demographic parameters. Indeed,
because seabirds are long-lived, population growth rate is most sensitive
to variation in adult annual survival (Croxall & Rothery 1991). Outside the
breeding season, Isle of May seabirds range much more widely than our
study area, in some cases over the entire North Atlantic Ocean. Seabirds
only recruit into the breeding population when they are several years old,
and during the pre-breeding period they range even more widely than
adults. As encouraging as our present results are, identifying, measuring
and modelling oceanographic variables at the appropriate spatial and tem-
poral scale to understand interactions between seabird survival and recruit-
ment still presents a major challenge.
More studies of this kind involving the use of one-dimensional physical–
biological models as tools for connecting past and predicted changes in
climate to higher trophic levels will bring us closer to identifying critical
linkages within ecosystems. In a constantly changing environment, where
future climate change is likely to have profound consequences for marine
ecosystems, these models could prove invaluable tools for understanding
and predicting impacts on higher trophic levels.
ACKNOWLEDGEMENTS
REFERENCES
Aebischer, N. J., Coulson, J. C. & Colebrook, J. M. (1990). Parallel long-term trends
across four marine trophic levels and weather. Nature, 347, 753–5.
Bertram, D. F., Mackas, D. L. & McKinnell, S. M. (2001). The seasonal cycle
revisited: interannual variation and ecosystem consequences. Prog. Oceanogr.,
49, 283–307.
60 B. E. Scott et al.
Lewis, S., Wanless, S., Wright, P. J. et al. (2001). Diet and breeding performance of
black-legged kittiwakes Rissa tridactyla at a North Sea colony. Mar. Ecol. Prog.
Ser., 221, 277–84.
Mann, K. H. & Lazier, J. R. N. (1996). Dynamics of Marine Ecosystems. Oxford, UK:
Blackwell Science.
Miller, G. B. (2004). Biological Oceanography. Oxford UK: Blackwell Science.
Monaghan, P. (1992). Seabirds and sandeels: the conflict between exploitation and
conservation in the northern North Sea. Biodiversity Conserv., 1, 98–111.
Monteleone, D. M. & Peterson, W. T. (1986). Feeding ecology of American sand
lance Ammodytes americanus larvae from Long Island Sound. Mar. Ecol. Prog.
Ser., 30, 133–43.
Munk, P., Wright, P. J. & Pihl, N. J. (2002). Distribution of the early larval stages of
cod, plaice and lesser sandeel across haline fronts in the North Sea. Estuar.
Coast. Shelf Sci., 55, 139–49.
Otto, L., Zimmerman, J. T. F., Furnes, G. K. et al. (1990). Review of the physical
oceanography of the North Sea. Neth. J. Sea Res., 26, 161–238.
Pearson, W. H., Woodruff, D. L. & Sugarmann, P. C. (1984). The burrowing
behaviour of sand lance, Ammodytes hexapterus: effects of oil-contaminated
sediment. Mar. Environ. Res., 11, 17–32.
Pingree, R. D., Pugh, P. R., Holligan, P. M. & Forster, G. R. (1975). Summer
phytoplankton blooms and red tides along tidal fronts in the approaches to the
English Channel. Nature, 258, 672–7.
Pingree, R. D., Bowman, M. J. & Esaias, W. E. (1978). Headland fronts. In Oceanic
Fronts in Coastal Processes, eds. M. J. Bowman & W. E. Esaias. Berlin:
Springer-Verlag, pp. 78–86.
Platt, T., Fuentes-Yaco, C. & Frank, K. T. (2003). Spring algal bloom and larval fish
survival. Nature, 423, 398–9.
Proctor, R., Wright, P. J. & Everitt, A. (1998). Modelling the transport of larval
sandeels on the north-west European shelf. Fish. Oceanogr., 7, 347–54.
Ratcliffe, N. (2004). Causes of seabird population change. In Seabird Populations of
Britain and Ireland, eds. P. I. Mitchell, S. F. Newton, N. Ratcliffe & T. E. Dunn.
London: T. & A. D. Poyser, pp. 407–37.
Reay, P. J. (1970). Synopsis of Biological Data on North Atlantic Sandeels of the genus
Ammodytes (A. tobianus, A. dubius, A. americanus and A. marinus). FAO
Fisheries Synopsis 82. Rome, Italy: FAO.
Rindorf, A., Wanless, S. & Harris, M. P. (2000). Effects of sandeel availability on
the reproductive output of seabirds. Mar. Ecol. Prog. Ser., 202, 241–52.
Sharples, J. (1999). Investigating the seasonal vertical structure of phytoplankton
in shelf seas. Mar. Models, 1, 3–38.
Sharples, J., Ross, O. N., Scott, B. E., Greenstreet, S. P. R. & Fraser, H. Inter-annual
variability in the timing of stratification and the spring bloom in a temperate
shelf sea. Cont. Shelf Res, in press.
Simpson, J. & Bowers, D. (1981). Models of stratification and frontal movement in
shelf seas. Deep-Sea Res. Part I, 28, 727–38.
Simpson, J. H. (1981). The shelf-sea fronts: implications of their existence and
behaviour. Phil. Trans. R. Soc. Lond. A, 302, 531–46.
62 B. E. Scott et al.
Variables Period
Behavioural
Individual Mean duration, depth, bottom time, % time at depth, 1992–2001
dives % square dives, descent and ascent rates, surface
time between dives, number of dives per day, total
duration diving per day, total bottom time per day
Dive bouts Mean duration, depth, % bout at depth, % square dives 1992–2001
per bout, % V-shaped dives per bout, number of
dives per bout, total time in bout per day
Diet % species composition, diversity, energy density 1994–2002
Life history
Maternal Postpartum body mass, lactation length 1992–2001
Offspring Weaning body mass, change in male and female 1992–2001
weaning mass
diet at any one time or place (Bowen et al. 1993, Bowen & Harrison 1994),
probably reflecting local prey abundance. Given their broad geographic dis-
tribution and their accessibility at the main breeding site on Sable Island,
grey seals may provide an opportunity to monitor changes in the Scotian
Shelf ecosystem. Here we address two questions: (1) do behavioural, dietary
and life-history variables of grey seals vary over time; (2) do these responses
correlate with particular features of environmental variability?
DATA COLLECTION
The data were collected from 1992 to 2003 on Sable Island (43◦ 55 N,
60◦ 00 W), a crescent-shaped, partially vegetated sandbar approximately
300 km southeast of Halifax, Nova Scotia, Canada. Sable Island is the largest
haul-out and breeding colony for grey seals in the northwest Atlantic popu-
lation. Seals congregate on the island in May and June to moult and again
in late December and January to rear offspring and mate. Thousands of
grey seals also haul out on the island throughout the year between foraging
trips. We studied a suite of behavioural, dietary and life-history variables to
investigate responses of grey seals to environmental variability (Table 5.1).
DISTRIBUTION
Fig. 5.1 Annual distribution of grey seals based on locations of 70 adults fitted
with Argos satellite tags on Sable Island (n = 18 for May/June, n = 38 for
September/October, n = 14 for January). To determine the spatial distribution
of the population (i.e. percentage usage), we divided the study area into 5 × 5
cells and counted the number of seals that entered each cell. Multiple use of a
cell by an individual seal was scored as a single use to avoid biasing the
population distribution by the behaviour of individual seals. Too few seals were
tagged in each year to permit the analysis of inter-annual changes in the use of
space. The 100-m isobath is indicated by a grey line.
Fig. 5.2 Percentage of square-shaped dives ( ± 1 SE) used by grey seals between
September and December of each year. Number of seals studied is given above
the symbol. Percentage of square-shaped dives differed among years (MANOVA:
F8,81 = 2.4, p = 0.024).
RESPONSE VARIABLES
Diving behaviour
(a)
200
180
Dives per day
160
140
120
100
1990 1992 1994 1996 1998 2000 2002
(b)
20
18
Time in foraging bouts (hours per day)
16
14
12
10
4
1990 1992 1994 1996 1998 2000 2002
Year
Fig. 5.3 Inter-annual indices of foraging effort ( ± 1 SE) (a) Mean number of
dives per day. (b) Time spent in foraging bouts. Sample sizes are as given in Fig.
5.2. Based on MANOVA, both indices varied inter-annually: dives per day,
F 8,81 = 2.9 and p = 0.007; hours in foraging bouts, F7,74 = 2.7 and p = 0.014.
Two of the three indices (number of dives per day, time spent in for-
aging bouts and time spent at depth) of foraging effort varied significantly
among years. Dives per day increased through the mid 1990s and declined
through the late 1990s into 2001 (Fig. 5.3a). Hours spent in foraging bouts
was relatively stable from 1992 to 1997, then declined significantly in
Predator behaviour and environmental variability 69
Diet
0.50
0.40
0.35
0.30
0.25
0.20
(b) 1.0
Other demersals
0.9 Squid
Flounders
0.8
Gadoids Small pelagics
0.7
0.6
Proportion
0.5
Sand lance
0.4
0.3
0.2
Redfish
0.1
0.0
1993 1994 1995 1996 1997 1998 1999 2000 2001
Year
Life history
ENVIRONMENTAL VARIABILITY
Fig. 5.6 Inter-annual estimates of the difference in mean weaning mass of male
and female grey seal pups.
the Scotian Shelf are among the most variable in the North Atlantic. On
the northeastern Scotian Shelf the cold, intermediate-layer water, repre-
sented by Misaine Bank at 100 m depth, fell sharply by the mid 1980s
and remained below normal through 1995, returning to the climatological
mean in the period 1997 to 2002 (DFO 2003). Despite these changes, the
estimated mean composition of the winter grey seal diet was similar during
the cold period and during years when temperature had returned to average
conditions (MANOVA: F8,322 = 1.8, p = 0.09; Table 5.2).
There were large inter-annual changes in the estimated biomass of some
prey species during the 1990s (Fig. 5.7). However, to a considerable degree
the magnitude of those changes are more difficult to assess as the bottom-
trawl inconsistently samples species such as redfish, sand lance, capelin and
←
Fig. 5.5 Inter-annual estimates of: (a) maternal postpartum mass (MPPM),
(b) lactation length and (c) pup weaning mass of grey seals. Estimated means of
MPPM did not differ among years (univariate general linear model (GLM) with
maternal age as a covariate, F9, 267 = 1.4, p = 0.21); estimated mean lactation
length among years (adjusted for differences in mean maternal age) differed
inter-annually (univariate GLM with maternal age and maternal age 5 squared as
covariates, F10, 811 = 2.0, p = 0.03); estimated mean weaning mass varied
among years (univariate GLM with maternal age and maternal age 5 squared as
covariates, F10, 1328 = 2.4, p = 0.01). Error bars are 95% confidence limits.
74 W. D. Bowen et al.
Period
a
Capelin (Mallotus villosus), herring (Clupea harengus), mackerel (Scomber scombrus),
snakeblenny (Lumpenus lumpretaeformis), gaspereau (Alosa pseudoharengus).
b
Pollock, haddock (Melanogrammus aeglefinus), lumpfish (Cyclopterus lumpus).
c
American plaice (Hippoglossoides platessoides), yellowtail ( Limanda ferruginea), witch
flounder, winter flounder, turbot (Rheinhardtius hippoglossoides).
d
Thorny skate (Raja radiata), winter skate (Raja ocellata).
300
200
100 Skates
Flounders
0 Pollock
1990 1992 1994 1996 1998 2000 2002
8000
6000
Biomass (× 1000 t)
4000
2000 Redfish
Capelin
0 Sand lance
1990 1992 1994 1996 1998 2000 2002
Year
Fig. 5.7 Biomass estimates of selected grey seal prey from July bottom-trawl
surveys, 1990–2002. Inter-annual variability in fish abundance was derived from
synoptic, stratified-random, bottom-trawl surveys conducted each July. Estimates
of species biomass, corrected for catchability to provide a better indication of true
relative abundance, were combined for the trawl survey strata primarily used by
grey seals. However, species such as sand lance, redfish, capelin and pollock are
poorly sampled by bottom trawls such that the resulting biomass estimates are
biased and observed trends may provide only a rough indication of true trends.
76 W. D. Bowen et al.
and commercial vessels. Nevertheless, the use of predators for this purpose
requires an understanding of how predator responses are linked to the vari-
ability in particular ecosystem components (Croxall et al. 1988, Boyd & Mur-
ray 2001, Hindell et al. 2003). We found significant inter-annual variation in
aspects of foraging behaviour, diet and several life-history variables of grey
seals over the 10 years of our study. Presumably, differences in foraging
behaviour and diet are causal, reflecting the need to use different foraging
tactics to locate and capture different prey species (e.g. redfish versus sand
lance) (Bowen et al. 2002). Similarly, differences in behaviour and diet are
presumably related to changes in the availability of prey. However, the links
among these variables are not clear in our data.
The continental-shelf ecosystems inhabited by grey seals in east-
ern Canada have exhibited considerable variability over the past several
decades – involving changes in physical and biological oceanography, fish-
eries exploitation rates and species abundance – with a general shift from
a system dominated by demersal fishes to one dominated by pelagic fish
species (Rice 2000, Swain & Sinclair 2000, Zwanenburg et al. 2002). Thus,
there were considerable changes in ecosystem state to test whether grey
seals revealed those changes. However, only grey seal pup production at
Sable Island was monitored over those earlier several decades. Measure-
ments of the behavioural, diet and life-history variables were only initiated
in the early 1990s after many of the larger changes had already occurred.
Grey seals are large, long-lived mammals with K-selected life histo-
ries. Despite the large environmental changes observed over the past four
decades on the Scotian Shelf (Zwanenburg et al. 2002), the grey seal pop-
ulation size has increased steadily from only a few thousand seals in the
1960s to about 175 000 in 1995 (Mohn & Bowen 1996). Pup production
on Sable Island increased exponentially, at a rate near the maximum possi-
ble (rmax ), through the late 1990s (Bowen et al. 2003). Although there is no
a-priori reason to have expected exponential population growth, the fact that
it occurred suggests that – from a grey seal perspective – the environment
was favourable throughout this period. This demographic history provides
an essential context for interpreting the performance of the response vari-
ables measured in this study.
Diving behaviour ought to reflect characteristics of the prey available to
pinnipeds since all foraging necessarily occurs during diving. The relation-
ship between diving behaviour and changes in prey availability is perhaps
best understood in Antarctic fur seals (Arctocephalus gazella) (e.g. Bengtson
1988, Boyd et al. 1994, McCafferty et al. 1998). Females in this species
altered both trip duration and number of dives in response to changes in
Predator behaviour and environmental variability 77
krill abundance and the amount of fish and squid in the diet. However,
these conclusions are limited to the period of offspring provisioning and
thus may not be representative of responses at other times of the year, or
in males. We studied diving behaviour of adult male and female grey seals
over the 4 months prior to arrival at the breeding colony. During this period
both sexes gain mass (Beck et al. 2003c), indicating that this is a period of
heavy feeding. Although most variables describing individual dives or bouts
of dives exhibited little inter-annual variability, number of dives per day, pro-
portion of square-shaped dives, proportion of dive bout spent at depth and
total time spent in diving bouts per day varied among years. However, for
the most part, inter-annual variation in foraging behaviour was not related
to differences in diet or estimated prey biomass. Number of dives per day
was positively correlated with the ratio of the two dominant prey in the diet,
redfish and sand lance. However, this finding is difficult to interpret with-
out knowing how predator foraging tactics differ for these prey types.
Inter-annual variation in pinniped diets is generally assumed to reflect
changes in prey abundance and encounter rates (Bowen & Siniff 1999).
Although demersal species accounted for ∼25% of the grey seal diet in
some years, diets were dominated by sand lance and redfish. The percent-
age of those two species in the diet varied significantly among years. How-
ever, there was no correlation between this variation and estimates of prey
biomass from trawl surveys conducted within grey seal habitat. There are
a number of possible reasons for this. Firstly, the estimate of prey biomass
was derived from the survey conducted in July, whereas our diet sam-
ples were collected about 5 months after the survey. Although fatty-acid-
based estimates of diet should integrate intake over several months (Iverson
et al. 2004), both prey availability and grey seal distribution are presumably
dynamic such that the July survey may not be a good measure of prey avail-
able to seals months later. Secondly, the trawl survey is known to sample
both redfish and sand lance inconsistently. Thus, the true abundance of
these species may not be reflected by the survey. Thirdly, the small number
of grey seals sampled in some years (e.g. 1997, 2000 and 2001) may not
have been representative of grey seal diets. Obtaining a representative sam-
ple may be difficult for a wide-ranging predator exploiting a spatially het-
erogeneous habitat. Fourthly, although we know little about the ontogeny
of foraging behaviour in grey seals and other pinnipeds, it is reasonable to
expect that learning plays an important role in the diet of individual seals
resulting in strong individual differences in diet among individuals for-
aging in the same habitat (Estes et al. 2003). Individual prey preferences
may partially obscure responses at the population level, particularly when
78 W. D. Bowen et al.
overall prey resources are not limited. Finally, and perhaps most impor-
tantly, given the favourable prey environment (as judged by the rapid rate of
population increase), it is possible that grey seals were foraging in the range
of the asymptotic limb of the non-linear functional response curve (Furness
(Chapter 14 in this volume)) where consumption is relatively insensitive to
changes in prey biomass. If true, the interpretation of predator responses
will be contingent on demography.
The significance of changes in foraging behaviour and diet to the preda-
tor can only be determined through their effects on demography (Croxall
et al. 1988). However, annual estimates of survival and fecundity are diffi-
cult to measure in most pinnipeds. Maternal and offspring size and con-
dition are attractive because they can be easily measured, ought to reflect
changes in prey availability and can affect demography. We found that
MPPM, an index of foraging success, did not vary among years. Duration
of maternal investment (i.e. lactation length) and pup weaning mass exhib-
ited significant, but relatively little, inter-annual variation. Interestingly, in
the year (1998) that the difference between male and female weaning mass
was greatest, combined pup weaning mass was also the highest, perhaps
suggesting that adult females were in particularly good condition that year.
However, in general, these life-history response variables in grey seals were
not informative of ecosystem state. As noted above, our data were collected
during a period when this population was experiencing favourable envi-
ronmental conditions and exponential population growth. The same vari-
ables measured during a period of population decline or stability may have
responded quite differently.
For the present, we conclude that despite the large changes in estimates
of invertebrate- and fish-species abundances (Zwanenburg et al. 2002),
MPPM, lactation length and offspring weaning mass provided little indica-
tion of those environmental changes. Although we observed greater inter-
annual variability in the foraging behaviour and diet, those response vari-
ables for the most part were also not informative with respect to specific
ecosystem changes that occurred during the 1990s. However, we believe
it would be premature to suggest that grey seals and similar species will
not be useful monitors on the basis of this initial exploratory analysis. It
is possible that grey seal diets are better indicators of abundance for many
of the species consumed than are the bottom-trawl surveys routinely used
for this purpose. Comparison of species estimates in the diet of grey seal
against reconstructed prey-population abundance from catch-at-age models
might provide a means of validating this hypothesis. Measurement of for-
aging behaviour and diet response variables at other times of the year may
Predator behaviour and environmental variability 79
ACKNOWLEDGEMENTS
Iverson, S. J., Frost, K. J. & Lang, S. (2002). Fat content and fatty acid composition
of forage fish and invertebrates in Prince William Sound, Alaska: factors
contributing to among and within species variability. Mar. Ecol. Prog. Ser., 241,
161–81.
Iverson, S. J., Field, C., Bowen, W. D. & Blanchard, W. (2004). Quantitative fatty
acid signature analysis: a new method of estimating predator diets. Ecol.
Monogr., 74, 211–35.
Levin, S. A. (1992). The problem of pattern and scale in ecology. Ecology, 73,
1943–67.
Mahon, R., Brown, S. K., Zwanenburg, K. C. T. et al. (1998). Assemblages and
biogeography of demersal fishes of the east coast of North America. Can. J.
Fish. Aquat. Sci., 55, 1704–38.
McCafferty, D. J., Boyd, I. L., Walker, T. R. & Taylor, R. I. (1998). Foraging
responses of Antarctic fur seals to changes in the marine environment. Mar.
Ecol. Prog. Ser., 166, 285–99.
McLaren, I. A. (1993). Growth in pinnipeds. Biol. Rev., 68, 1–79.
Mellish, J.-A. E., Iverson, S. J. & Bowen, W. D. (1999). Individual variation in
maternal energy allocation and milk production in grey seals and
consequences for pup growth and weaning characteristics. Physiol. Biochem.
Zool., 72, 677–90.
Mohn, R. & Bowen, W. D. (1996). Grey seal predation on the eastern Scotian Shelf:
modelling the impact on Atlantic cod. Can. J. Fish. Aquat. Sci., 53, 2722–38.
Montevecchi, W. A. (1993). Birds as indicators of changes in marine prey stocks. In
Birds as Monitors of Environmental Change, eds. R. W. Furness and J. J. D.
Greenwood. London: Chapman and Hall, pp. 217–66.
Pomeroy, P. P., Fedak, M. A., Rothery, P. & Anderson, S. (1999). Consequences of
maternal size for reproductive expenditure and pupping success of grey seals at
North Rona, Scotland. J. Anim. Ecol., 68, 235–53.
Reid, K. & Croxall, J. P. (2001). Environmental response of upper trophic-level
predators reveals a system change in an Antarctic marine ecosystem. Proc. R.
Soc. Lond. B, 268, 377–84.
Rice, J. C. (2000). Evaluating fishery impacts using metrics of community
structure. ICES J. Mar. Sci., 57, 682–8.
Sherman, K., Solow, A. Jossi, J. & Kane, J. (1998). Biodiversity and abundance of
the zooplankton of the Northeast Shelf ecosystem. ICES J. Mar. Sci., 55, 730–8.
Stobo, W. T., Beck, B. & Horne, J. K. (1990). Seasonal movements of grey seals
(Halichoerus grypus) in the Northwest Atlantic. In Population Biology of
Sealworm (Pseudoterranova decipiens) in Relation to its Intermediate and Seal
Hosts, ed. W. D. Bowen. Can. Bull. Fish. Aquat. Sci., 222, 199–213.
Swain, D. P. & Sinclair, A. F. (2000). Pelagic fishes and the cod recruitment
dilemma in the Northwest Atlantic. Can. J. Fish. Aquat. Sci., 57, 1321–5.
Zwanenburg, K. C. T., Bowen, D. W., Bundy, A. et al. (2002). Decadal changes in
the Scotian Shelf large marine ecosystem. In Large Marine Ecosystems of the
North Atlantic: Changing States and Sustainability, eds. K. Sherman & H. R.
Skjoldal. Amsterdam, the Netherlands: Elsevier Science.
6
Distribution and foraging interactions of
seabirds and marine mammals in the North Sea:
multispecies foraging assemblages and
habitat-specific feeding strategies
C. J. CAMPHUYSEN, B. E. SCOTT AND S. WANLESS
Fig. 6.1 Study area (54◦ to 59◦ N, 2◦ E to the coast) and locations of seabird
colonies and oceanographic areas mentioned in the text. Isobaths for 30-, 50- and
100-m depths are shown, horizontal lines indicate ship-based transects (see Box
6.1).
Distribution and foraging interactions 85
0
20 40 60 80 100 150 200 250 300
(b) 100%
Auks
90%
Tems
80% Kittiwake
Large Larus gulls
70% Small Larus gulls
Skuas
60%
Seaduck
50% Cormorants
Gannet
40% Storm-petrels
Shearwaters
30%
Fulmar
20% Grebes
Divers
10%
0%
Coastal 20 40 60 80 100 150 200 250 300
(c) 10 000
Seabirds
Mar. mammals
1000
Biomass (kg km −2)
100
10
1
20 40 60 80 100 150 200 250 300
Distance strata off UK coast (km)
FORAGING RANGE
were using spatially discrete foraging areas. Comparing these results with
information on diving locations, obtained using activity loggers deployed on
chick-rearing adults on the Isle of May in 2003, showed that there was close
agreement between the two methods, with birds feeding predominantly on
the western side of the Wee Bankie. In addition, the at-sea surveys suggested
that common guillemots from the Farne Islands and St Abb’s Head were
using the southern part of the Marr Bank, while birds from Fowlsheugh
foraged mainly in the northern part (Fig. 6.1). These results indicate max-
imum foraging ranges of 50 km for common guillemots from the Isle of
May, 55 km for St Abb’s Head, 70 km for the Farne Islands and at least 110
km for Fowlsheugh.
FORAGING-HABITAT CHARACTERISTICS
The study area is part of the Northeast Atlantic shelves province of the
Atlantic coastal biome (Longhurst 1999) and contains two distinct hydro-
graphic regions: North Atlantic waters, which occupy most of the central
North Sea, and Scottish coastal waters (Otto et al. 1990, Scott et al. (Chapter
4 in this volume)). During the winter months, lower levels of solar radia-
tion combined with stronger winds and tidal friction leave the water col-
umn throughout the North Sea completely mixed. Only in the spring does
the surface layer in deeper areas begin to warm due to increasing amounts
of sunlight and decreasing winds. This warming creates a difference in den-
sity between the upper and lower layers of the water column and the onset
of the resulting stratification allows plankton to stay above the critical depth
needed for population growth and marks the beginning of seasonal primary
production (Scott et al. (Chapter 4 in this volume)). In shelf seas, shallow
sea fronts, also known as tidal mixing fronts, separate inshore areas that
are permanently vertically mixed due to their shallow depth and/or strong
tidal currents, from areas that stratify due to deeper depths and/or weaker
tidal currents (Simpson 1981, Scott et al. this volume). Top predators fre-
quently congregate around these shallow sea fronts that are associated with
increased abundances of fish, larvae and zooplankton (Pingree et al. 1975,
Pingree & Griffiths 1978, Richardson et al. 1986). The exact locations of the
fronts change over the spring and summer months in response to weather
conditions, and the monthly and daily rhythm of tidal speeds. A ‘stratifica-
tion index’, defined as the difference in density between the sea surface and
the bottom, can be used to identify the locations of fronts (Heath & Brander
2001). The offshore boundary of the area used by many seabirds and marine
mammals repeatedly identified from at-sea surveys, typically coincided with
Distribution and foraging interactions 89
this frontal zone, where the stratification index ranged from 0.6 to 0.8
(cf. Ollason 2000).
Small, short-lived MSFAs (Box 6.2) were frequently recorded in the coastal
foraging zone, particularly around the shallow sea front. The tendency to
participate in such MSFAs differed among the various species (Table 6.1).
Black-legged kittiwakes frequently acted as catalysts or initiators in MSFA
formation, large gulls and skuas quickly joined in, with the former acting
as scroungers or suppressors, while the latter were peripheral, aerial klep-
toparasites (see Box 6.2 for definitions of these terms). Small species such
as storm-petrels and terns rarely joined feeding aggregations, except at the
periphery, possibly because such birds are likely to lose out in direct com-
petition with other predators. Auks were normally joined by other seabirds
and rarely joined existing aggregations (0.3% of cases, n = 3277 MSFAs
recorded within 100 km of the coast). The most common type of MSFA in
coastal waters formed over groups of feeding common guillemots and/or
razorbills (76%, n = 3277), puffins (13%) or harbour porpoises (3%). Within
40 km of the coast, about one-quarter of MSFAs (26%, n = 1518) were tar-
geted by large Larus gulls, and the arrival of these species rapidly prevented
further access by catalysts. In contrast, only 6% (n = 1759) of MSFAs more
than 40 km from land were targeted by large gulls, and black-legged kitti-
wake foraging activities tended to be concentrated in these aggregations.
The apparent avoidance by black-legged kittiwakes of the inshore areas
used by the large gulls resulted in a counter-intuitive, positive relationship
between kittiwake annual breeding success and foraging range (rS = 0.68,
n = 9, p < 0.05) such that success tended to be lower in years when the
shallow sea front occurred closer inshore. Northern gannets joined 18% of
MSFAs (n = 3277), and their arrival typically rapidly disrupted the forag-
ing opportunities of all the other participants, including other gannets and
auks.
Large differences in feeding activity, as well as in the frequency of occur-
rence of MSFAs, were recorded when comparing transects crossing the
shallow sea front. On some occasions only large flocks of inactive (resting or
preening) seabirds were encountered while on others high numbers of birds
and MSFAs were recorded. A dedicated cruise in 2003 revealed that forag-
ing activity in these areas varied during the day in relation to changes in tidal
currents, suggesting that physical processes may help drive prey towards the
Table 6.1. Proportion of surface-feeding and plunge-diving seabirds participating in MSFAsa , behavioural characteristics and
role within MSFAs (see Box 6.2), main feeding areab and the type of diving predator producing the MSFA (see Box 6.2)
MSFA
Species (%) Feeding behaviour and MSFA role Distance Producer
surface (Camphuysen & Scott 2003). Inactive periods were recorded more
frequently in black-legged kittiwakes than in common guillemots (Fig. 6.4),
with the latter continuing to feed at certain phases of the tide when black-
legged kittiwakes had stopped entirely. Common guillemot feeding activ-
ity was more evenly spread over the day than that of kittiwakes. Clearly
more surveys are needed to investigate these interspecific differences
further.
92 C. J. Camphuysen et al.
Surface feeders
Joiners
Joiners
(Scroungers,
(Scroungers, (Initiators) kleptoparasites)
kleptoparasites) (Catalysts) (Suppressors)
Sea surface
Fish
ball
Fig. 6.3 Schematic representation of an MSFA with diving auks and facilitated
surface-feeders. (Redrawn from Camphuysen and Webb (1999).)
40
Kittiwakes
35 Guillemots
Numbers per 5-min observation
30
25
20
15
10
0
Inc. south Max. south Dec. south Inc. north Max. north Dec. north
Flow direction of tidal current
DISCUSSION
ACKNOWLEDGEMENTS
REFERENCES
Bailey, R. S., Furness, R. W., Gauld, J. A. & Kunzlik, P. A. (1991). Recent changes in
the population of the sandeel (Ammodytes marinus Raitt) at Shetland in relation
to estimates of seabird predation. ICES Mar. Sci. Symp., 193, 209–16.
96 C. J. Camphuysen et al.
Polar bears (Ursus maritimus) are broadly distributed in the Arctic and, as
such, have the potential to provide information about changes in ecosystem
structure and functioning over broad scales in time and space. Yet, because
they are so wide-ranging and difficult to observe, there are few quantita-
tive data on polar bear diets or on the ecological (e.g. climate change) and
demographic factors that influence prey selection. We used quantitative
fatty acid signature analysis of polar bear adipose tissue to estimate their
diets in the 1980s/90s across three major regions of the Canadian Arctic:
Davis Strait (n = 70), western Hudson Bay (n = 217) and the Beaufort Sea
(n = 34), using a database of the major prey species in each region (n =
292). Although polar bears consumed ringed and bearded seals throughout
their range, diets differed greatly among regions. Ringed seals accounted
for ≥98% of diet in the Beaufort Sea. In western Hudson Bay, ringed seals
accounted for about 80% of intake in the early 1990s, indicating the impor-
tance of foraging in ice-covered habitat. However, ringed seal consumption
declined throughout the 1990s concurrent with progressively earlier ice
breakup, while the proportions of bearded and harbour seals increased, sug-
gesting reduced reliance on ice. Throughout Davis Strait, harp seals com-
prised 50% of bears’ diets, consistent with the increase in the harp seal
population in this region. Off southern Labrador near the whelping patch,
harp seals accounted for 90% of diets. Hooded seals made up the highest
proportion of bear diets in northern Davis Strait, near their major north-
ern whelping patch. Our results demonstrate that polar bears have a high
SAMPLE COLLECTION
We obtained 321 adipose tissue samples from 295 bears (some bears were
sampled in more than one year) across three regions of the Canadian Arc-
tic, representing four subpopulations (Fig. 7.1). Subcutaneous adipose tis-
sue samples from muscle to skin were obtained from the rump either from
biopsies (6-mm biopsy punch) from anaesthetized bears – in association
with tagging and monitoring programmes of the Canadian Wildlife Ser-
vice – or from bears harvested by Inuit hunters. FA composition of subcu-
taneous adipose tissue does not differ significantly among body locations
within individual bears (Thiemann et al. in press).
Given that polar bears consume primarily the blubber of seals and that
the blubber will represent the majority of FAs stored in a seal even if
the entire carcass was consumed, we assumed that the blubber of seals
was representative of prey FAs. We obtained full-depth blubber samples
by live biopsy or from subsistence harvests from 292 individuals (across
age classes) of six species: bearded (Erignathus barbatus), harbour (Phoca
Spatial and temporal variation in the diets of polar bears 101
80°00' N
KB
NW 75°00' N
NB
n = 17
SB Amundsen LS
70°00' N
Gulf N VM BB
S
n = 17 65°00' N
MC GB
Cumberland Sound
FB n = 28
Frobisher Bay
60°00' N
n = 24
DS 55°00' N
Northern Labrador
n = 11
Churchill WHn =217 Southern
Labrador
n=7
SH
26 (a)
Prey
24
22 Bearded seal (n = 39)
20 Harbour seal (n = 7)
18
Harp seal (n = 133)
Hooded seal (n = 17)
16
Ringed seal (n = 86)
14
Walrus (n = 10)
12
10
8
Mass per cent of total fatty acids
6
4
2
0
14:0 16:0 16:1 18:0 18:1 18:1 18:2 20:1 20:1 20:4 20:5 22:1 22:1 22:5 22:6
n-7 n-9 n-7 n-6 n-11 n-9 n-6 n-3 n-11 n-9 n-3 n-3
26 (b)
24 Polar bears
22 SB (n = 17)
20 NB (n = 17)
18 WH (n = 217)
16 DS south (n = 18)
14
DS north (n = 52)
12
10
8
6
4
2
0
14:0 16:0 16:1 18:0 18:1 18:1 18:2 20:1 20:1 20:4 20:5 22:1 22:1 22:5 22:6
n-7 n-9 n-7 n-6 n-11 n-9 n-6 n-3 n-11 n-9 n-3 n-3
Fig. 7.2 Selected FAs (mean + sem; 15 of the 67 FAs identified and quantified,
representing 87% to 88% of all FAs in both seals and polar bears) that were most
abundant and/or exhibited the greatest average variance across all groups: (a) in
prey species averaged across all regions of the Canadian Arctic sampled and (b)
in polar bears within the four sampled subpopulations. Given the large range in
latitude of bears sampled in the DS population, these were separated into
northern and southern areas ( for abbreviations and areas see Fig. 7.1).
and Iverson et al. (2002). All seals and bears contained the same major
FAs, however, the FA composition varied markedly among prey species and
among bears across locations (Fig. 7.2a and b). Many of the key dietary FAs
such as 20:1n-9, 20:5n-3, 22:1n-11 and 22:6n-3, tended to characterize both
prey and bears in different regions, but means of individual FAs are a lim-
ited way to view data. Discriminant function analyses test predicted group
Spatial and temporal variation in the diets of polar bears 103
6
(a) Prey
Ringed
2nd discriminant function 4
Harbour BS
WH
WH
2 DS
WH
BS Harp
0 DSn
DS
Bearded
DSo
2 DS
Walrus
Hooded
4
6
10 8 6 4 2 0 2 4
1st discriminant function
6
(b) Bears
4
2nd discriminant function
DS
2 N
WH s
6 BS
8
6 4 2 0 2 4 6 8 10
1st discriminant function
Fig. 7.3 Results of discriminant function analyses of seals and polar bears
across the Canadian Arctic. Due to small sample sizes for some prey species (e.g.
n = 7) and bears in some locations (e.g. n = 17, see Fig. 7.2), we used a reduced
set of FAs (one minus size of the smallest group, respectively) for discriminant
analyses, to offer some assurance that covariance matrices were homogeneous
(Stevens 1986). Here we used FAs that were most abundant and/or exhibited the
greatest average variance across all groups, and analyses were performed
according to Iverson et al. (2002). (a) Plot of the group centroids (within-group
Spatial and temporal variation in the diets of polar bears 105
do not generally occur in the areas of bears sampled in the northern and
southern Beaufort (NB and SB, Fig. 7.1). Thus these bears were modelled
only on ringed and bearded seals (n = 24 prey) collected in the Beaufort
Sea. Hooded seals do not occur in WH and harp seals or walrus are avail-
able only occasionally. Nevertheless, we modelled WH bears on ringed,
bearded and harbour seals (obtained in WH), and also included harp seals
and walrus (obtained in DS, Baffin Bay (BB) and Foxe Basin (FB), Fig. 7.1;
n = 120 prey). We were not able to obtain samples of harbour seals in DS
given their rarity there, but all other species are potentially available to bears
in this area. Thus, DS bears were modelled using ringed, bearded, harp
and hooded seals and walrus collected in the DS area (50◦ 00 to 70◦ 0 N,
Fig. 7.1; n = 231 prey).
Consistent with differences in FA signatures of polar bears (Figs 7.2b
and 7.3b), their estimated diets differed dramatically across their geographic
range (Fig. 7.6). Although ringed and bearded seals occurred in bear diets
throughout the Canadian Arctic, their relative importance differed greatly.
In both NB and SB, ringed seals comprised about 95% of the FA signa-
tures of polar bear adipose tissue, with the remainder being made up by
bearded seals (Fig 7.6a). However, on average in WH, ringed seal con-
sumption accounted for only 56% of FA signatures, followed by about 38%
bearded and 5% harbour seal; trace levels (<1%) of harp seal also appeared.
In the DS subpopulation, ringed and bearded seals accounted for 18% and
26% of signatures, respectively, whereas harp seals (49%) dominated; some
hooded seals (12%) were also present. While both harp and hooded seals
were reasonably well differentiated in simulations, there was evidence for
some degree of overlap in signatures (Box 7.1; Fig. 7.3a). Thus, the precise
proportions estimated for these two species are somewhat less certain. Wal-
rus appeared at about 1% of diets in DS overall (Fig. 7.6a).
←
Fig. 7.3 (cont.) mean for each discriminant function) for the first and second
discriminant functions for prey species by region. Ellipses represent 95% of the
data points for each species. The first two functions accounted for 86.9% of the
variance among the 11 groups tested (n = 292; Wilk’s λ < 0.001). Species overall
were separated with 86.6% accuracy. For harp seals, DSn and DSo represent
individuals sampled in DS nearshore and DS offshore, respectively. (b) Plot of
the discriminant scores for each individual polar bear (n = 257, removing
repeat-sampled bears and cubs), as well as the group centroids, for the first and
second discriminant functions. NB and SB bears were combined as BS (Beaufort
Sea), due to close proximity of sampling. The first two functions accounted for
95.0% of the variance (Wilk’s λ < 0.001) and individuals were grouped to major
region (Fig. 7.1) with 99.6% accuracy overall; the only misclassifications were
between northern and southern DS.
106 S. J. Iverson et al.
100
90 WH DS
% of FA signature
80
70 a
60
50 a
40
30
20 a a a
10 a
0
a a
Ringed
Walrus
Ringed
Walrus
Harp
Harp
Hooded
Bearded
Bearded
Harbour
Fig. 7.4
set of those prey species plus all other prey in the region’s database.
This creation of the pseudo bear and subsequent modelling was then
repeated 1000 times and gave an indication of mean reliability of
estimates and noise around those estimates.
Figure 7.4 shows the results of the simulation studies for WH
and DS presented as box plots, showing the 25th, median, and 75th
percentiles of the 1000 diet estimates as horizontal bars and with
dots representing outliers. The diet composition specified is repre-
sented in plots as ‘a’ and was designated as follows. For WH, com-
position was designated as 75% ringed seal, 20% bearded seal and
5% harbour seal; for DS, it was 20% ringed seal, 20% bearded seal,
50% harp seal, 8% hooded seal and 2% walrus. The simulations
performed for each region demonstrated the reliability with which
we could differentiate prey species of polar bears in the QFASA
model. In WH, ringed, bearded and harbour seals were well esti-
mated in diets, with only trace amounts of noise from the appearance
of harp seals or walrus. When harp seal and walrus were removed
from the simulations, results were similar except that the propor-
tions of all prey items were precisely predicted as specified. In DS,
species were again well estimated with some overlap between harp
and hooded seals. Ringed seals were only slightly overestimated
and bearded seals only slightly underestimated. Results for the
Beaufort Sea (not shown) also showed precise estimates of specified
diets.
108 S. J. Iverson et al.
Calibration coefficient 4
3
2
1
0
20:1n11
16:0
18:0
14:0
16:1n7
18:1n9
18:1n7
18:2n6
20:1n9
20:4n6
22:1n11
22:6n3
20:5n3
22:1n9
22:5n3
Fatty Acids 1–67
Fig. 7.5
The above diet estimates represent the total mass contribution of each
prey species’ blubber FAs to polar bear diets. However, these species dif-
fer greatly in body size. Although by no means definitive, we used one
simple approach to illustrate the relative difference in diet estimates if one
accounts for these differences in species body mass and thus total blubber
110 S. J. Iverson et al.
90
% of Fatty acid signature
80
70
60
50
40
30
20
10
0
Ringed
Ringed
Ringed
Ringed
Bearded
Hooded
Bearded
Bearded
Bearded
Walrus
Walrus
Harp
Harp
Harbour
(b) 100
90
80
70
% no. of seals
60
50
40
30
20
10
0
Ringed
Ringed
Ringed
Ringed
Bearded
Bearded
Bearded
Hooded
Walrus
Bearded
Walrus
Harp
Harbour
Harp
Fig. 7.6 QFASA model estimates (mean + SEM) of polar bear diets across the
three major geographical regions and four subpopulations. (a) Data are
presented as the QFASA model output, or the proportional contribution of prey
species to polar bear adipose tissue FA signatures. (b) Data are presented as the
approximate percentage number of seals that would be consumed to account for
these signatures given the large size differences among prey species. For these
estimations, we made the coarse assumption that the entire carcass’s blubber
layer was consumed, that seals contained similar percentages of blubber and,
finally, that both juveniles and adults of each species scaled similarly by body
size. We used the following adult body mass values for scaling: ringed (65 kg),
bearded (300 kg), harbour (87 kg), harp (110 kg) and hooded seal (250 kg); walrus
(1040 kg); ringed seals were used as the divisor.
Spatial and temporal variation in the diets of polar bears 111
40
30
20
10
0
Females Males Females Males Females Males
Ringed seals Bearded seals Harbour seals
Fig. 7.7 QFASA estimates of female (n = 86) and male (n = 99) polar bear
diets (mean percentage of relative number of seals ± SEM; see Fig. 7.4) in WH
during a 5-year period in relation to timing of ice breakup (Stirling et al. 1999) in
those years. Only bears >1.5 years of age and sampled in summer after coming
ashore with the ice melt were included. Diets of females and males differed in all
components (p ≤ 0.01), diets decreased in levels of ringed seals and increased in
harbour seals in both females and males (p ≤ 0.01), and diets were variable in
levels of bearded seals (P = 0.236; two-way ANOVA). Results were the same
when expressed as a percentage of FA signature or as converted to a proxy of
relative number of seals (shown).
65° 90' N
Ringed Cumberland
Sound
Bearded
Harp
Hooded
Walrus
63° 10' N
Ringed Frobisher
Bearded Bay
Harp
Hooded
Walrus
58° 40' N
Ringed
Northern
Bearded Labrador
Harp
Hooded
Walrus
52° 50' N
Ringed Southern
Bearded Labrador
Harp
Hooded
Walrus
0 10 20 30 40 50 60 70 80 90 100
% no. of seals
Fig. 7.8 QFASA estimates of polar bear diets (mean percentage of the relative
number of seals + SEM; see Fig. 7.6) across the wide range of latitudes of bears
sampled within the DS population (see Fig. 7.1). Bears in southern Labrador
consumed more harp seals and fewer (none) ringed seals than in all other areas
(p < 0.01), and consumed fewer bearded seals than in all areas except
Cumberland Sound (p < 0.01). The greatest numbers of hooded seals were
consumed by bears sampled near Frobisher Bay (p < 0.05). Walrus consumption
did not differ with area (ANOVA with Fisher’s PLSD post-hoc tests).
Spatial and temporal variation in the diets of polar bears 113
The results of our study provide new insight into the foraging ecology of
polar bears across their Canadian range. Although FAs have been used
previously to make inferences about changes in lower trophic levels in an
ecosystem (Iverson et al. 1997, 2001b), this is the first time FAs have been
used to estimate quantitatively the diets of a top predator in relation to spa-
tial and temporal changes in prey availability.
Previous studies, based on direct observation, have concluded that
ringed seals are the dominant prey of polar bears in the Arctic, followed by
bearded seals (e.g. Stirling 2002). However, climate and access have lim-
ited most studies of polar bear foraging to late winter and early spring, and
to areas of landfast sea-ice and immediately adjacent pack-ice where ringed
seals are most abundant, thus potentially biasing our concept of their impor-
tance in the diet. Nevertheless, ringed seals clearly remained an important
component of polar bear diets in this study (Fig. 7.6). However, their dom-
inance (and even presence) in diets across spatial and temporal scales dif-
fered dramatically. In the bears sampled in NB and SB in the early 1990s,
ringed seals were consumed almost exclusively, consistent with their distri-
bution and abundance relative to bearded seals in those areas (e.g. Stirling
2002). Similarly, ringed seals were the dominant seal consumed in WH,
but this varied substantially by individual, sex and year (e.g. Fig. 7.7). Male
bears fed more on bearded seals than did females, which is consistent with
expectations based on their larger body size and more frequent (albeit lim-
ited) observations of male bears seen over bearded seal kills than females
114 S. J. Iverson et al.
(Stirling & Derocher 1990). This has interesting implications for evaluating
male foraging and consumption. A bearded seal is over four times heavier
than a ringed seal, and thus a bear would only need to catch a single bearded
seal to equal the blubber intake of four to five ringed seals. Future determi-
nation of age classes of seals taken will also influence these conclusions.
Despite potential sex differences in foraging tactics, both sexes responded
to environmental variability in similar ways and thus both served as indica-
tors of short-term ecosystem change.
During the 1990s in WH, the trend towards progressively earlier sea-
ice breakup dates was accompanied by significant decreases in ringed seals
in polar bear diets. Through the same period, Holst et al. (1999) and Stir-
ling (2004) documented low apparent survival of ringed seal pups in 1998–
2000. It is not clear whether the increase in the proportion of bearded and
harbour seals in the diet of WH bears through the 1990s reflects a decline
in the availability of ringed seals, an increase in the population size and
availability of bearded and harbour seals, or both. However, all species are
mainly only available to bears on the ice so these changes in diet, espe-
cially reduction in ringed seals, complement evidence that during the same
period bears came ashore earlier and in progressively poorer condition, with
a decline in both physical and reproductive characteristics (Stirling et al.
1999).
Besides short-term temporal changes, diet composition of apex preda-
tors can characterize, and signal a shift in, the abundance of lower trophic
levels. In WH, bear diets reflected their prey field as an assemblage dom-
inated by ringed and bearded seals, with increasing dependence on both
harbour and bearded seals, coincident with climate warming as predicted
by Stirling and Derocher (1993). Little or no harp seal and walrus occurred
in diets, consistent with their known geographic distributions, further sup-
porting the notion that the FAs of polar bears reflect the prey available. In
DS, bear diets reflected a longer-term trend in prey abundances. Although
it was previously known, mostly from chance encounters, that polar bears
hunted harp seals, we have for the first time shown that harp seals are the
dominant prey in this region, along with some intake of hooded seals. This
coincides with the large and well-documented increases that have occurred
in these seal populations since the early 1970s (NOAA 1999, DFO 2000)
and with apparently increasing polar bear numbers observed in this area
(I. Stirling, unpublished data, 2005). Harp seals were most abundant in
the diets of bears in southern Labrador, which is closest to the harp seal
whelping patch, while hooded seals were most abundant in diets of bears
near Frobisher Bay (63◦ 10 N), which is closest to the northern hooded
seal whelping patch (62◦ to 64◦ N). These results are consistent with both
Spatial and temporal variation in the diets of polar bears 115
ACKNOWLEDGEMENTS
We thank the following for their support of the research that made this
project possible: the Natural Sciences and Engineering Research Council
(NSERC) Canada; the Canadian Wildlife Service (CWS); the Polar Continen-
tal Shelf Project; the World Wildlife Fund and the Churchill Northern Stud-
ies Centre. We are particularly grateful to the many people from the North-
west Territories Department of Resources, Wildlife, and Economic Devel-
opment; the Nunavut Department of Sustainable Development (especially
F. Piugattuk and J. Beauchesne); the Labrador Inuit Association; the Inu-
vialuit Fisheries Joint Management Committee and the CWS for collecting
fat samples from subsistence harvesting programmes and ongoing research
projects on both polar bears and seals. We also thank G. Stenson for pro-
viding most of the harp and hooded seal samples. J. Lasner and C. Beck
assisted with some sample analyses and W. Blanchard assisted with the
QFASA modelling. We thank W. D. Bowen, I. L. Boyd and A. E. Derocher
for providing helpful comments on an earlier version of the manuscript.
REFERENCES
Amstrup, S. C., Durner, G. M., Stirling, I., Lunn, N. & Messier, F. (2000).
Movements and distribution of polar bears in the Beaufort Sea. Can. J. Zool.,
78, 948–66.
116 S. J. Iverson et al.
define thermal habitats and ‘hotspots’ used by predators and prey. Regional
and global comparisons aid in understanding changing meso-scale pro-
cesses and ocean-scale patterns for effective ocean conservation and
management.
Top-predator responses to changes in prey and environmental con-
ditions pose pressing questions about food webs, ecosystem dynamics,
climate change and conservation. In marine systems, physical processes
impose more pervasive influences on animals than in fresh water. Physi-
cal events as well as biological interactions determine the distributions and
abundance of exothermic animals which in turn drive the distributions and
activities of their endothermic predators (Springer et al. 1984).
Owing to mobility, high metabolic rates and limited lipid storage,
many seabirds maintain relatively close spatial and temporal associations
with prey (Russell et al. 1992, Davoren et al. 2002). Comparisons of
short- and long-term behavioural, dietary, distributional, production and
population responses of surface-feeding and diving seabirds that feed at
different trophic levels reflect changes in prey and ecosystem conditions
(Montevecchi 1993, Springer & Speckman 1997). Surface-feeders often
have more extensive foraging ranges and are more sensitive to surface-
water effects on prey distributions than divers. Feeding at lower trophic lev-
els, planktivores are often more affected by oceanographic fluctuations than
piscivores.
Seabird–prey interactions are most spatially constrained, yet highly
energy-demanding, during breeding seasons when birds forage from land-
based colonies (fixed-place foraging). During these periods, birds are highly
attuned to prey conditions and often develop traditional colony-specific for-
aging areas (e.g. Cairns 1989, Grémillet et al. 2004). Over decades and cen-
turies, colonies tend to be in areas of predictable prey availability and, conse-
quently, changes in breeding distributions and populations can be related to
long-term oceanographic change (Montevecchi & Myers 1997, Ainley 2002,
Emslie et al. 2003).
Seabird foraging ecology has been studied indirectly by sampling the
diets of chicks and adults at colonies and by surveying distributions and den-
sities from vessels. Diet samples at colonies reveal what birds have preyed
on but not where it came from or its distribution or relative abundance
compared with other prey. Surveys at sea associate avian occurrences with
oceanography and prey fields. Synoptic integration of these approaches is
significantly more informative about where predators captured prey and
options that they might have had in the process. Yet, to determine mech-
anisms of predator responses, direct studies of the foraging tactics of
120 W. A. Montevecchi et al.
INDICATOR RESPONSES
Multispecies assessments
Order of magnitude
Index
species Mass (g) Population Max. dive Trophic level
avian predators in the region and were selected to maximize the efficient
assessment of prey and environmental conditions using top predators.
REGIME SHIFTS
(a)
Gulls Surface feeders
Colony
Puffins Murres
Capelin
(b) Colony
Puffins Murres
Capelin
Colony
(c)
Puffins Murres
Capelin
Fig. 8.2 Circumstances in which forage fish are: (a) available to surface-feeders
(arrows represent horizontal ranges) and divers (vertical ranges are indicated by
the position of the ellipsoid beneath the sea level); (b) unavailable to surface-
feeders due to a change in vertical distribution; (c) available to surface-feeders
and divers despite a change in abundance represented by the size of the dashed
ellipsoids.
Cury & Shannon 2004). Definitions of regime shift converge on a few com-
monalities, i.e. rapid onset, physically forced, ‘bottom-up’ perturbations that
initiate decadal or longer systemic food-web and ecological changes (Steele
1996, 1998, Hare & Mantua 2000). Shifts can also be forced from the ‘top’
by over-fishing, and ecosystems with food webs dominated by mid-trophic-
level forage fishes are especially sensitive to such effects (Cury et al. 2000).
The persistence of regime shifts is due in part to lagged biological responses
to physical perturbation. Following a regime shift, a system must be ‘forced’
again to change states, being unlikely to return to its former condition.
124 W. A. Montevecchi et al.
100
Cold
80
Percentage of landings
Warm
60
40
20
0
1978 1983 1988 1993 1998 2003
Year
Fig. 8.3 Shift from landings of warm-water prey (e.g. mackerel, short-finned
squid) to cold-water prey (capelin, Atlantic salmon) by breeding northern
gannets at Funk Island in the northwest Atlantic.
sprats in the Baltic Sea; capelin in the Barents Sea and northwest Atlantic;
pollack in Alaskan waters; krill in the Antarctic. Biophysical influences on
forage species generate systemic changes that permeate pelagic food webs
(Anderson & Piatt 1999, Cury et al. 2000) in the Baltic Sea. Unfortunately,
forage species are often not well studied, either because they are not com-
mercially pursued or because they are undervalued as a raw material by
industrial fisheries (Aikman 1997). Seabird research involving focal for-
age species provides insight and predictive capability about their population
resiliency and about their roles in potential trophic shifts (Hatch & Sanger
1992, Springer & Speckman 1997, Barrett 2002, Davoren & Montevecchi
2003, Miller & Sydeman 2004, Wanless et al. 2004).
THERMAL HABITATS
‘hotspots’, especially when they are within marine protected areas (Schmidt
1997), will help provide the archival oceanographic information required
to place environmental changes in an ecological context and will facilitate
effective conservation and management.
ACKNOWLEDGEMENTS
REFERENCES
Aikman, P. (1997). Industrial ‘Hoover’ Fishing: A Policy Vacuum. Amsterdam, the
Netherlands: Greenpeace.
Ainley, D. (2002). Adélie Penguin: Bellweather of Climate Change. New York:
Columbia University Press.
Ancel, A., Kooyman, G. L., Ponganis, P. J. et al. (1992). Foraging behaviour of
emperor penguins as a resource detector. Nature, 360, 336–9.
Anderson, P. J. & Piatt, J. F. (1999). Community reorganization in the Gulf of
Alaska following ocean climate regime shift. Mar. Ecol. Prog. Ser., 189, 117–23.
Baird, P. H. (1990). Influence of abiotic factors and prey distribution on diet and
reproductive success of three seabird speices in Alaska. Ornis Scand., 21,
224–35.
Barrett, R. T. (1996). Prey harvest, chick growth, and production of three seabird
species on Bleiksøy, North Norway during years of variable food availability. In
Studies of High-Latitude Seabirds. 4. Trophic Relationships and Energetics of
Endotherms in Cold Ocean Systems, ed. W. A. Montevecchi. Canadian Wildlife
Service Occasional Paper 91. Ottowa, Ontario, Canada: Canadian Wildlife
Service, pp. 20–26.
(2002). Puffin and guillemot chick food and growth as indicators of fish stocks
in the Barents Sea. Mar. Ecol. Prog. Ser., 230, 275–87.
Benvenuti, S., Bonadonna, F., Dall’Antonia, L. & Gudmundsson, G. A. (1998).
Foraging flights of breeding thick-billed murres carrying direction recorders.
Auk, 115, 57–66.
Birt-Friesen, V., Montevecchi, W. A., Cairns, D. K. & Macko, S. A. (1989). Activity
specific metabolic rates of free living seabirds with emphasis on Northern
Gannets Sula bassanus. Ecology, 70, 357–67.
Boyd, I. L. (1996). Temporal scales of foraging in a marine predator. Ecology, 77,
426–34.
Boyd, I. L. & Murray, W. (2001). Monitoring marine ecosystems with responses of
upper trophic level predators. J. Anim. Ecol., 70, 747–60.
128 W. A. Montevecchi et al.
Jones, I. L., Hunter, F. M. & Robertson, G. J. (2002). Annual adult survival of Least
Auklets (Aves, Alcidae) varies with large-scale climatic conditions of the North
Pacific Ocean. Oecologia, 133, 38–44.
Jouventin, J. & Weimerskirch, H. (1990). Satellite tracking of wandering
albatrosses. Nature, 343, 746–8.
Kitaysky, A. S. & Golubova, E. G. (2000). Climate change causes contrasting
trends in reproductive performance of planktivorous and piscivorous alcids.
J. Anim. Ecol., 69, 248–69.
May, R., Beddington, J. R., Clark, C. W., Holt, S. J. & Laws, R. M. (1979).
Management of multispecies fisheries. Science, 205, 267–77.
Miller, A. K. & Sydeman, W. J. (2004). Rockfish response to low-frequency
ocean climate change as revealed by the diet of a marine bird over multiple
time scales. Mar. Ecol. Prog. Ser., 281, 207–16.
Montevecchi, W. A. (1993). Birds as indicators of change in marine prey stocks. In
Birds as Monitors of Environmental Change, eds. R. W. Furness & J. Greenwood.
London: Chapman and Hall, pp. 217–66.
(ed.) (1996). Studies of High-Latitude Seabirds. 4. Trophic Relationships and
Energetics of Endotherms in Cold Ocean Systems. Canadian Wildlife Service
Occasional Paper 91. Ontario, Canada: Canadian Wildlife Service.
(2001). Interactions between fisheries and seabirds. In Biology of Marine Birds,
eds. E. A. Schrieber & J. Burger. Boca Raton, FL: CRC Press, pp. 527–57.
Montevecchi, W. A. & Berruti, A. (1991). Avian indication of pelagic fishery
conditions in the southeast and northwest Atlantic. Acta Int. Ornithol. Cong.,
20, 2246–56.
Montevecchi, W. A. & Myers, R. A. (1995). Seabird prey harvests reflect pelagic fish
and squid abundance on multiple spatial and temporal scales. Mar. Ecol. Prog.
Ser., 117, 1–9.
(1996). Dietary changes of seabirds reflect shifts in pelagic food webs. Sarsia, 80,
313–22.
(1997). Centurial and decadal oceanographic influences on changes in northern
gannet populations and diets in the Northwest Atlantic: implications for
climate change. ICES J. Mar. Sci., 54, 608–14.
Mowbray, F. (2003). Changes in the vertical distribution of capelin off
Newfoundland. ICES J. Mar. Sci., 59, 942–9.
Regehr, H. & Montevecchi, W. A. (1997). Effects of food shortage and predation on
breeding failure of kittiwakes. Mar. Ecol. Prog. Ser., 155, 249–60.
Reid, K. & Croxall, J. P. (2000). Environmental response of upper trophic-level
predators reveals system change in an Antarctic ecosystem. Proc. R. Soc. Lond.
B, 268, 377–84.
Russell, R., Hunt, G. L., Coyle, K. & Cooney, K. (1992). Foraging in a fractal
environment: spatial patterns in a marine predator–prey system. Landscape
Ecol., 7, 195–209.
Scheffer, M., Carpenter, S., Foley, J. A., Folke, C. & Walker, B. (2001). Catastrophic
shifts in ecosystems. Nature, 413, 591–6.
Schmidt, K. (1997). ‘No take’ zones spark fisheries debate. Science, 277, 489–91.
Skov, H., Durinck, J. & Andell, P. (2000). Associations between wintering avian
predators and schooling fish. J. Avian Biol., 31, 135–43.
130 W. A. Montevecchi et al.
During the summer months at South Georgia the diet of female Antarc-
tic fur seals is dominated (88% frequency of occurrence) by krill (Reid
& Arnould 1996). With such a reliance on one prey species, changes in
the abundance of krill influences fur seal foraging (Mori & Boyd 2004)
and breeding success (Reid & Croxall 2001) at South Georgia. In years
of low krill availability females must increase their foraging effort, spend-
ing around 40% longer at sea and 25% less time ashore (McCafferty et al.
1998).
Monitoring basic behavioural responses, such as trip duration, can
therefore potentially be used as a measure of krill availability (CCAMLR
1995). However, trip duration is highly variable between individual seals,
even for those operating in the same environmental conditions. Recent
work has shown that there is a strong individual component to their
behaviour, with some seals consistently undertaking longer trips than oth-
ers (Staniland et al. 2004). Fur seals are able to adjust their behaviour at a
number of scales (not just overall trip length) in order to maximize their
rate of energy gain (Boyd 1996, Mori & Boyd 2004). As a result, they
are able to reproduce successfully under a wide range of environmental
conditions.
Prey distribution and top-predator foraging behaviour 133
Days Dives Per trip Per dive No. of krill per divea
a Based on the consumption of gravid female krill (60-mm long and containing 5.45
kJ g−1 ).
Satellite telemetry data have allowed the important foraging areas of top
predators to be described (Fig. 9.1). Clearly concentrations of foraging
predators are likely to correspond to areas of high prey availability. Indeed
Prey distribution and top-predator foraging behaviour 135
(a)
53° S
54° S
55° S
Relative time
(b)
53° S
54° S
55° S
Relative time
Fig. 9.1 Density distribution maps for (a) Antarctic fur seals and (b) macaroni
penguins foraging from Bird Island, South Georgia. The density distributions
are contour plots using linear spline interpolation based on the relative time
spent within each 0.1 km × 0.1 km area. Fur seal data represent 31 animals
throughout the breeding season, from December until March. Macaroni penguin
data represent 46 individual trips during January.
such data have been used to define small-scale management units for
Antarctic fisheries (CCAMLR, 2002).
Work on fur seal foraging over the last 20 years has given an insight into the
distribution of krill in the water column. Krill swarms, like many other zoo-
plankton, undergo diel vertical migration that is known to vary depending
136 I. J. Staniland et al.
Using simple models, it is evident that fur seals need to exploit krill at higher
densities than macaroni penguins (Box 9.1). This is further highlighted by
the contrast in the organization of diving in the two species.
Prey distribution and top-predator foraging behaviour 137
(a)
0
0
Depth (m)
50
100
00:00 00:00 00:00 00:00
(b)
0
Depth (m)
50
100
17:00 19:00 21:00
Local time
Fig. 9.2 Examples of typical dive records from an individual Antarctic fur seal
(a and b) and macaroni penguin (c and d). Organization of dives is shown at the
trip level and expanded to show four hours in detail. The unfilled and filled
rectangles represent the day–night cycle. Barred lines show dives grouped into
bouts separated by inter-bout periods (curled brackets). Dives marked with a
cross were not included in bouts.
138 I. J. Staniland et al.
(c)
0
Depth (m)
50
100
00:00 00:00
(d)
0
Depth (m)
50
100
11:00 01:00 03:00
Local time
water made longer, deeper dives and spent longer at the bottom of the dive
than those foraging in deep water.
CONCLUSION
allow us to measure the behaviour of these predators while they are for-
aging. This can provide an insight into the dispersion and patchiness of
prey populations in a way that would otherwise be impossible or at least
impractical. However, perhaps the biggest challenge is to condense the
results of such analyses to a form that is useful for management purposes.
In this context, there is a growing recognition of the need for functional
response relationships between the environment and marine predators,
and for an understanding of the behavioural mechanisms that lead to par-
ticular response functions (Mori & Boyd 2004). These are critical to using
predator information in the monitoring and management of marine ecosys-
tems. By understanding their behaviour when foraging, we can make pre-
dictions about critical prey densities at which predators are unable to main-
tain their energy intake. Thus behavioural data are critical for any under-
standing of how functional response curves relate the abundance of prey to
predator performance indices.
REFERENCES
Dall, S. R. X. & Boyd, I. L. (2002). Provisioning under the risk of starvation. Evol.
Ecol. Res., 4, 883–96.
Godlewska, M. (1996). Vertical migrations of krill (Euphausia superba Dana). Pol.
Arch. Hydrobiol., 43, 9–63.
Green, J. A., Butler, P. J., Woakes, A. J. & Boyd, I. L. (2002). Energy requirements
of female Macaroni penguins breeding at South Georgia. Funct. Ecol., 16,
671–81.
(2003). Energetics of diving in Macaroni penguins. J. Exp. Biol., 206, 43–57.
McCafferty, D. J., Boyd, I. L., Walker, T. R. & Taylor, R. I. (1998). Foraging
responses of Antarctic fur seals to changes in the marine environment. Mar.
Ecol. Prog. Ser., 166, 285–99.
Miller, D. G. M. & Hampton, I. (1989). Krill aggregation characteristics: spatial
distribution patterns from hydroacoustic observations. Polar Biol., 10, 125–34.
Miller, D. G. M., Barange, M., Klindt, H.et al. (1993). Antarctic krill aggregation
characteristics from acoustic observations in the South West Atlantic Ocean.
Mar. Biol., 117, 171–83.
Mori, Y. & Boyd, I. L. (2004). The behavioral basis for nonlinear functional
responses and optimal foraging in Antarctic fur seals. Ecology, 85, 398–410.
Reid, K. & Arnould, J. P. Y. (1996). The diet of Antarctic fur seals Arctocephalus
gazella during the breeding season at South Georgia. Polar Biol., 16, 105–14.
Reid, K. & Croxall, J. P. (2001). Environmental response of upper trophic-level
predators reveals a system change in an Antarctic marine ecosystem. Proc. R.
Soc. Lond. B, 268: 377–84.
Staniland, I. J. & Boyd, I. L. (2003). Variation in the foraging location of Antarctic
fur seals (Arctocephalus gazella), the effects on diving behaviour. Mar. Mamm.
Sci., 19, 331–43.
Staniland, I. J., Boyd, I. L. & Reid, K. (2004). Comparing individual and spatial
influences on foraging behaviour in Antarctic fur seals. Mar. Ecol. Prog. Ser.,
275, 263–74.
Trathan, P. N., Croxall, J. P., Murphy, E. J. & Everson, I. (1998). Use of at-sea data
to derive potential foraging ranges of Macaroni penguins during the breeding
season. Mar. Ecol. Prog. Ser., 169, 263–75.
10
Identifying drivers of change: did fisheries play a
role in the spread of North Atlantic fulmars?
P. M. THOMPSON
(a) (b)
1800
1800–99 100000
1900–49
Number of nests
10000
1950–2000
1000
100
(c)
100000
Number of nests
10000
1000
100
Fig. 10.1 Changes in the distribution and abundance of North Atlantic fulmars.
(a) Schematic distribution of boreal populations, illustrating the timing of their
range expansion. (b) Changes in the number of apparently occupied nest sites at
newly colonized breeding sites in Britain and Ireland, excluding St. Kilda. Data
from Fisher and Waterston (1941), Fisher (1966), Cramp et al. (1974) and
Mitchell et al. (2004). (c) Changes in the number of apparently occupied nest
sites at all breeding sites in Britain and Ireland, including St Kilda. Data sources
as for Fig. 10.1b, with an assumed stable population of 21 000 on St Kilda
between 1875 and 1939 (see Fisher and Waterston (1941)).
Hatch 1991, Thompson & Ollason 2001). Throughout the expansion, there
has been intense debate about the underlying causes of these changes,
particularly in relation to the potential influence of commercial fisheries
(Box 10.2). As such, this provides an interesting case study where inferences
can be integrated from both broad-scale and fine-scale ecological studies of
this species. More recently, fulmar populations in European waters have
stabilized (Mitchell et al. 2004), and investigations of earlier trends may
provide useful insights into current management issues involving both ful-
mars and other marine top predators.
It has been strongly argued, and is now widely believed, that the spread
of fulmars was driven largely by increases in food availability from whal-
ing and commercial fisheries (Fisher 1952). At the time, little was known
about fulmar diet and feeding ecology but, over the last decade, there has
been more research in this area. Fulmars often scavenge around fishing
boats, but wider-scale studies indicate that their distribution is more closely
related to hydrographic features than to fisheries (Camphuysen & Garthe
Identifying drivers of change 149
1997, Skov & Durinck 2000). Similarly, comparisons of diet across their
range suggest that fishery-derived offal and discards can form an important
part of the diet in some areas, but that birds at many other colonies tend to
forage on pelagic crustaceans and small fish (Furness & Todd 1984, Hamer
et al. 1997, Phillips et al. 1999). Nevertheless, while these prey appear to be
taken directly, their availability may also have increased indirectly as a result
of fishing pressure (Pauly et al. 1998). More generally, this work has shown
that fulmars are extremely catholic in their diet, and that diet composition
can differ markedly between years at a single site (Phillips et al. 1999). Such
inter-annual variation makes assessments of longer-term trends difficult to
evaluate, and further highlights that seasonal variation in diet is likely. The
predominance of breeding-season studies may therefore bias our under-
standing of the overall diet of these birds.
The development of bio-energetic models can also be used to assess
whether current levels of discarding could support the energetic require-
ments of different seabird populations. Even in heavily fished areas such as
the North Sea, fewer than 50% of fulmars could be fully supported by fishery
waste (Camphuysen & Garthe 1997). Together, these studies suggest that
fulmars are not, at least currently, heavily dependent upon fishery waste.
Nevertheless, the availability of discards may be important to these birds at
times when, or in areas where, natural prey are more limited (Mitchell et al.
2004); probably during the winter (when natural prey are less available),
and during the early chick-rearing period (when adults are constrained to
shorter foraging trips).
(a) (c) 9
8
7
% recruiting
6
5
4
3
2
1
0
0.2 0.1 0.0 0.1
Temperature anomaly (°C)
(b) (d)
60 9
8
Reproductive success (%)
50 7
% recruiting
6
40 5
4
30
3
2
20
1
10 0
6 4 2 0 2 4 6 10 20 30 40 50 60
Winter North Atlantic Oscillation Annual reproductive success
Fig. 10.2 Studies of individually marked fulmars on Eynhallow (a) have shown
that reproductive success and recruitment are influenced by climate variation. (b)
Reproductive success is negatively related to the winter North Atlantic Oscillation
index. (c) Cohort recruitment rates are positively related to northern-hemisphere
temperature anomalies. (d) Cohorts experiencing higher reproductive success do
not necessarily exhibit higher levels of recruitment Annual breeding success is
calculated as the percentage of eggs laid that produce successful fledglings.
Panels b and c are redrawn from Thompson and Ollason (2001).
are not involved. Indeed, these analyses show that there remains a strong,
unexplained, linear increase in colony size, which earlier work suggests
must have been driven by immigration from other areas (Ollason & Dunnet
1983).
One of the key constraints when interpreting recent data on diet composi-
tion is that these patterns may not be representative of earlier critical periods
during the fulmar’s expansion. Even the retrospective analysis of long-term
individual-based data is restricted to the later phases of the expansion, and
there is limited information on relevant environmental covariates. Attempts
to understand historical patterns of ecological changes have therefore often
drawn upon a wide variety of indirect proxies of abundance, diet or environ-
mental changes that may have driven such changes. In other systems, these
have included economic records of catches (Allen & Keay 2001), analyses
of hair of seals and fish scales in seabed sediments (Hodgson et al. 1998,
O’Connell & Tunnicliffe 2001) and an increasing array of molecular and
biochemical techniques for understanding variation in abundance (Roman
& Palumbi, 2003), dispersal and feeding ecology (Hobson 1993, 1999,
Smith et al. 1997). Such studies provide useful new insights into the nature
of the fulmar’s spread, and highlight the potential for similar approaches to
extend these findings in the future.
The first of these insights involve paleoecological studies, where
Montevecchi and Hufthammer (1990) described the distribution of ful-
mar bones from archaeological sites in Norway. They found evidence
of fulmars at 26 sites, extending through northern and southern Nor-
way. Most recovered bones were dated at between 1000 and 4000 years
before present, with a peak in the period 1000 to 2000 years ago. The
recent colonization of Norway in 1920 therefore appears to have been
a re-colonization, and clearly shows that recent changes in distribution
did not necessarily depend upon human-induced changes in food sup-
plies. Records of fulmars at Scottish archaeological sites are also scattered
through the literature. Fisher and Waterston (1941) briefly mention, but
then ignore, the fact that fulmar bones were found in a midden from
the west of Scotland; more recent studies have recorded fulmars from
excavations dating from the Neolithic to the early medieval (Serjeantson
1988). Further review of archaeological data from other parts of the ful-
mar’s contemporary range, together with carbon dating of the Scottish spec-
imens, would provide valuable insights into their historical distribution.
152 P. M. Thompson
While fisheries waste has proved an important source of food for fulmars
in certain areas, it is clearly not the sole cause underlying the population’s
expansion. Instead it seems much more likely that the population was
responding to multiple driving forces, leaving us with the challenge of deter-
mining the relative importance of different drivers. While this is beyond the
scope of this chapter, examination of the process by which explanations for
the spread were developed and interpreted provides some general lessons
for evaluating this and similar issues in the future.
One clear feature of criticisms of alternative hypotheses was that they
were often constrained by the limited information available on seabird life-
history patterns at that time. Perhaps the most obvious shortcoming is
that complex descriptions of the spatial pattern of spread were based on
the assumption that successful breeders produced young that recruited to
nearby colonies at just 1 or 2 years old (Fisher 1952). Similarly, lack of knowl-
edge about age at maturity and longevity meant that the sensitivity of popu-
lation trends to variations in reproduction and adult survival were not fully
appreciated (Fisher & Waterston 1941). There remains uncertainty over the
detailed ecology of many species involved in current fisheries interactions,
Identifying drivers of change 153
and we should be careful not to dismiss hypotheses simply because they lie
outside our current understanding of the system.
Secondly, issues of scale heavily influenced perceptions of the impor-
tance of fisheries. At a fine scale, large feeding flocks of fulmars around
whale carcasses and trawlers appeared to provide convincing support for
the importance of these artificial prey supplies. Only since larger-scale stud-
ies have been conducted has it become clear that attraction to vessels is a
relatively local process (Skov & Durinck 2000). Indeed, our perception of
the rapid rise in abundance is itself biased because most studies have been
carried out at a local scale in newly colonized parts of the United King-
dom. Broader-scale assessments of population increases are less dramatic
(see Box 10.1), and may not be so very different from those observed in
other seabirds in the region (Mitchell et al. 2004). Another aspect of scale
that links to our understanding of life history, is that driving forces acting
upon reproduction or early survival will have lagged effects on measures of
population abundance. Specifically, given the low power of many marine
monitoring programmes, this means that we may need to look well back
into the past to identify drivers of recently detected changes.
The lack of opportunities to test alternative hypotheses directly appears
to have encouraged the champions of different hypotheses to become
increasingly entrenched in their opinions (Box 10.2). The dismissal of some
hypotheses now seems premature, as several arguments used against these
ideas are much less convincing in the light of current ecological understand-
ing. Fisher and Waterston’s (1941) assessment of the impact of hunting on
populations in both Iceland and St Kilda was based on the assumption that
only young birds were taken. However, other sources indicate that harvests
were of both adults and young, and there would also be additional losses
from egg harvesting. When Martin Martin visited St Kilda in 1697, he esti-
mated that the people of St Kilda had given their party 16 000 seabird eggs
during their stay; while his description of the delicate taste of the adult ful-
mars highlights the existence of a mixed harvest (cited in Gordon (1936)).
Given the sensitivity of populations of long-lived vertebrates to changes in
adult mortality, slight changes in the ratio of adults to young in reported har-
vest levels (for example in response to a known decline in the market for the
oil and down during the late nineteenth century (Harman 1997)) could have
had important impacts on population growth rates. Alternative hypotheses
clearly need to be kept under review; particularly where there may be oppor-
tunities to develop new techniques to explore some of these old questions.
Even if changes in top-down processes did not influence the early stages
of the expansion, reductions in hunting after the 1930s must have con-
tributed to the faster increases during the next few decades (see Box 10.1).
154 P. M. Thompson
REFERENCES
Allen, R. C. & Keay, I. (2001). The first great whale extinction: the end of the
bowhead whale in the eastern arctic. Explor. Econ. Hist., 38, 448–77.
Barbraud, C., Weimerskirch, H., Guinet, C. & Jouventin, P. (2000). Effect of
sea-ice extent on adult survival of an Antarctic top predator: the snow petrel
Pagodroma nivea. Oecologia, 125, 483–8.
Beaugrand, G., Reid, P. C., Ibanez, F., Lindley, J. A. & Edwards, M. (2002).
Reorganization of North Atlantic marine copepod biodiversity and climate.
Science, 296, 1692–4.
Brown, R. G. B. (1970). Fulmar distribution: a Canadian perspective. Ibis, 112,
44–51.
Burg, T. M., Lomax, J., Almond, R., Brooke, M. D. & Amos, W. (2003). Unravelling
dispersal patterns in an expanding population of a highly mobile seabird, the
northern fulmar (Fulmarus glacialis). Proc. R. Soc. Lond. B Biol. Sci., 270,
979–84.
Camphuysen, C. J. & Garthe, S. (1997). An evaluation of the distribution and
scavenging habits of northern fulmars (Fulmarus glacialis) in the North Sea.
ICES J. Mar. Sci., 54, 654–83.
Cramp, S., Bourne, W. & Saunders, D. (1974). The seabirds of Britain and Ireland.
London: Collins.
Dunn, E. & Steel, C. (2001). The Impact of Longline Fishing on Seabirds in the
North-east Atlantic: Recommendations for Reducing Mortality. Report of the
RSPB/NOF/JNCC/BirdLife International, Report No. 5. Sandy, UK: RSPB.
Identifying drivers of change 155
Ollason, J. & Dunnet, G. (1978). Age, experience and other factors affecting the
breeding success of the Fulmar, Fulmarus glacialis in Orkney. J. Anim. Ecol., 47,
961–76.
(1983). Modelling annual changes in numbers of breeding Fulmars, Fulmarus
glacialis, at a colony in Orkney. J. Anim. Ecol., 52, 185–98.
Pauly, D., Christensen, V., Dalsgaard, J., Froese, R. & Torres, F. (1998). Fishing
down marine food webs. Science, 279, 860–3.
Phillips, R. A., Petersen, M. K., Lilliendahl, K. et al. (1999). Diet of the northern
fulmar Fulmarus glacialis: reliance on commercial fisheries? Mar. Biol., 135,
159–70.
Rindorf, A., Wanless, S. & Harris, M. P. (2000). Effects of changes in sandeel
availability on the reproductive output of seabirds. Mar. Ecol. Prog. Ser., 202,
241–52.
Roman, J. & Palumbi, S. R. (2003). Whales before whaling in the North Atlantic.
Science, 301, 508–10.
Salomonsen, F. (1965). The geographical variation of the fulmar (Fulmarus glacialis)
and the zones of marine environment in the North Atlantic. Auk, 82, 327–55.
Serjeantson, D. (1988). Archaeological and ethnographic evidence for seabird
exploitation in Scotland. Archaeozoologia, 11, 209–24.
Skov, H. & Durinck, J. (2000). Seabird distribution in relation to hydrography in
the Skagerrak. Cont. Shelf Res., 20, 169–87.
Smith, S. J., Iverson, S. J. & Bowen, W. D. (1997). Fatty acid signatures and
classification trees: new tools for investigating the foraging ecology of seals.
Can. J. Fish. Aquat. Sci., 54, 1377–86.
Springer, A. M., Estes, J. A., Vliet, G. B. van et al. (2003). Sequential megafaunal
collapse in the North Pacific Ocean: an ongoing legacy of industrial whaling?
Proc. Natl. Acad. Sci. U. S. A., 100, 12 223–8.
Stenhouse, I. J. & Montevecchi, W. A. (1999). Increasing and expanding
populations of breeding Northern Fulmars in Atlantic Canada. Waterbirds, 22,
382–91.
Thompson, D. R., Furness, R. W. & Lewis, S. A. (1995). Diets and long-term
changes in delta-N-15 and delta-C-13 values in Northern Fulmars Fulmarus
glacialis from two Northeast Atlantic colonies. Mar. Ecol. Prog. Ser., 125, 3–11.
Thompson, P. M. & Ollason, J. C. (2001). Lagged effects of ocean-climate change
on fulmar population dynamics. Nature, 413, 417–20.
Trites, A. W. & Donnelly, C. P. (2003). The decline of Steller sea lions Eumetopias
jubatus in Alaska: a review of the nutritional stress hypothesis. Mamm. Rev., 33,
3–28.
Wooller, R. D., Bradley, J. S. & Croxall, J. P. (1992). Long-term population studies
of seabirds. Trends Ecol. Evol., 7, 111–14.
Wynne-Edwards, V. (1962). Animal Dispersion in Relation to Social Behaviour.
Edinburgh, UK: Oliver and Boyd.
11
Monitoring predator–prey interactions using
multiple predator species: the
South Georgia experience
J . P . C R O XA L L
These principles (in effect pioneering the potential practical use of the pre-
cautionary principle) and the explicit need to balance resource exploita-
tion and conservation within a single management regime, led to the first
attempts for marine systems explicitly to try to develop approaches to man-
agement on an ecosystem basis. This was developed through two main
approaches: creating an appropriate conceptual framework (e.g. as illus-
trated in CCAMLR (1995) and Fig. 11.3 later in this chapter) and establishing
a CCAMLR Ecosystem Monitoring Programme (CEMP). The CEMP was
designed mainly to:
(a) detect significant changes in critical components of the ecosystem;
(b) distinguish between changes due to harvesting and changes due to
environmental variability.
By 1985, when CEMP planning started, the principles and practices
established in the BAS monitoring and population programmes at Bird
Island had been expanded and endorsed by the Scientific Committee for
Antarctic Research (SCAR 1979) as an appropriate model for an interna-
tional network and scheme. They also formed the foundation for the devel-
opment of the CEMP (CCAMLR 1985, 1986, Croxall et al. 1988, Agnew
1997, Croxall & Nicol 2004) which started in 1987. The focus of this moni-
toring programme was on species dependent on krill; criteria for the selec-
tion of monitored species and variables are summarized in Box 11.1.
Box 11.1 Choice of monitoring species and variables
The criteria used to select species for the predator-based monitoring
at Bird Island (and CEMP) were:
r significance (important consumer of krill);
r specialist (krill consistently forming the main element of the
diet);
r widely distributed (a range of additional study sites potentially
available);
r feasibility (readily accessible breeding sites and tolerant of
human activity).
The resulting species chosen at Bird Island were gentoo and mac-
aroni penguins, black-browed albatross and Antarctic fur seal. (Addi-
tional species for CEMP were Adélie and chinstrap penguins, Cape
and Antarctic petrels and crabeater seal).
The variables selected were designed to cover a range of spatial
and temporal scales (Fig. 11.1). Inevitably the majority of these reflect
performance and conditions during the austral summer breeding
season, whereas only arrival mass (and breeding-population size, at
least in part) reflect conditions in winter (Fig. 11.1a). The variables
were also selected on the basis of ease and accuracy of measurement
(given that detecting change over relatively short periods requires
repeated acquisition of large samples of data). A handbook provid-
ing standard methods for measuring each variable was produced for
CEMP (CCAMLR 1987) and updated annually.
Most variables integrate processes and conditions operating on
scales of months (i.e. the duration of offspring-rearing events);
although the constituent elements of some of these variables (e.g.
foraging-trip duration, diet) are collected, and can reveal effects at,
smaller scales. Population variables represent variation at scales of
at least one year, with population size – which integrates annual and
multi-year effects – being particularly complex to interpret.
The spatial scales at which these variables integrate prey
and environmental conditions are particularly diverse (Fig. 11.1b),
ranging from 10 to 100 km in gentoo penguin to over 10 000 km in
black-browed albatross. Knowledge and awareness of the relation-
ship between temporal and spatial scales is often crucial to interpret-
ing monitoring data in terms of influence of prey and environment.
Monitoring and linking these at congruent scales to the particular
predator variables is obviously essential. The advent of satellite track-
ing has revolutionized our ability to understand the nature of forag-
ing range (and key feeding areas within these ranges) of predators.
160 J. P. Croxall
Breeding population
Chicks fledging
Adult arrival mass Foraging-trip duration
Growth rate
Incubation shift duration
Adult departure mass
Breeding success Chick fledging mass
(b)
Multi-year
n
tio
p ula
Po size
Year
ult l
Ad viva
r
su
Time
Months
g
din
r ee cess
B uc
s
g
rin g
fsp ionin
Weeks
f
O is
v
ing r pro
o rag viou
F ha
be
Black-browed
albatross 10 100 1000 10 000
Gentoo
penguin 1 10 100
Space (km)
Fig. 11.1 Examples of temporal and spatial scales of relevance to variables
selected for monitoring by the Bird Island and CCAMLR Ecosystem Monitoring
Programmes. (a) Timing and duration of various parameters for macaroni
penguin. (b) Temporal and spatial scales integrated by main categories of
monitoring variables for the two most extreme species at South Georgia:
black-browed albatross (upper spatial scale) and gentoo penguin (lower spatial
scale). After Reid et al. (2005).
Monitoring predator–prey interactions 161
Detection of change
3
Reproductive output index
2
−1
Year
Fig. 11.2 Fluctuations in breeding performance – as measured by a reproductive
output index (ROI) – for Antarctic fur seals, gentoo penguins, macaroni
penguins and black-browed albatrosses breeding at Bird Island, South Georgia.
Details of variables comprising the ROI are given in Reid & Croxall (2001), from
which this figure is taken.
Grey-headed albatross 9 80 53 41 45
Black-browed albatross 12 96 53 47 9
Gentoo penguin 13 9 78 9 10
Macaroni penguin 14 62 93 50 92
performance and krill availability came from data in single years of widely
different estimates of krill abundance (Box 11.2). The results presented
in Box 11.2 indicate that even with sufficiently major changes in prey
abundance, such that all dependent species show statistically significant
responses in terms of provisioning and productivity:
In some cases (e.g. Fig. 11.4) there was also evidence of strong relation-
ships between krill abundance from acoustic surveys and one or more of
the diet indices of krill availability to predators. The ability to generate valu-
able data on the population structure of prey, particularly krill, from preda-
tor sampling became a particular achievement of the programme (e.g. Reid
et al. 1996, 1999a, 1999b) and is reviewed by Reid et al. (Chapter 17 in
this volume). Although these data are unlikely to provide measures of prey
stock abundance, they are likely to be increasingly valuable as proxies for
understanding the nature of krill availability to predators and potentially
contributing to predictions for future years (Reid et al. 1999a).
Now that over a decade of annual prey-abundance estimates are avail-
able for South Georgia, more progress is being made in defining and
understanding the functional relationships involved. This is greatly assisted
1986 0.3 0.3 0.3
n = 2673 n = 3215 n = 324
0.2 0.2 0.2
Proportion
Proportion
Proportion
0 0 0
Proportion
Proportion
Proportion
0 0 0
26 30 34 38 42 46 50 54 58 62 66 70 26 30 34 38 42 46 50 54 58 62 66 70 26 30 34 38 42 46 50 54 58 62 66 70
Krill length (mm) Krill length (mm) Krill length (mm)
Fig. 11.3 Differences in length–frequency distribution of Antarctic krill taken by predators in years of average (1986) and low (1994) availability
of krill. From Croxall et al. (1999).
166 J. P. Croxall
1996
80
1998
60
40
1999
20
1994 2000
0
10 15 20 25
Krill density (g m−2)
Fig. 11.4 Relationship between proportion (% wet mass) of krill in the diet of
macaroni penguins from Bird Island and estimated krill density (g m−2 ) at the
nearby north-western end of South Georgia between 1994 and 2000 (r2 = 0.96,
F1,5 = 92.9, p = 0.0006). From Barlow et al. (2002).
Vector CSI
a
See Asseburg et al. (Chapter 18 in this volume) for more details.
80
S
70 S
S
60 S
C
50 P
F
CV
40 F F F
P F S
P
P F
30 S
P
20 P
M C M
10 M S
M M M M M
M
0
0 5 10 15 20 25 30
Ranking
1.5
1.0
0.5
CSI summer
0.0
0.5
1.0
1.5
2.0
2.5
0 20 40 60 80 100 120
Krill density (g m )
Fig. 11.6 Relationship between krill density and the summer CSI for
predators breeding at Bird Island (F1,98 = 28.73, p < 0.001); from Reid et al.
(2005). For examples including typical standard deviations see Boyd and
Murray (2001).
This last result (Fig. 11.6) reinforces the conclusions of Boyd and
Murray (2001) and Boyd (2002) that such relationships provide a
potential basis for management of krill to minimize adverse effects
on dependent species. Thus in the example shown, krill densities of
below 20 to 30 g m−2 could be taken as levels at which fishing should
be particularly constrained (e.g. by closed season, area restrictions
or reduction in catch level) to avoid exacerbating problems already
being encountered by the dependent species (see Boyd (2002)). Sug-
gestions for future work on methodologies for using these preda-
tor response curves in assessing krill availability and/or identifying
anomalous (low) years of krill availability are provided by Constable
and Murphy (2003).
170 J. P. Croxall
6000 1600
(c) (d)
5000 1400
Population
1200
4000
1000
3000
800
2000
600
1000 400
1980 1985 1990 1995 2000 1980 1985 1990 1995 2000
Year Year
Fig. 11.7 Population changes in (a) Antarctic fur seals, (b) black-browed
albatrosses, (c) gentoo penguins, and (d) macaroni penguins breeding at Bird
Island, South Georgia, 1980–2000. Dotted lines show the measured values;
continuous lines show a smoothed spline through these points. From Reid and
Croxall (2001).
ACKNOWLEDGEMENTS
I thank particularly the late Peter Prince with whom I started these pro-
grammes that were made possible by the vision and foresight of Dick
Laws; successive Heads of Biosciences at BAS – Nigel Bonner, Andrew
Clarke and Paul Rodhouse – gave unstinting support. The commitment of
Steve Hunter, Bill Doidge, Nick Lunn, Tony Williams, Simon Berrow, Iain
Staniland, Kate Barlow and Ian Boyd ensured the programme prospered,
however, without the exceptional field work of the research-assistant team
at Bird Island over the last 30 years this would have been impossible. Keith
Reid, Dirk Briggs and Andy Wood played vital roles in data management
and quality control; Chris Ricketts and Pete Rothery contributed statisti-
cal rigour to our efforts; to these and many others we owe the success of
the long-term monitoring programme at Bird Island and of the science it
underpins and complements.
I also thank Elizabeth Dixon, Julie Leland and Janet Silk for assistance
in the preparation of this chapter, which benefited from comments by Ian
Boyd, Keith Reid, Phil Trathan and an anonymous reviewer.
REFERENCES
Agnew, D. J. (1997). The CCAMLR ecosystem monitoring programme. Antarct.
Sci., 9, 235–42.
Ainley, D. G., Sydeman, W. J. & Norton, J. (1995). Upper trophic level predators
indicate interannual negative and positive anomalies in the Californian
Current food web. Mar. Ecol. Prog. Ser., 118, 69–79.
Arnold, J. M., Brault, S. & Croxall, J. P. Albatross populations in peril? A
population trajectory for black-browed albatrosses at South Georgia. Ecological
Applic., in press.
Barlow, K. E., Boyd, I. L., Croxall, J. P. et al. (2002). Are penguins and seals in
competition for Antarctic krill at South Georgia? Mar. Biol., 140, 205–13.
Bost, C. A. & Le Maho, Y. (1993). Seabirds as bio-indicators of changing marine
ecosystems: new perspectives. Acta Ecol., 14, 463–70.
Boyd, I. L. (1993). Pup production and distribution of breeding Antarctic fur seals
(Arctocephalus gazella) at South Georgia. Antarct. Sci., 5, 17–24.
174 J. P. Croxall
Weimerskirch, H., Inchausti, P., Guinet, C. & Barbraud, C. (2003). Trends in bird
and seal populations as indicators of a system shift in the Southern Ocean.
Antarct. Sci., 15, 249–56.
Williams, T. D. & Rothery, P. (1990). Factors affecting variation in foraging and
activity patterns of gentoo penguins Pygoscelis papua during the breeding
season at Bird Island, South Georgia. J. Appl. Ecol., 27, 1042–54.
12
Impacts of oceanography on the foraging
dynamics of seabirds in the North Sea
F . D AU N T , S . W A N L E S S , G . P E T E R S , S . B E N V E N U T I ,
J . S H A R P L E S , D . G R É M I L L E T A N D B . S C O T T
Prey densities of at least 100× the average are necessary for profitable
foraging by auks A. G. Gaston (2004)
Animal-borne instrumentation
The data collected by a variety of instruments attached to birds
describe the three-dimensional distribution and foraging behaviour
of individuals in detail (Wilson et al. 2002). It is also possible to
demonstrate preferences for particular hydrographic conditions, if
data are collected concurrently with oceanography. However, ship-
based cruises may not coincide in timing or location with the bird
deployments, and satellite images are frequently lacking because of
cloud cover. In addition, the number of deployments is usually low,
and so population inferences are often challenging.
At-sea surveys
Unlike animal-borne instrumentation, at-sea surveys generally have
the advantage of large sample sizes. In addition, oceanographic data
can be collected concurrently, although the slow speed of ships
results in a lack of synoptic measurement across a study area. Direct
observation of certain behaviours – especially by surface-feeders –
and interactions between individuals and species, can be made.
However, population-based inferences are problematic because the
status and origin of birds is usually unknown, and thus the intrinsic
constraints within which individuals are operating cannot be incor-
porated. In addition, data collection is biased towards what is visible
at the sea surface, and therefore depth usage is unknown.
The North Sea is a semi-enclosed shelf sea (Otto et al. 1990). In spring,
the interaction between tidal currents, solar irradiation, wind patterns and
bathymetry create a mosaic of mixed, stratified and frontal regions (see Scott
et al. (Chapter 4 in this volume)). The principal frontal zone (the tidal or
shallow sea front) occurs at a point between the shelf break and the coast
where the water is shallow enough for tidal mixing to reach the surface.
Inshore of the front the water is mixed, although under certain bathymetric
Impacts of oceanography on seabird foraging dynamics 179
conditions small regions of stratified water and fronts may occur. Pri-
mary production is typically concentrated at frontal regions (Pingree
et al. 1975, Franks 1992) and thermoclines in stratified water (Harder 1968,
Barraclough et al. 1969). Superimposed on this broad seasonal pattern are
variations in water structure due to the tidal cycle and wind. The tides
cause changes in water depth, current speed and direction (Mann & Lazier
1996).
The lesser sandeel is the principal prey of most North Sea seabirds (Fur-
ness & Tasker 2000). For much of the summer, autumn and winter, adult
sandeels are buried in the substrate, only entering the water column briefly
in winter to spawn (Robards et al. 1999). In spring, adult sandeels are active
in the water column during the daytime, returning to the sand at night
(Winslade 1974). Within this diurnal pattern, their distribution in the water
column is expected to be dependent on the vertical distribution of their prin-
cipal prey, calanoid copepods. However, diurnal movements of calanoid
copepods are highly flexible, diverging from the typical migration from shal-
low depths at night to deep depths during the daytime depending on pre-
dation pressure (Frost & Bollens 1992). Because of this complex dynamic,
the distribution of adult lesser sandeels in relation to ocean physics and pri-
mary production is poorly understood. Larval sandeels metamorphose in
early spring into young-of-the-year fish which have an extended pre-settled
phase where they are present in the water column throughout the daily cycle
(Jensen et al. 2003). These young fish are also preyed upon by seabirds,
and are regularly aggregated in frontal zones (Camphuysen & Webb 1999,
Camphuysen et al. (Chapter 6 in this volume)), but other precise distri-
butions in the water column – including association with phytoplankton
biomass at the thermocline – are unknown.
Clupeids, and in particular sprats Sprattus sprattus, are important alter-
native prey for a number of seabird species in the North Sea. Like sandeels,
clupeids feed primarily on zooplankton but have a very different diurnal
distribution, foraging actively near the surface at night but being inactive
at deep depths during the daytime (Blaxter & Hunter 1982). As such, they
appear to match the typical vertical migration of their prey more closely,
but the extent to which this ties in with frontal features and thermoclines is
untested.
180 F. Daunt et al .
Surface-feeding species
Surface-feeding seabirds require processes that bring prey to the sea sur-
face (Garthe 1997, Camphuysen & Webb 1999). Horizontal frontal systems
are predicted to provide such opportunities, by driving prey such as young-
of-the-year lesser sandeels to the surface, in particular under certain tidal
phases when strong currents interact with bathymetry (see Camphuysen
et al. (Chapter 6 in this volume)).
We tested the prediction that chick-rearing black-legged kittiwakes Rissa
tridactyla, which feed predominantly on young-of-the-year lesser sandeels
at this time (Lewis et al. 2001), would target frontal regions. We equipped
breeding birds with activity loggers (Box 12.2) that allowed us to estimate a
maximum foraging range of 69 ± 6 km (see Fig. 12.1a and b). This distance
accords well with the distance from the breeding colony to the shallow sea
front, which runs parallel to the coast in our study area. Thus, the front
appears to form an outer barrier for breeding black-legged kittiwakes, with
foraging occurring throughout the zone between the colony and the front.
Our findings agree with at-sea surveys and telemetry that demonstrate a
consistent pattern of distribution of kittiwakes and other surface feeders
Impacts of oceanography on seabird foraging dynamics 181
at – and westwards of – the front, with very few east of the front (Cam-
phuysen & Webb 1999, Humphreys 2002). However, unlike at-sea surveys,
we found no strong evidence that birds were targeting the front over other
regions within the birds’ foraging range.
Mid-water divers
In the North Sea, the shallow sea front is important for mid-water divers
(Camphuysen & Webb 1999). In addition, these species can exploit the
water column, and would be expected to target depths where prey are con-
centrated. Primary production is aggregated at the thermocline in stratified
Impacts of oceanography on seabird foraging dynamics 183
250
200
179
150
100
50
0
414
0 100 200 300 400 500 600 700 800
Foraging trip duration (min)
Fig. 12.1 (a) Relationship between travelling flight duration and trip duration
during foraging trips of black-legged kittiwakes in 1999–2003. A broken-stick
model with flat asymptote provided the best fit to the model (77.9% of the
variation explained). The slope of the line is initially estimated at 0.43 ± 0.018,
before flattening at a trip duration of 414 min and a flight duration of 179 min.
Thereafter, there is no increase in distance travelled with increasing trip
duration. (b) Map of the study area, showing the inferred mean ± SE maximum
foraging range (69 ± 6 km) of kittiwakes breeding on the Isle of May. Maximum
range coincided with the position of the shallow sea front, and kittiwakes foraged
throughout the zone between the colony and the front. The 30-, 40- and 50-m
bathymetric contours are shown. Also shown are 45◦ and 135◦ bearings relative to
the colony, between which most kittiwake foraging trips are located (Humphreys
2002).
184 F. Daunt et al .
6 1.0
5
Chl A concentration (mg m ) 0.8
4
0.6
Proportion of dives
3
0.4
2
0.2
1
0 0
Above thermocline In thermocline Below thermocline
Fig. 12.2 Chlorophyll A concentration (mean + SD, open bars) and proportion
of dives of common guillemots (filled bars) above, in and below the thermocline,
18–26 June 2002.
waters (Mann & Lazier 1996). However, the extent to which the thermo-
cline is an important zone for foraging seabirds is dependent on its use by
their prey.
We examined the importance of the thermocline to a mid-water diver,
the common guillemot Uria aalge, a species that feeds on both lesser
sandeels and sprats (Harris & Wanless 1985). We equipped common guille-
mots with rapid-response temperature–depth loggers that record external
temperature very accurately (Daunt et al. 2003). Data collected concurrently
by fixed moorings allowed primary production to be modelled throughout
the water column (Box 12.2), and demonstrated that primary production
was highest at the thermocline, lowest below the thermocline and inter-
mediate above the thermocline (Fig. 12.2). Although guillemots foraged
almost exclusively in stratified water (Daunt et al. 2003), foraging effort
was strongly targeted at the zone below the thermocline (generalized lin-
ear mixed model (GLMM) with individual as random effect, W = 28.0,
p < 0.001; Fig. 12.2; Box 12.2). This distribution matched the daytime dis-
tribution of sprats rather than lesser sandeels and accorded well with sprat
being the principal prey delivered to chicks during the study period (78% of
prey deliveries).
Benthic divers
23
22 0.34
21
0.32
20
0.30
19
18 0.28
0 1 2 3 4 5 0 1 2 3 4 5
Tidal height band (m) Tidal height band (m)
(c) 0.20
Proportion of foraging time
0.15
0.10
0.05
0
0 1 2 3 4 5
Tidal height band (m)
Fig. 12.3 (a) Foraging depth; (b) proportion of time in water spent at foraging
depth; (c) proportion of time spent foraging by European shags in each tidal
height band (1 = 0.5 to 1.5 m, 2 = 1.5 to 2.5 m, etc.) corrected for availability.
1993), resulted in the proportion of time at foraging depth being 20% lower
at high tide than low tide (ratio of time at foraging depth to dive + surface
duration, REML, tidal height: F1,46 = 76.1, p < 0.001; Fig. 12.3b). Despite
this, there was no evidence that birds preferred to forage during low tide
(Fig. 12.3c; GLMM: W = 0.2, p = 0.94).
CONCLUSIONS
There is support both for and against a direct association between oceanog-
raphy and the foraging dynamics of seabirds in the North Sea. The shallow
sea front is favoured by both surface-feeding and mid-water-diving species,
supporting the view that these physical features form an important focus
for marine life at all trophic levels. There is also some indication that the
tidal cycle has an impact on temporal foraging opportunities for surface-
feeders (see Camphuysen et al. (Chapter 6 in this volume)), although the
potential impact of tidal current and direction on horizontal distribution
remains unclear. In contrast, the shallow sea front was not favoured by
breeding black-legged kittiwakes over other habitats within their foraging
range. In addition, we found no evidence for a preference for the thermo-
cline by common guillemots diving in stratified regions, with birds consis-
tently diving through the thermocline to areas with the lowest levels of pri-
mary production. Finally, there was no evidence that benthic-feeding shags
adjusted their foraging activity in response to the tidal cycle.
There are a number of reasons why the link between oceanography and
avian top predators may be weak. The most significant of these is likely to be
the number of trophic links between ocean physics and seabirds (typically
four levels in the North Sea). The association between primary production
and oceanography is strong. However, as one moves up the food web the
interactions become more complex. North Sea seabirds are feeding primar-
ily on small fish that are active swimmers and may only aggregate at regions
of high productivity when zooplankton, their principal prey, are feeding in
these zones. Depth utilization and observations of prey delivered to chicks
both suggest that common guillemots are targeting sprats which are inac-
tive at the sea floor during the day. As such, a close association between
guillemots and primary production is not expected. The lack of a preference
for different phases of the tide by European shags suggests that there are
patterns in the behaviour of adult lesser sandeels that are more important
in determining the timing of foraging than reduced distance between sea
surface and foraging habitat apparent during low tide. Clearly, an important
focus for future research is to gain a better understanding of lesser sandeel
Impacts of oceanography on seabird foraging dynamics 187
ACKNOWLEDGEMENTS
REFERENCES
Barraclough, W. E., LeBrasseur, R. J. & Kennedy, O. D. (1969). Shallow scattering
layer in the subarctic Pacific Ocean: detection by high-frequency echo sounder.
Science, 166, 611–13.
Beaugrand, G. (2004). The North Sea regime shift: evidence, causes, mechanisms
and consequences. Prog. Oceanogr., 60, 245–62.
Blaxter, J. H. S. & Hunter, J. R. (1982). The biology of the clupeoid fishes. Adv. Mar.
Biol., 20, 3–203.
Boyd, I. L. & Arnbom, T. (1991). Diving behavior in relation to water temperature
in the Southern elephant seal: foraging implications. Polar Biol., 11, 259–66.
Camphuysen, C. J. & Webb, A. (1999). Multi-species feeding associations in
North Sea seabirds: jointly exploiting a patchy environment. Ardea, 87, 177–98.
Coyle, K. O., Hunt, G. L., Decker, M. B. & Weingartner, T. J. (1992). Murre
foraging, epibenthic sound scattering and tidal advection over a shoal near St
George Island, Bering Sea. Mar. Ecol. Prog. Ser., 83, 1–14.
Daunt, F., Benvenuti, S., Harris, M. P. et al. (2002). Foraging strategies of the
black-legged kittiwake Rissa tridactyla at a North Sea colony: evidence for a
maximum foraging range. Mar. Ecol. Prog. Ser., 245, 239–47.
Daunt, F., Peters, G., Scott, B., Grémillet, D. & Wanless, S. (2003). Rapid response
recorders reveal interplay between marine physics and seabird behaviour. Mar.
Ecol. Prog. Ser., 255, 283–8.
Impacts of oceanography on seabird foraging dynamics 189
Time–activity/energy budgets
The daily time–activity budget indicated that all species except shags spent
about 50% of their time at the colony and 50% at sea. Shags on the other
hand allocated only about 15% of their time towards food acquisition, and
stayed at the colony for the remainder of the time. Kittiwakes, gannets and
guillemots spent a considerable amount of their time at sea resting (15% to
30%), but resting at sea was negligible in shags. Shags and guillemots spent
a much smaller proportion of their time flying than kittiwakes and gannets,
reflecting the use of prey patches closer to the colony. Daily energy expendi-
ture (DEE) calculated for the four species considered (Table 13.2) compared
well with reported energy expenditures measured in the field using doubly
labelled water (DLW), where available. The time–energy budget emphasized
the relative importance of energetically expensive activities, especially flight,
on the overall daily energy expenditure. While birds spent only between 13%
and 34% of their day active at sea, this period accounted for 39% to 60%
of their daily energy expenditure. Gannets worked the hardest with a field
metabolic rate (FMR) of 3.9 × BMR, while all other species worked at a level
of around 3 × BMR (Table 13.2).
CPUE values (based on active time spent at sea; see Table 13.2) for
shags and gannets were high compared with the other species, with shags
foraging most efficiently (Table 13.3; foraging efficiency is defined as the
ratio of metabolizable energy gained during foraging to energy used during
foraging).
Sensitivity analysis
rate at the nest that was twice the BMR; this value was used for all
species except for the kittiwake where we used the measured value
from Humphreys (2002). To incorporate the effect of water tem-
perature on metabolic rate during resting at sea for kittiwakes and
gannets we used the slope given by Croll and McLaren (1993) for
guillemots. In the absence of data we assumed that metabolic costs
of travel flight and forage flight for the kittiwake are identical and
the same assumption was made for flying and plunge-diving for the
gannet. All estimates of energetic costs during flight were calculated
using the aerodynamic model of Pennycuick (1989), using the latest
version ‘Flight 1.13’. Wing morphology values were taken from Pen-
nycuick (1987). We accounted for the presumably higher flight costs
during the return trip, after birds have ingested food and carry food
for their chicks. Estimates of the daily energy expenditure of chicks
were based on those provided by Visser (2002) for all species except
the guillemot – which was taken from Harris and Wanless (1985),
corrected for assimilation efficiency.
Diet samples were collected as regurgitations, observations of
prey delivered to chicks or from food dropped at the ledge. A mean
calorific value for prey taken was established for each species based
on the biomass proportions of prey and its size. Calorific values of
the various prey items were taken from the literature (Hislop et al.
1991, Bennet & Hart 1993, Pedersen & Hislop 2001) accounting
for seasonal effects. We took assimilation efficiencies for the gan-
net from Cooper (1978) and for all other species from Hilton et al.
(2000b). Assimilation efficiency for chicks was assumed to be the
same as in adults except in kittiwakes, for which we took the value
from Gabrielsen et al. (1992).
Body masses were obtained from birds during routine handling
associated with ringing. Breeding success was determined as the
number of chicks fledged from surveyed nests where eggs had been
laid. We took water temperatures from Daunt et al. (2003) who, in
the same area, measured water temperatures directly from foraging
shags and guillemots during chick rearing.
The algorithm used to compile the time–energy budgets (‘base-
line situation’, see Table 13.1 for key input values) and to investi-
gate the different scenarios was based on Grémillet et al. (2003)
but incorporated the energetic requirements of chicks. CPUE val-
ues (Table 13.2) are based on the time spent underwater for shags
196 M. R. Enstipp et al.
and guillemots, the time spent in forage flight for kittiwakes and the
total time spent at sea for gannets (a CPUE value based on the active
time spent at sea is included in brackets to allow comparison across
species). We conducted a sensitivity analysis (Table 13.4) to test the
robustness of our algorithm (Grémillet et al. 2003).
Body mass (g) 361.64 ± 36.14 1780.43 ± 97.63 2998 ± 234 920.34 ± 57.44
Assimilation efficiency for chick (%) 80.00 ± 1.25
Calorific value of fish (kJ g−1 wet mass) 5.0 ± 0.5 5.4 ± 0.5 5.8 ± 0.6 5.1 ± 0.5
Water temperature at surface (◦ C) 11.1 ± 0.5 11.1 ± 0.5 11.1 ± 0.5 12.0 ± 0.5
Water temperature at bottom (◦ C) 10.3 ± 0.4 8.8 ± 0.5
BMR (kJ day−1 ) 267.28 726.07 ± 46.15 1256.28 ± 227.94 584.48
Energy costs, resting at colony (W kg−1 ) 13.69 ± 1.20 9.44 ± 0.6 9.70 ± 1.76 14.70 ± 1.47
Energy costs, resting at sea (W kg−1 ) 12.82 ± 2.56 17.18 ± 2.02 12.46 ± 2.16 10.19 ± 1.02
Energy costs, flying (W kg−1 ) 44.83 ± 4.48 98.07 ± 9.81 43.69 ± 4.37 92.58 ± 9.26
Energy costs, foraging (W kg−1 ) 44.83 ± 4.48 20.58 ± 2.8 43.69 ± 4.37 23.83 ± 2.38
DEE of chick (kJ day−1 ) 525.71 ± 52.57 1203.98 ± 120.40 1593.30 ± 159.33 221.71 ± 22.17
Adult
DEE (kJ day−1 ) 786.74 2249.25 4856.01 1641.01
FMR (× BMR) 2.9 3.1 3.9 2.8
DFI (g fish day−1 ) 211 514 1114 415
Chick
DEE (kJ day−1 ) 525.71 1203.98 1593.30 221.71
DFI (g fish day−1 ) 131 275 366 56
No. of chicks fledged/pair 0.71 1.51 0.67 0.69
DFI (g fish day−1 , portion/adult) 47 208 122 19
Total
DFI (g fish day−1 ) 258 722 1237 434
CPUE (g fish min−1 ) 1.35 (0.50) 10.10 (3.84) 1.63 (3.89) 2.45 (1.18)
CPUE values are based on the time spent underwater for shags and guillemots, the time spent in forage flight for kittiwakes and the total time
spent at sea for gannets. To allow comparison across species a CPUE value based on the active time spent at sea (excluding periods of rest at sea)
is included in brackets.
Table 13.3. Foraging efficiency (ratio of metabolizable energy gained during foraging to energy used during foraging) and
foraging range of four North Sea seabirds
a
Excludes periods of rest at sea.
Table 13.4. Sensitivity analysis for the time–energy budget of four North Sea seabirds
Minimum and maximum input values for each parameter were used (see Table 13.1) to compute the individual variation in mean DFI (%). Minimum
and maximum values for all parametersa combined were computed for the most and least demanding situation, which indicates the maximum range of
potential DFI values for the birds.
Foraging energetics of North Sea birds 201
4.2
Energy expenditure (× BMR) 4 x BMR
4.0
3.8
+222%
3.6
+75%
3.4
+111%
3.2
Kittiwake
3.0 Shag
Gannet
2.8 Guillemot
1400
1300
Daily food intake (g day−1)
800
700
600
500
400
300
200
0 50 100 150 200 250 300
−1
Increase in foraging time (min day )
Fig. 13.1 Scenario 1 (increasing foraging time within a prey patch). Zero
indicates the ‘baseline situation’ (i.e. before increasing foraging time). See
Box 13.2 for details.
3.6 +69%
3.4
3.2
Kittiwake
3.0 Shag
Gannet
2.8 Guillemot
1400
1300
Daily food intake (g day−1)
900
800
700
600
500
400
300
200
0 25 50 75 100 125 150 175 200
Fig. 13.2 Scenario 2 (foraging at a more distant prey patch). Zero indicates the
‘baseline situation’ (i.e. before increasing flight time). See Box 13.2 for details.
their foraging time by about 222%, their flight time by about 106% and
their total active time at sea by about 112%. Comparable values for guille-
mots were 111%, 190% and 65%, respectively. Increased flight times in sce-
nario 2 potentially doubled the foraging range of shags while it tripled that of
guillemots (Table 13.3). Kittiwakes and gannets had the potential to increase
their foraging range by 69% and 9% respectively.
Food responses
Birds are constrained by the amount of food available and by the rates at
which they can acquire food. In many cases we know little about sandeel
abundance in the North Sea but we know even less about the prey-capture
capacities of seabirds and the fish densities they require to forage effectively.
Foraging energetics of North Sea birds 205
4.2
Energy expenditure (× BMR) 4 x BMR
4.0
+9% +112% +65%
3.8
3.6
3.4 +29%
3.2
Kittiwake
3.0 Shag
Gannet
2.8 Guillemot
1400
1300
Daily food intake (g day−1)
900
800
700
600
500
400
300
200
0 50 100 150 200 250
−1
Increase in active time at sea (min day )
Fig. 13.3 Scenario 3 (foraging at a more distant prey patch and for a longer time
within that prey patch). Zero indicates the ‘baseline situation’ (i.e. before
increasing flight and foraging time). See Box 13.2 for details.
(a)
CPUE (g min )
−1
4.0
2200
3.5
2000
3.0
1800
2.5
1600 2.0
0 50 100 150
Increase in flight time (min day−1)
(b)
Daily energy expenditure (kJ day −1)
3200 15
Shag
14
3000
4 x BMR 13
CPUE (g min )
−1
12
2800
11
2600
10
9
2400
8
2200 7
0 50 100 150
Fig. 13.4 (a) Scenario 2 (foraging at a more distant prey patch), and combined
with feeding on prey of reduced caloric density for the guillemot. (b) Scenario 3
(foraging at a more distant prey patch and for a longer time within that prey
patch), and combined with feeding on prey of reduced caloric density for the
shag. Circles indicate DEE, while squares indicate CPUE. Filled symbols indicate
scenarios 2 and 3, while open symbols indicate combination of the respective
scenario with feeding on less profitable prey. CPUE values are based on the time
spent underwater.
Foraging energetics of North Sea birds 207
shows scenario 3 in the shag (foraging at a distant prey patch for a longer
period).
How steep the increase in foraging effort will need to be depends on how
energetically expensive the associated activities are. Making use of a prey
patch at a greater distance from the colony, requiring longer flight times,
greatly increased daily food requirements for most species considered here
(Fig. 13.2). There will be a limit, imposed by the food availability, at which a
further increase in foraging effort becomes unsustainable.
Feeding rates reported in the literature (6 to 12 g fish min−1 underwater
for shags, Wanless et al. (1998); 0.5 to 1.3 g fish min−1 at sea for gannets,
Garthe et al. (1999)) are typically within the range or slightly lower than our
estimates for the ‘baseline situation’ (before increasing foraging effort). The
same holds true for DFI values reported for kittiwakes and guillemots. This
could indicate that birds might not be able to achieve the feeding rates that
would be required in the above scenarios when foraging effort is drastically
increased.
ACKNOWLEDGEMENTS
Most of this work was funded by the European Commission project ‘Inter-
actions Between the Marine Environment, Predators and Prey: Implica-
tions for Sustainable Sandeel Fisheries (IMPRESS; QRRS 2000-30864)’.
We thank Svein-Håkon Lorentsen, Claus Bech, Geir Håvard Nymoen,
Hans Jakob Runde and Magali Grantham for help with the captive work
on shags and guillemots. The Norwegian Animal Research Authority
granted research permits for this study (reference numbers 7/01 and
1997/09618/432.41/ME). Thanks also to the many people who helped with
field work, particularly Stefan Garthe, Mike Harris, Janos Hennicke, Sue
Lewis, Gerrit Peters and Linda Wilson. Scottish Natural Heritage and Sir
Hew Hamilton-Dalrymple granted permission to work on the Isle of May
and the Bass Rock, respectively. Luigi Dall’Antonia and Alberto Ribolini
developed the flight activity sensors used on kittiwakes and provided soft-
ware for data analysis. Thanks to Ian Boyd and two anonymous referees for
improving an earlier version of this chapter.
REFERENCES
Bennet, D. C. & Hart, L. E. (1993). Metabolizable energy of fish when fed to captive
great blue herons (Ardea herodias). Can. J. Zool., 71, 1767–71.
Birt-Friesen, V. L., Montevecchi, W. A., Cairns, D. K. & Macko, S. A. (1989).
Activity-specific metabolic rates of free-living northern gannets and other
seabirds. Ecology, 70, 357–67.
Foraging energetics of North Sea birds 209
BIOENERGETICS CONSIDERATIONS
Consumption by seabirds
Estimated food consumption by seabirds in various regions.
Consumption
Region (t km−2 ) Reference
Sandeel consumption
Consumption of sandeels Ammodytes by major consumer groups in
the North Sea (ICES Areas IV a,b,c) in 1969 and 1999; from Furness
(2002).
consume may be much less than the amount that they need to have in the
environment in order to forage effectively.
of sandeels.
(Figs. 14.2 and 14.3), yet (following an equivalent calculation to that above)
Shetland kittiwakes consumed less than 1% of this amount. Similar log-
arithmic or sigmoidal relationships between predator local abundance or
breeding success and food fish abundance have been reported for guille-
mot species (Mehlum et al. 1999, Fauchald et al. 2000, Fauchald & Erik-
stad 2002), for Atlantic puffins Fratercula arctica (Durant et al. 2003), for
least auklets Aethia pusilla (Hunt 1997), for kelp bass Paralabrax clathra-
tus (Anderson 2001) and for basking sharks Cetorhinus maximus (Sims
2000).
Another example where predator performance can be related to fish
stock biomass is the case of the common guillemot Uria aalge in the Barents
Sea, where this species feeds almost exclusively on capelin Mallotus villosus
(Barrett & Krasnov 1996). Food requirements of common guillemots in the
Barents Sea have been estimated at 70 000 t per year (Mehlum & Gabrielsen
1995). The capelin stock biomass was around 6 million tonnes in 1980, but
fell rapidly in the mid 1980s to about 500 000 t in 1985. In 1985–7 over 90%
How many fish should we leave in the sea? 217
1.2
1.0
0.8
0.6
0.4
y = 0.5428 ln(x) − 5.2379
0.2 R 2 = 0.8032
0
0 50000 100000 150000 200 000
−0.2
Shetland sandeel total stock biomass (tonnes)
Fig. 14.1 Breeding success of Arctic skuas in Foula, Shetland from 1975 to
1994, in relation to the estimated abundance of sandeels at Shetland (VPA
estimate of total stock biomass in tonnes; data from ICES). The fitted line is a
logarithmic regression.
Kittiwake breeding success at Foula
1.6
1.4
1.2
1.0
0.8
0.6
0.4 y = 0.4735 ln(x ) − 4.3601
0.2 R 2 = 0.602
0
0 50 000 100 000 150 000 200 000
Shetland sandeel total stock biomass (tonnes)
Fig. 14.2 Breeding success of kittiwakes in Foula, Shetland from 1975 to 1994,
in relation to the estimated abundance of sandeels at Shetland (VPA estimate of
total stock biomass in tonnes; data from ICES). The fitted line is a logarithmic
regression.
CONCLUSION
It seems clear that calculating the amount of fish eaten by top predators
does not provide a useful approach to defining the minimum stock biomass
necessary to support these top-predator populations. The seabird examples
218 R. W. Furness
1.2
0.8
Shetland
0.6
0.4
y = 0.3673 ln(x) − 3.4806
0.2
R 2 = 0.7351
0
0 20000 40000 60000 80000 100000 120000 140000
Shetland sandeel total stock biomass (tonnes)
quoted above show that populations require very much larger biomasses of
food fish in order to allow them to forage to obtain the relatively smaller
quantity of food that they need. The few empirical examples available from
the literature suggest that a decrease in food-fish stock biomass to about
20% of ‘normal’ may cause catastrophic reproductive failure in sensitive
seabird species, at least in the short term. However, the impact of such a
large reduction may be very slight for species of low sensitivity. Whether the
concept of species-specific sensitivity based on the ecology of each species
can be extended to other top predators – such as marine mammals and pis-
civorous fish – remains to be seen, but such an approach identifying ‘sen-
tinel’ species of high sensitivity as a proxy to represent the community of
top predators may provide a way to define prey-fish stock biomass (or den-
sity) limits that should provide adequate foraging opportunities for most top
predators in particular ecosystems. This approach has already been put into
effect with the management of the North Sea sandeel fishery; low breed-
ing success of black-legged kittiwakes in southeast Scotland and northeast
England was taken as a signal that sandeel fishing in that area should be
halted due to its effect on populations of dependent species.
It is noteworthy that at least the more sensitive seabird species suffer
reproductive failure at a food-fish stock biomass level that is somewhat
higher than might be set as a limit reference point on the basis of single-
stock assessment to protect recruitment. Reference-point management for
ecosystem sustainability probably cannot simply adopt reference points set
on the basis of fish stock recruitment needs alone. Although this paper
How many fish should we leave in the sea? 219
ACKNOWLEDGEMENTS
I thank the European Commission for funding the project ‘DISCBIRD’ and
the Natural Environment Research Council and the Shetland Oil Termi-
nal Environmental Advisory Group (SOTEAG) for funding research into
seabird ecology in Shetland over much of the period 1971–2003. I am grate-
ful to Mark Tasker for helpful comments on an earlier draft of this chapter.
REFERENCES
Anderson, T. W. (2001). Predator responses, prey refuges, and density-dependent
mortality of a marine fish. Ecology, 82, 245–57.
220 R. W. Furness
Industrial fisheries remove large quantities of small fish from the North Sea
ecosystem each year. Since these small fish constitute the prey of marine top
predators, such activities are considered to pose a potential threat to marine
food-web dynamics. The risk to seabird and marine mammal communities
has in the past received most attention, but more recently concern has been
expressed regarding the possible consequences of industrial fishing for pis-
civorous fish populations, often the target of fisheries for human consump-
tion. These concerns are addressed in this chapter. A major industrial fish-
ery for sandeels opened on the Wee Bankie in the northwestern North Sea in
the early 1990s. Subsequently, in 2000, this fishery was closed in response
to concern over its possible impact on local seabird populations. The effect
of this closure on the abundance of sandeels in the area – and on local gadoid
population abundance, diet, food consumption rates and body condition –
are described to examine the effects of the sandeel fishery on these pis-
civorous, predatory-fish populations. Although closing the sandeel fishery
resulted in an immediate increase in the local abundance of sandeels, no
beneficial effect on local gadoid populations was detected. Gadoid predators
in the area prey almost entirely on 0-group sandeels (fish‘born’ in the cur-
rent year), while the fishery took predominantly older-aged sandeels. Thus
these two consumers appear not to have directly competed for the same
resource.
Industrial fishing, the catching of fish for industrial purposes, such as
for the production of fishmeal, rather than for human consumption, has
taken place in the North Sea for several decades. However, in the mid 1970s
the industrial fishery expanded considerably. At this time the percentage of
the total fish removals from the North Sea that was landed for industrial use
increased from less than 20% to over 50%. This expansion coincided with
increased targeting of sandeels by the fishery (ICES 2002). The industrial
fishery has long been a source of controversy and several different concerns
have been raised. Firstly, the large biomass of fish removed by the fishery
each year has worried many people, over 1.7 million tonnes in 1975 and close
to or exceeding 1 million tonnes every year since (Arnott et al. 2002, ICES
2002). Such catch levels represent the removal by the industrial fishery of
between 10% and 15% of the North Sea fish standing-crop biomass each
year (Yang 1982, Daan et al. 1990, Sparholt 1990). Secondly, in times when
fisheries for many species are in crisis, it is possible that large bycatches of
the juvenile stages of more valuable human-consumption species – such
as herring, cod and haddock – might be included in the catches taken by
the industrial fishery. Fishermen fear that these bycatches reduce recruit-
ment potential, contributing to stock declines and inhibiting recovery in
depleted stocks. Finally, the fish landed by the industrial fisheries consti-
tute the prey of predators higher up the food web. Consequently, there has
been concern over the potential for competition between industrial fisheries
and marine top predators (Avery & Green 1989, Furness 1996, Naylor et al.
2000).
When these concerns come to the fore, the traditional management
response has been to adopt a precautionary approach and to implement
‘local’ industrial-fishery moratoria. Thus in 1983 the winter sprat fishery in
the Moray Firth, northeast Scotland, was closed due to increasing concern
over the growing bycatch of juvenile herring among the fish caught (Hop-
kins 1986). In 1991 the fishery for sandeels around the Shetland Islands, to
the north of Scotland, was closed in an attempt to reverse a worrying down-
ward trend in the breeding success of seabirds in the area (Wright 1996).
Similar concerns were raised when, in the mid 1990s, seabird breeding suc-
cess started to decline at colonies in the Firth of Forth in southeast Scotland,
shortly after a new sandeel fishery commenced on the nearby Wee Bankie
and Marr Bank. In 2000, with no significant indication of any improvement
in the situation, a precautionary stance was taken and fishing for sandeels
along much of the east coast of Scotland was prohibited (Wright et al.
2002).
It is interesting to note that, to date, the top-predator concerns that have
elicited a management response have always involved seabirds. However,
during the 1990s fishermen were also making similar comments to those
Prohibition of industrial fishing for sandeels 225
The closure of the sandeel fishery along the Scottish east coast in 2000 pro-
vided an opportunity to explore the validity of the fishermen’s concerns. In
1997, in response to the outcry over declining breeding success at seabird
colonies in the Firth of Forth and elsewhere along the Scottish east coast, a
research programme was initiated to examine interactions between marine
top predators and their sandeel prey. Field studies focused on the nearby
Wee Bankie–Marr Bank region, which in the early 1990s had become a
major sandeel fishing ground. By 2000 therefore, 3 years of data regard-
ing changes in the abundance, distribution and age–length composition
of the sandeel population – and changes in the diet, food consumption
rates and body condition of the main piscivorous fish species populations
in the area – had already been gathered. During this time the sandeel
Prohibition of industrial fishing for sandeels 227
Fishing for sandeels off the east coast of Scotland – principally on the Wee
Bankie, Marr Bank and Berwick Bank – commenced in the early 1990s. In
1993 catches from the area exceeded 100 000 t. In 2000 the fishery was
closed to protect breeding seabirds, and since this time fishing removals
(continued for scientific purposes) have not exceeded 5000 t in any year
(Fig. 15.1). The largest proportion of the sandeels removed by the fish-
ery over this entire period has consisted of sandeels of at least 1 year old.
Two separate fisheries-independent assessment methods indicated a con-
siderable increase in the biomass of sandeels present in the Wee Bankie–
Marr Bank–Berwick Bank complex following closure of the sandeel fishery
(Fig. 15.1).
The predominant piscivorous-fish species in the area were cod, haddock
and whiting. Sandeels constituted around 50% of the diet of cod up to 30
cm in length, falling to less than 10% of the diet of cod greater than 45 cm in
length. Approximately 50% of the diet of haddock smaller than 20 cm con-
sisted of sandeels, and this percentage increased to around 75% in haddock
of 45 cm in length. Whiting were fish specialists; 85% of the diet of 15-cm
whiting consisted of sandeels and this percentage declined to around 75% in
whiting of over 30 cm in length (Fig. 15.2). The biomass of all three gadoid
predators declined over the period 1997–2003 (Fig. 15.2). This change in
predator biomass was opposite to the change expected given the increase
in sandeel abundance in the area. Variation in cod and haddock biomass
in the area was strongly correlated with changes in stock size in the entire
North Sea (Fig. 15.2). These two predator populations in the Wee Bankie–
Marr Bank region appeared more strongly influenced by factors affecting
stocks in the whole of the North Sea, such as variability in recruitment and
fishing mortality, than by local changes in management policy regarding
the sandeel fishery.
Annual variation in the proportion of sandeels in the diets of the three
gadoid predators and variation in their daily food consumption rates were
228 S. P. R. Greenstreet
56.50
50-m depth
contour
Marr
Degrees latitude Bank
56.25
56.00
−3 −2.75 −2.5 −2.25 −2 −1.75 −1.5 −1.25 −1
Degrees longitude
120000 300 25
10
40000 Fishery 100
closed
Fishery 5
closed
0 0 0
1990 1995 2000 2005 1996 1998 2000 2002 2004
Year Year
Fig. 15.1 Chart of the study area showing the main sandbanks where the
sandeel fishery was carried out. The size of the symbol indicates the numbers of
vessels recorded at each location by the Scottish Fisheries Protection Agency in
1996 and 1997. Landings of sandeels taken from the study area are presented
and the period of the moratorium is indicated. Trends in two
fisheries-independent, sandeel-abundance indicators – a demersal trawl survey
(see Fig. 15.2 legend) and an acoustic survey (Mosteiro et al. 2004) – show
changes in the size of the sandeel population within the area over the duration of
the study; again the fishery closure period is shown.
0.8 0
Diet
0.6 Invertebrates 0.4
Other fish
0.4 Flatfish
0.8 Cod
Gadoids
Clupeids
0.2 1.2
Sandeels
0 Y = 2.85114 X 1.7938
1.6
0 20 40 60 80 10 20 30 40 50 10 15 20 25 30 35 40 0.2 0.4 0.6 0.8
Length (cm) Length (cm) Length (cm) 2.0
r = 0.917, p = 0.010
1.8
2000 50 Whiting
1.85
0 0 1.80
Fig. 15.2 Nineteen evenly spaced demersal trawl samples were collected in each
year and the catches quantified (number per size class per square meter swept).
These trawl densities were raised to the size of study area to estimate the total
biomass of each species present in each year. Size-stratified (12–14.9, 15–19.9,
20–24.9, 25–29.9, 30–34.9, 35–39.9, 40–49.9 and then every 10 cm thereafter)
sub-samples of the catch of cod, haddock and whiting were dissected for stomach
analysis. Sufficient fish, up to a limit of 25, were examined so as to obtain
between 10 and 15 stomachs with food contents per size class of each predator in
each trawl. Changes in the diet with length of cod, haddock and whiting; trends
in the abundance of these species in the area over the period 1997 – 2003 (note
different Y-axis for cod abundance); and correlations between these study-area
trends and similar trends for the entire North Sea stock are illustrated.
certainly no ‘release’ of their populations when the fishery was closed. The
proportion of sandeels in the diets of the gadoid predators, and their daily
food consumption rates, were not positively related to changes in sandeel
abundance; and there was no increase in either factor when the sandeel fish-
ery was closed. Consequently gadoid-predator body condition also remained
unaffected by the fishery moratorium. Why was this?
The hypotheses posed by the fishermen assume that the industrial fish-
ery and the gadoid predators in the Wee Bankie–Marr Bank region com-
peted directly for the same sandeel resource, but was this actually the case?
230 S. P. R. Greenstreet
100
Cod Haddock
Mann–Whitney p = 0.827 Mann–Whitney p = 0.827
Diet (% sandeel) 80
60
Whiting
40 Mann–Whitney p = 0.827
20
0
Daily consumption (% BM)
3
Cod Haddock
Mann–Whitney p = 0.827 Mann–Whitney p = 0.127
2.5
2
Whiting
1.5 Mann–Whitney p < 0.05
1
Yes No Yes No Yes No
Sandeel fishery in operation
Fig. 15.3 The mean weight of stomach contents was determined for each size
class of each predator in each cruise and Jones’ (1974, 1978) digestion rate model
applied to estimate predator daily rations (g. day−1 ). Daily food consumption
rates as a percentage of predator biomass were determined by dividing daily
rations by predator body weights. The effects of closure of the sandeel fishery on
the percentage (by weight) of sandeels in the diet of each gadoid predator, and on
daily food consumption rates (expressed as a percentage of predator biomass
(BM)), are illustrated.
CONCLUDING REMARKS
3
Cod Haddock Whiting
Daily ration (% of predator biomass)
2.5
1.5
1.05 1.06
1.10
1.04 1.05
Condition factor
Condition factor
Condition factor
1.08 1.04
1.03
1.03
1.02 1.06
1.02
1.01
1.04 1.01
1.00 1.00
1.02
0.99 0.99
0.98 1.00 0.98
No Yes No Yes No Yes
Sandeel fishing in operation Sandeel fishing in operation Sandeel fishing in operation
Fig. 15.5 Predator body condition (a ratio of weight per unit length) factors were
determined following the methods proposed by Kruuk et al. (1987) and
Thompson et al. (1996). Single weight-at-length relationships (of the form W =
aLb ) were derived by linear regression on all the log-transformed data collected
over the whole 7-year period for each species. Individual condition factors were
determined by dividing observed individual weights by the weight predicted by
the regression. The individual weight data used were fish tissue weight only, i.e.
with all the weight of food in the stomachs and intestines subtracted. The effects
of closure of the sandeel fishery on fish condition factors within the cod, haddock
and whiting populations in the Wee Bankie–Marr Bank study area are illustrated.
232 S. P. R. Greenstreet
1200 600
40
Number
800 400
20
400 200
0 0 0
0 5 10 15 20 0 5 10 15 20 0 5 10 15 20
Length (cm)
1.0
0.15 0.15
Proportion
Proportion
0.8
0.4
0.05 0.05
0.2
ACKNOWLEDGEMENTS
Many colleagues at the Marine Laboratory assisted both with the collection
of the data at sea and in its subsequent analysis; particularly Eric Arm-
strong, Iain Gibb, Helen Fraser, Gayle Holland, Cathy Doyle, Rose Li and
Mike Robertson. Conversations with Mike Heath, Peter Wright and others
influenced the direction of the study and I am grateful to Peter Wright for
his comments on the manuscript. I also thank Ian Boyd and two anony-
mous referees for their help in improving the manuscript. These data were
collected as part of two EC projects: The Effect of Large-Scale Industrial
234 S. P. R. Greenstreet
REFERENCES
Arnott, S. A. & Ruxton, G. D. (2002). Sandeel recruitment in the North Sea:
demographic, climatic and trophic effects. Mar. Ecol. Prog. Ser., 238, 199–210.
Arnott, S. A., Ruxton, G. D. & Poloczanska, E. S. (2002). Stochastic dynamic
population model of North Sea sandeels, and its application to precautionary
management procedures. Mar. Ecol. Prog. Ser., 235, 223–34.
Avery, M. & Green, R. (1989). Not enough fish in the sea. New Sci., 1674, 28–9.
Bogstad, B. & Gjøsæter, H. (2001). Predation by cod (Gadus morhua) on capelin
(Mallotus villosus) in the Barents Sea: implications for capelin stock assessment.
Fish. Res., 53, 197–209.
Bogstad B. & Mehl, S. (1997). Interactions between Atlantic cod (Gadus morhua)
and its prey in the Barents Sea. In Proceedings of the International Symposium on
the Role of Forage Fishes in Marine Ecosystems. Alaska Sea Grant College
Program Report AK-SG-97-01. Fairbanks, Alaska: University of Alaska
Fairbanks, pp. 591–616.
Daan, N. (1989). Database Report of the Stomach Sampling Project 1981. ICES
Cooperative Research Report 164. Copenhagen, Denmark: ICES, pp. 1–144.
Daan, N., Bromley, P. J., Hislop, J. R. G. & Nielsen, N. A. (1990). Ecology of North
Sea fish. Neth. J. Sea Res., 26, 343–86.
Furness, R. W. (1996). A review of seabird responses to natural or
fisheries-induced changes in food supply. In Aquatic Predators and Their Prey,
eds. S. P. R. Greenstreet & M. L. Tasker. Oxford, UK:, Blackwell Science, pp.
166–73.
(2002). Management implications of interactions between fisheries and sandeel
dependent seabirds and seals in the North Sea. ICES J. Mar. Sci., 59, 261–9.
Greenstreet, S. P. R., McMillan, J. A. & Armstrong, F. (1998). Seasonal variation in
the importance of pelagic fish in the diet of piscivorous fish in the Moray Firth,
NE Scotland: a response to variation in prey abundance? ICES J. Mar. Sci., 55,
121–33.
Hislop, J. R. G. (1997). Database Report of the Stomach Sampling Project 1991. ICES
Cooperative Research Report 219. Copenhagen, Denmark: ICES, pp. 1–442.
Hopkins, P. J. (1986). Exploited fish and shellfish populations in the Moray Firth.
Proc. R. Soc. Edinburgh, Ser. B, 91, 57–72.
ICES (2002). Report of the Working Group on the Assessment of Demersal Stocks in the
North Sea and Skagerrak. ICES CM 2002, ACFM01. Copenhagen, Denmark:
ICES, pp. 1–555.
Jones, R. (1974). The rate of elimination of food from the stomachs of haddock
Melanogrammus aeglefinus, cod Gadus morhua, and whiting Merlangius
merlangus. J. Cons. Int. Explor. Mer, 35, 225–43.
(1978). Estimates of the food consumption of haddock (Melanogrammus
aeglefinus) and cod (Gadus morhua). J. Cons. Int. Explor. Mer., 38, 18–27.
Prohibition of industrial fishing for sandeels 235
Large seabirds such as northern gannets Morus bassanus have very flex-
ible time–activity budgets; this means that changes in variables such as
foraging-trip duration could provide a rapid indicator of changes in food
supply, an indicator that could not be obtained from smaller species. More-
over, larger birds often have longer foraging ranges, giving them the poten-
tial to integrate information about changes in food availability over large
areas of ocean. There is insufficient information on temporal variation in
fish stocks exploited by far-ranging species to determine how the birds’ for-
aging ecology varies with prey abundance, but comparing colonies of dif-
ferent size potentially presents an opportunity to examine empirically how
foraging ecology varies in relation to prey availability. For gannets in Britain
and Ireland, there were major differences between colonies in foraging and
food-provisioning behaviour, but the relationship between trip duration and
foraging range was remarkably constant. Moreover, there was a strong rela-
tionship between trip duration and the square root of colony size, which was
very similar within colonies between years and between colonies within a
single year. This relationship could provide a powerful tool for gauging the
importance of changes in trip duration in terms of changes in per-capita
prey availability. Over 4 years, annual variation in diet at one colony to some
extent reflected variation in trip durations and foraging locations, although
one year was anomalous and a combination of trip duration and diet pro-
vided a much more complete picture than either did on their own.
Many studies have shown that seabirds are sensitive to changes in food
supply and so have the potential to act as monitors of fish stocks (see reviews
by Montevecchi (1993), and Furness and Camphuysen (1997)). Responses
vary among species but small seabirds such as terns, which spend a high
proportion of available time foraging, are generally considered the most sen-
sitive indicators (Furness & Ainley 1984, Monaghan 1992). Larger birds
have more flexibility in their time–activity budgets and so are often consid-
ered to be less sensitive because they can buffer the impacts of reductions
in food supply on variables such as breeding success and chick growth by
increasing their time spent foraging. However, the very fact that they can do
this means that changes in variables such as foraging-trip duration could
provide a rapid indicator of changes in food supply, which could not be
obtained from smaller species. Moreover, larger birds often have longer for-
aging ranges (e.g. >500 km in northern gannets Morus bassanus, hereafter
termed gannets; Hamer et al. (2000)), giving them the potential to integrate
information about changes in food availability over large areas of ocean.
The main problem to date in using far-ranging seabirds such as gan-
nets as monitors of marine fish stocks is that there is insufficient infor-
mation over sufficiently large spatial scales on temporal variation in fish
stocks exploited as prey, or on how variables other than stock size affect
prey availability. Thus the relationships between foraging-trip duration and
prey availability or abundance are unknown. This is a particular problem
for gannets, which exploit a wide range of species and sizes of prey (Hamer
et al. 2000, Lewis et al. 2003) – each of which would need to be assessed
independently to give a full picture of prey population sizes. Moreover,
while a reduction in prey availability would intuitively be expected to result
in longer foraging trips, this need not be the case. For instance, individuals
of many species, including gannets, spend a proportion of the time at sea
apparently inactive (Hamer et al. 2001, Lewis et al. 2002), and may respond
to a reduction in food supply by reducing this time without any change in
overall trip duration (Enstipp et al. (Chapter 13 in this volume)).
There are currently 21 separate gannet colonies in the British Isles, rang-
ing in size from <150 to >60 000 breeding pairs (Wanless & Harris 2004).
If birds at a colony are competing for food resources, then those at larger
colonies should experience lower per-capita prey availability as a result of
greater competition. If gannets compete for available prey that is randomly
distributed in discrete patches, then they should forage over approximately
the same total area per bird to obtain the same amount of food, irrespec-
tive of colony size. The area covered increases with the square of the mean
238 K. C. Hamer et al.
One of the largest gannet colonies in Britain is at the Bass Rock, southeast
Scotland (56◦ 6 N, 2◦ 36 W; >40 000 breeding pairs) and one of the small-
est is at Great Saltee, southeast Ireland (52◦ 7 N, 6◦ 37 W; 2000 pairs). We
examined foraging-trip durations and ranges, and at-sea behaviour of birds
at these two colonies in 1998 and 1999 respectively, using satellite teleme-
try (Box 16.1). At the Bass Rock, the mean duration of foraging trips was
31.3 h (SD, ± 11.0h; range, 13.1 to 84.0 h) and the mean distance to desti-
nation was 223 km (SD, ± 95 km; range, 39 to 540 km). Destinations of
foraging trips covered a wide area of ocean encompassing >200 000 km2
within the northwest, west and central North Sea (Fig. 16.1). Mean trip dura-
tion at Great Saltee was significantly shorter than at the Bass Rock (11.9 ±
6.7 h; range, 2.8 to 42.8 h; t-test using mean values for each bird, t12 = 10.5,
p < 0.001) as was the mean distance to destination (89 ± 49 km; range, 14
to 238 km; t12 = 9.5, p < 0.001). Destinations of foraging trips from Great
Saltee encompassed an area of about 45 000 km2 between the coasts of
northwest Wales, southwest England and southern Ireland (Fig. 16.1). This
was about one-quarter of the area covered by birds from the Bass Rock. In
association with this difference in foraging area, individual birds at the Bass
Rock showed a high degree of foraging-area fidelity, with successive trips
having very similar bearings and significant repeatability in distances trav-
elled; in contrast, birds at Great Saltee did not show this behaviour (Hamer
et al. 2001).
There was a highly significant relationship between maximum dis-
tance from the colony and trip duration at both the Bass Rock (F1,67 =
988.7, p < 0.0001, R2 = 0.94) and Great Saltee (F1,58 = 305.4, p < 0.0001,
R2 = 0.84). Average speed ( ± SE) over complete foraging trips was
14.1 ± 0.4 km h−1 at the Bass Rock and 13.8 ± 0.8 km h−1 at Great Saltee
Use of gannets to monitor prey availability 239
60°N
Bass Rock
58°N
56°N
54°N
52°N
100 km
Great Saltee
50°N
Fig. 16.1 Foraging ranges and destinations of foraging trips by gannets from the
Bass Rock, southeast Scotland and Great Saltee, southeast Ireland. Open circles,
locations of adults at sea; filled circles, destinations of foraging trips. Reproduced
from Marine Ecology Progress Series, with permission (Hamer et al. 2000).
600
500
Maximum distance (km)
400
300
200
100
0
0 20 40 60 80 100
Trip duration (h)
Fig. 16.2 Maximum distance from the colony and foraging-trip duration at the
Bass Rock (filled circles and continuous line) and at Great Saltee (open circles
and dashed line).
Bass Rock: Maximum distance (km) = 7.05 (SE, ± 0.22) × trip duration (h)
Great Saltee: Maximum distance (km) = 6.88 (SE, ± 0.39) × trip duration (h)
Use of gannets to monitor prey availability 241
Speeds of travel over intervals within trips were calculated using con-
secutive pairs of locations with >1 h between them (Box 16.1). The mean
( ± SD) of these values at the Bass Rock (18.1 ± 16.6 km h−1 , n = 797)
was very similar to the mean at Great Saltee (17.3 ± 17.2 km h−1 , n = 237)
and there was no difference in the frequency distribution of travel speeds
at the two colonies (Kolmogorov–Smirnov two-sample test, Z = 0.47, n =
1034, p = 0.98). At both the Bass Rock and Great Saltee, the mean ( ± SD)
speed during hours of darkness was very low (1.6 ± 1.9 km h−1 , n = 30
and 4.0 ± 5.4 km h−1 , n = 11 respectively), with much higher speeds dur-
ing daylight (22.3 ± 31.6 km h−1 , n = 767 and 21.8 ± 27.3 km h−1 , n =
226 respectively).
These data indicate that despite the large differences between colonies in
trip durations, distances travelled and foraging-area fidelity, the behaviour
of birds during foraging trips was very similar at the two colonies.
There was, however, a major difference in foraging and food-provisioning
behaviour that was not indicated by these data. At Great Saltee, chicks were
always attended by at least one parent whereas at the Bass Rock, chicks >4
weeks old were sometimes left unattended at the nest while both parents
foraged simultaneously (Lewis et al. 2004). Such trips were only about half
as long on average as attended trips and increased in frequency as chicks
grew, comprising almost 40% of all trips by the time chicks were >9 weeks
old. These unattended trips typically occurred after a longer than average
period of attendance at the nest, and so they are likely to be more frequent at
colonies where long foraging trips are relatively common. Since these unat-
tended trips were much shorter than attended trips, their occurrence could
have influenced the relationship between colony size and mean trip dura-
tion, although this is unlikely: the duration of unattended trips increased
as chicks grew (Lewis et al. 2004) and by the time they were sufficiently
frequent to have a significant impact on overall trip duration, they were of
similar duration to attended trips.
N N
BR
100 km
Fig. 16.3 Location and size of gannet colonies studied. The area of each circle is
proportional to the colony’s size in 2000 (A, Ailsa Craig; B, Bempton; BR, Bass
Rock; F, Fair Isle; G, Great Saltee; H, Hermaness; I, Ireland’s Eye; N, Noss; T,
Troup Head).
day for foraging (local daylight time minus the time birds were together at
the nest) by the estimated changeover rate. We also obtained historical data
on trip durations at four colonies from personal (S. Wanless, unpublished
data for 1980) and published data (Nelson 1978, Garthe et al. 1999).
Across the nine colonies studied, there was a significant positive corre-
lation between trip duration (hence foraging range) and the square root of
Use of gannets to monitor prey availability 243
25 175
150
20 BR
125
A
15
100
H
G
H 75
10 N
F A
BR
I 50
B
5
T
25
B
0 0
0 50 100 150 200 250
Square root of colony size
colony size (r = 0.90, d.f. = 7, p < 0.005) with trip duration increasing
three-fold between the smallest and largest colony (Fig. 16.4). Pairs at the
smaller colonies spent significantly more time together at the nest (square-
root-transformed data: r = −0.73, d.f. = 7, p < 0.05), but the rate of
provisioning of chicks by the parents still tended to decrease with increas-
ing population size (square-root-transformed data: r = −0.66, d.f. = 7,
p = 0.052; log-transformed data: r = −0.67, d.f. = 7, p < 0.05). His-
torical data on foraging-trip duration at four of these colonies (Fig. 16.4)
fitted the same regression model well (2000-only regression slope =
0.0492, SE = 0.009 07; pooled within colony regression slope for the four
colonies = 0.0622, SE = 0.0102; combined slope from a linear mixed
model = 0.0548, SE = 0.0128), indicating that the relationship within
colonies between years was similar to the relationship between colonies
within a single year. Data for St Kilda (the largest colony in the North
Atlantic) confirmed the generality of this relationship: in 1980 the colony
held 40 000 breeding sites and the average trip duration was around 21 h
244 K. C. Hamer et al.
(S. Wanless, unpublished data). Finally, of two pairs of colonies that were
very similar in size (Great Saltee (G) and Bempton (B); Fair Isle (F) and
Troup Head (T)) one of each pair was above the regression line of trip dura-
tion against colony size, and one was below. The colonies that were above
the regression line (G and F) are both approximately 50 km from other
large colonies, whereas the colonies below the regression line (B and T)
are both over 100 km away from other large colonies. This suggests that
foraging birds from neighbouring colonies may to some extent compete for
food.
Foraging-trip durations were recorded at the Bass Rock every year from
1998 to 2003 except 1999. In four of these years (1998 and 2001–3) we
also assessed the diets of birds at the colony, from regurgitates from adults
(Box 16.2; Hamer et al. 2000, Lewis et al. 2003). In addition, we used satel-
lite telemetry to examine the foraging ranges of birds in 1998, 2002 and
2003. Dietary data were not available for 2000 and so we used data on trip
durations and diets in 2001 as a baseline for comparison with other years.
We first used the regression of trip duration on colony size (Fig. 16.4) to
calculate an equivalent population size for each year i, based on observed
Use of gannets to monitor prey availability 245
Clupeidae were herring Clupea harengus and sprat Sprattus sprattus. Gadidae were
mainly haddock Melanogrammus aeglefinus, whiting Merlangius merlangus and cod
Gadus morhua. Other species were plaice Pleuronectes platessa, salmon Salmo salar,
trout S. trutta, grey gurnard Eutriglia gurnadus, garfish Belone belone, greater fork-
beard Phycis blennoides, dragonet Callionymus lyra and scad Trachurus trachurus.
mean trip durations, if there had been no change between years in prey
abundance or availability. We then calculated an index of relative per-capita
prey availability for each year with dietary information, as the equivalent
population size in year i expressed as a proportion of the population size in
2001 (Table 16.1). This indicated that in comparison with 2001, per-capita
prey availability was less than one-half as high in 1998, only one-quarter as
high in 2002 but very similar in 2003.
Accompanying these differences in trip duration, there was marked vari-
ation in diet between years. In particular, birds made relatively short trips
(about 160 km on average) in 2001 and 2003 and there was a high propor-
tion of sandeel (>50% by frequency) in the diet in both these years (Table
16.1). In 1998, birds made much longer trips (range = 232 km on aver-
age) and the proportion of sandeel in regurgitates was much lower than in
any other year. However, 2002 was apparently a year of very low prey avail-
ability judging from foraging-trip durations, yet the proportion of sandeel
in the diet in 2002 was more than twice that in 1998 and similar to that
in 2003, when the index of per-capita prey availability was >3 times that
246 K. C. Hamer et al.
in 2002 (Table 16.1). The mean foraging range in 2002 (303 km) was the
longest yet recorded for gannets, with the majority of trips heading north-
east of the colony and some extending as far as Bergen Bank, southwest
Norway (580 km from the Bass Rock). Hence despite the presence of a high
proportion of sandeel in the diet, sandeel availability close to the colony was
evidently low in 2002. That was also the only year when neither mackerel
nor Clupeidae (herring and sprats) made a major contribution (>25% by
frequency) to the diet (Table 16.1). In all four years, the majority of sandeels
in the diet (94% ± to 99%) were 0-group (Table 16.1).
DISCUSSION
could provide a powerful tool for gauging the importance of changes in trip
duration in terms of changes in per-capita prey availability. Further data are
now needed to determine the efficacy of this approach for other species and
different marine ecosystems. Annual variation in diet at the Bass Rock to
some extent reflected variation in trip durations, although 2002 was anoma-
lous and a combination of trip duration and diet provided a much more
complete picture than either did on their own.
ACKNOWLEDGEMENTS
Sincere thanks to Tracey Begg, Alan Bull, Jenny Bull, Francis Daunt, Cather-
ine Gray, Mike Harris, Chas Holt, Micky Maher, Linda Milne, Diana de
Palacio, Kelly Redman and Chris Roger for recording changeover rates at
gannet colonies. Thanks to Sir Hew Hamilton-Dalrymple, the Marr family,
Declan Bates, Paul Harvey (Scottish Natural Heritage), Oscar Merne and
Alyn Walsh (Duchas), the Neale family, Deryk Shaw (Fair Isle Bird Obser-
vatory Trust), Martin and Lynda Smyth, Bernie Zonfrillo, the Royal Soci-
ety for the Protection of Birds and the Scottish Seabird Centre for logis-
tical support. This work was funded by Natural Environment Research
Council CASE studentship GT4/99/55 and by grants from the European
Union (projects CEC 96-079 and Q5RS-2000-30864) and the Joint Nature
Conservation Committee (project F90-01-154).
REFERENCES
Anon. (1995). Review of sandeel biology. In Report of the ICES Workshop on Sandeel
Otolith Analysis. ICES CM 1995/9: 4. Copenhagen, Denmark: ICES.
Furness, R. W. & Ainley, D. G. (1984). Threat to Seabird Populations Presented by
Commercial Fisheries. ICBP Technical Publication 2. Cambridge, UK: ICBP,
pp. 701–708.
Furness, R. W. & Camphuysen, C. J. (1997). Seabirds as monitors of the marine
environment. ICES J. Mar. Sci., 54, 726–37.
Furness, R. W. & Tasker, M. L. (1997). Seabird consumption in sand lance MSVPA
models for the North Sea, and the impact of industrial fishing on seabird
populations. In Marine Ecosystems. Proceedings of the International Symposium on
the Role of Forage Fishes in Marine Ecosystems: Alaska Sea Grant College
Program Report Ak-SG-97–01. Fairbanks, Alaska: University of Alaska
Fairbanks, pp. 147–69.
Garthe, S., Gremillet, D. & Furness, R. W. (1999). At-sea activity and foraging
efficiency in chick-rearing northern gannets Sula bassana: a case study in
Shetland. Mar. Ecol. Prog. Ser., 185, 93–9.
Hamer K. C., Phillips, R. A., Hill, J. K., Wanless, S. & Wood, A. G. (2001).
Contrasting foraging strategies of gannets Morus bassanus at two North Atlantic
248 K. C. Hamer et al.
colonies: foraging trip duration and foraging area fidelity. Mar. Ecol. Prog. Ser.,
224, 283–90.
Hamer, K. C., Phillips, R. A., Wanless, S., Harris, M. P. & Wood, A. G. (2000).
Foraging ranges, diets and feeding locations of gannets in the North Sea:
evidence from satellite telemetry. Mar. Ecol. Prog. Ser., 200, 257–64.
Härkönen, T. (1986). Guide to the Otoliths of the Bony Fishes of the Northeast Atlantic.
Hellerup, Denmark: Danbiu ApS.
Lewis, S., Sherratt, T. N., Hamer, K. C. & Wanless, S. (2001). Evidence for
intra-specific competition for food in a pelagic seabird. Nature, 412, 816–19
Lewis, S., Benvenuti, S., Dall’Antonia, L. et al. (2002). Sex-specific foraging
behaviour in a monomorphic seabird. Proc. R. Soc. Lond. B, 269, 1687–93.
Lewis, S., Sherratt, T. N., Hamer, K. C., Harris, M. P. & Wanless, S. (2003).
Contrasting diet quality of Northern Gannets at two colonies. Ardea, 91,
167–76.
Lewis, S., Hamer, K. C., Money, L. et al. (2004). Brood neglect and contingent
foraging behaviour in a pelagic seabird. Behav. Ecol. Sociobiol., 56, 81–8.
Monaghan, P. (1992). Seabirds and sandeels: the conflict between exploitation and
conservation in the northern North Sea. Biodiversity Conserv., 1, 98–111.
Montevecchi, W. A. (1993). Birds as indicators of change in marine prey stocks. In
Birds as Monitors of Environmental Change, eds. R. W. Furness & J. J. D.
Greenwood. London: Chapman and Hall, pp. 217–65.
Nelson, J. B. (1978). The Sulidae: Gannets and Boobies. Oxford, UK: Aberdeen
University Press.
Wanless, S. & Harris, M. P. (2004). Northern gannet Morus bassanus. In Seabird
Populations of Britain and Ireland. Results of the Seabird 2000 Census
(1998–2002) eds. P. I. Mitchell, S. F. Newton, N. Ratcliffe & T. E, Dunn.
London: T. & A. D. Poyser, pp. 115–127.
Watt, J., Pierce, G. J. & Boyle, P. R. (1997). Guide to the Identification of North Sea
Fish using Premaxillae and Vertebrae. ICES Co-operative Research Report 220.
Copenhagen, Denmark: ICES, pp. 1–231.
17
Population dynamics of Antarctic krill Euphausia
superba at South Georgia: sampling with
predators provides new insights
K . R E I D , E . J . M U R P H Y, J . P . C R O XA L L
A N D P . N . T R AT H A N
A SUITABLE SAMPLER
During January 1994, a period of low krill biomass and very poor repro-
ductive success of krill-dependent predators at South Georgia (Croxall et al.
1999), the krill population structure in simultaneous samples from scien-
tific nets and land-based krill predators both showed a distinctly bimodal
distribution (Reid et al. 1999). However, samples from Antarctic fur seals,
in the weeks prior to and after this period, revealed a progressive change
Population dynamics of Antarctic krill 253
03-Jan-94
10-Jan-94
18-Jan-94
24-Jan-94 30 38 46 54 62
31-Jan-94
06-Feb-94
30 38 46 54 62
Krill length (mm) Krill length (mm)
Antarctic fur seal Scientific nets
from a dominance of large krill with a complete absence of small krill; the
gradual appearance of small krill producing a short period of bimodality
(when the net samples were collected) followed by a complete dominance
of small krill with no large krill present (Fig. 17.2). In contrast to this change
in the krill population structure, the modal size of krill remained more or
less constant throughout the predator breeding seasons of 1993 and 1995,
when predator reproductive success (and by inference krill abundance) was
considered ‘normal’ (Reid et al. 1999). These findings provided the first
evidence that distinct changes in the population structure of krill accompa-
nied periods of low krill biomass and, in turn, suggested that the population
dynamics of krill is a major factor in driving the inter-annual variability
in abundance. This pattern of change in the length–frequency distribu-
tion of krill provided a fundamental insight into the population dynamics
254 K. Reid et al.
55
53
51
Fig. 17.3 The mean length of Antarctic krill in the diet of Antarctic fur seals
during the first 14 days of March 1990–2003.
Information from the diets of Antarctic fur seals was used to examine the
population dynamics of Antarctic krill by Murphy and Reid (2001); they
showed that the absence of a single age class of krill could indeed produce
the changes in the length–frequency distribution observed in the diet of
predators. Moreover, the analysis suggested that the observed size compo-
sition of krill in samples from both predators and nets was consistent with a
relatively high rate of krill mortality at South Georgia compared with other
areas in the Scotia Sea. This increased rate of mortality also means that the
changes in length–frequency distribution associated with the failure of a
single age class to recruit into the population would also be consistent with
the observed reduction in biomass (cf. Priddle et al. 1988).
The growth of Antarctic krill is thought to be highly seasonal and the
work of Mackintosh (1974) indicated a relatively rapid increase in the mean
Population dynamics of Antarctic krill 255
(a)
10 20 30 40 50 60 70
(b)
10 20 30 40 50 60 70
(c)
10 20 30 40 50 60 70
(d)
10 20 30 40 50 60 70
Krill length (mm)
The relatively high growth and mortality rates at South Georgia places an
increased importance on the role of recruitment variability at South Geor-
gia – particularly the occurrence of years of recruitment failure – in driving
ecosystem variability. Since the analysis of population dynamics across the
Scotia Sea indicated a large-scale concordance in the pattern of recruitment,
this suggests a link between recruitment and some large-scale forcing fac-
tor. To further investigate any such environment–krill recruitment relation-
ship would require a quantifiable index of recruitment. Although no such
index exists, at least at South Georgia, the index of mean krill length in the
diet of Antarctic fur seals in March (Fig. 17.3) provides a relative index of
recruitment insofar as it indicates periods of recruitment failure.
A relationship between extensive sea-ice cover during the winter and
subsequent high recruitment of krill in the Antarctic Peninsula region has
been suggested by several authors (Loeb et al. 1997, Hewitt et al. 2003). In
addition, Trathan et al. (2003) found a relationship between krill biomass
and sea-surface temperature at South Georgia over the period 1997–9.
Therefore, since sea-ice does not reach as far north as South Georgia,
we have used an index of sea-surface temperature to examine the poten-
tial environment–recruitment relationship at South Georgia. A preliminary
examination of the relationship between the index of krill length and the
sea-surface temperature at the start of the summer suggested that the years
in which there was a recruitment failure were those of high sea-surface tem-
perature; or at least those years at the high point of the sea-surface temper-
ature cycle (Fig. 17.5).
Sea-surface temperature is frequently used as a proxy for physical
changes in the marine environment as it is relatively easy to measure, par-
ticularly using remote systems such as satellites (Yuan & Martinson 2000).
It is important to recognize that in a correlational analysis such as this we
are not necessarily suggesting that warm sea-surface temperatures cause
a krill recruitment failure; rather, there may be some other process that
causes recruitment failure – but sea-surface temperature is a suitable proxy
for this process. Indeed the time series shown in Fig. 17.5 suggests that
the relationship between sea-surface temperature and recruitment was less
well defined during the period 1999–2003. Similarly an analysis of the
258 K. Reid et al.
55 2.50
53
51 2.00
SST (°C)
45
43 1.00
41
39 0.50
37
35 0.00
1989 1990 1991 1992 1993 1994 1995 1996 1997 1998 1999 2000 2001 2002 2003 2004
Year
Fig. 17.5 The mean length of Antarctic krill in the diet of Antarctic fur seals
during the first 14 days of March 1990–2003 and the mean sea-surface
temperature (SST) at South Georgia (SST data are from Reynolds et al. 2002).
PREDATORS AS SAMPLERS
ACKNOWLEDGEMENTS
We thank all of the British Antarctic Survey staff who have collected, sorted
and measured krill in the diet of predators at Bird Island and to Drs Jon
Watkins and Bill Montevecchi, and Professor Ian Boyd for their advice and
encouragement.
REFERENCES
Boyd, I. L., Lunn, N. J. & Barton, T. (1991). Time budgets and foraging
characteristics of lactating Antarctic fur seals. J. Anim. Ecol., 60, 577–92.
Brierley, A. S., Watkins, J. L., Goss, C., Wilkinson, M. T. & Everson, I. (1999).
Acoustic estimates of krill density at South Georgia, 1981 to 1998. CCAMLR
Sci., 6, 47–57.
Broeke, M. R. van den (2000). On the interpretation of Antarctic temperature
trends. J. Clim., 13, 3885–9.
Croxall, J. P., McCann, T. S., Prince, P. A. & Rothery, P. (1988). Reproductive
performance of seabirds and seals at South Georgia and Signy Island, South
Orkney Islands, 1976–1987: implications for Southern Ocean monitoring
studies. In Antarctic Ocean and Resources Variability, ed. D. Sahrhage. Berlin:
Springer-Verlag, pp. 261–85.
Croxall, J. P., Reid, K. & Prince, P. (1999). Diet, provisioning and productivity
responses of marine predators to differences in availability of Antarctic krill.
Mar. Ecol. Prog. Ser., 177, 115–31.
Everson, I., Kock, K. H. & Parkes, G. (1997). Interannual variation in condition of
the mackerel icefish. J. Fish Biol., 51, 146–54.
Harmer, S. F. (1931). Southern whaling. Proc. Linn. Soc. Lond., 142, 85–163.
Hewitt, R. P., Demer, D. A. & Emery, J. H. (2003). An 8-year cycle in krill biomass
density inferred from acoustic surveys conducted in the vicinity of the South
260 K. Reid et al.
Reid, K., Watkins, J. Croxall, J. & Murphy, E. (1999). Krill population dynamics at
South Georgia 1991–1997, based on data from predators and nets. Mar. Ecol.
Prog. Ser., 117, 103–14.
Reid, K., Murphy, E. J., Loeb, V. & Hewitt, R. P. (2002). Krill population dynamics
in the Scotia Sea: variability in growth and mortality within a single population.
J. Mar. Syst., 36, 1.
Reynolds, R. W., Rayner, N. A., Smith, T. M., Stokes, D. C. & Wang, W. Q. (2002).
An improved in situ and satellite SST analysis for climate. J. Clim., 15, 1609–25.
Siegel, V. (1987). Age and growth of Antarctic Euphausiacea (Crustacea) under
natural conditions. Mar. Biol., 96, 483–95.
Staniland, I. J. (2002). Investigating the biases in the use of hard prey remains to
identify diet composition using Antarctic fur seals (Arctocephalus gazella) in
captive feeding trials. Mar. Mamm. Sci., 18, 223–43.
Trathan, P. N., Brierley, A. S., Brandon, M. A. et al. (2003). Oceanographic
variability and changes in Antarctic krill (Euphausia superba) abundance at
South Georgia. Fish. Oceanogr., 12, 569.
Watkins, J. L., Murray, A. W. A. & Daly, H. I. (1999). Variation in the distribution
of Antarctic krill Euphausia superba around South Georgia. Mar. Ecol. Prog. Ser.,
188, 149–60.
Yuan, X. J. & Martinson, D. G. (2000). Antarctic sea ice extent variability and its
global connectivity. J. Clim., 13, 1697–717.
18
The functional response of generalist predators
and its implications for the monitoring
of marine ecosystems
C . A S S E B U R G , J . H A R W O O D , J . M AT T H I O P O U L O S
AND S. SMOUT
0.6
0.5
0.4
Consumption rate
0.3
0.2
0.1
0.0
0 20 40 60 80 100
Focal prey density
Fig. 18.1
0.6
0.5
0.4
Consumption rate
0.3
0.2
0.1
0.0
0 20 40 60 80 100
Focal prey density
Fig. 18.2
268 C. Asseburg et al.
0.8
0.6
Consumption rate
0.4
0.2
0
0 20 40 60 80 100
Focal prey density
Fig. 18.3
20 000
19 000
Total energy intake (kJ)
18 000
17 000
16 000
15 000
14 000
0 500 1000 1500 2000
Vole density (individuals km −2)
Fig. 18.4
18000
17000
16500
16000
15500
15000
0 500 1000 1500 2000
Vole density (individuals km−2)
Fig. 18.5
26000
24000
Total energy intake (kJ)
22000
20000
18000
16000
14000
12000
0 500 1000 1500 2000
Vole density (individuals km−2)
Fig. 18.6
Figure 18.6 shows the 80% credibility interval around the rela-
tionship between mean energy intake and prey density from Exam-
ple 1. When we take account of this uncertainty, the positive rela-
tionship between vole density and total energy intake is no longer so
clear.
the density of a prey species. The predicted energy intake derived from
this MSFR does show the hoped-for simple monotonic relationship to the
density of a focal prey species when other prey species are rare or absent
(Example 1 in Box 18.3). However, when alternative prey are more abundant
(Example 2 in Box 18.3), there may be virtually no relationship between total
energy intake and prey density. In addition, because the MSFR was derived
using the relatively sparse data typical of field situations, the uncertainty
in our knowledge of this relationship results in imprecise estimates of prey
density, even when energy intake and prey density should be strongly corre-
lated. Clearly, we need to know the form of the MSFR, and the uncertainty
associated with it, in order to identify when predator performance might
provide a reliable index of prey density.
272 C. Asseburg et al.
ACKNOWLEDGEMENTS
We thank Simon Thirgood and Steve Redpath for sharing with us their data
and insights on hen harrier foraging, and Carmen Fernández for statistical
advice.
Christian Asseburg and Sophie Smout were supported by studentships
from the Scottish Higher Education Funding Council and the Natural Envi-
ronment Research Council respectively.
REFERENCES
Aqorau, T. (2003). Obligations to protect marine ecosystems under international
conventions and other legal instruments. In Responsible Fisheries in the Marine
Ecosystem, eds. M. Sinclair & G. Valdimarsson. Wallingford, UK: CABI
Publishing, pp. 25–41.
Asseburg, C., (2005). Modelling uncertainty in multi-species predator–prey
interactions. Unpublished Ph.D. thesis, University of St Andrews.
Beddington, J., Hassell, M. P. & Lawton, J. H. (1976). The components of arthropod
predation. II. The predator rate of increase. J. Anim. Ecol., 45, 165–86.
CCAMLR (1982). Convention on the Conservation of Antarctic Marine Living
Resources. http://www.ccamlr.org/pu/e/pubs/bd/pt1p2.htm.
(1985). CCAMLR Ecosystem Monitoring Program: Standard Methods (revised
2004). http://www.ccamlr.org/pu/e/pubs/std-meth04.pdf
Constable, A. J. (2001). The ecosystem approach to managing fisheries: achieving
conservation objectives for predators of fished species. CCAMLR Sci., 8, 37–64.
Constable, A. J., de la Mare, W. K., Agnew, D. J., Everson, I. & Miller, D. (2000).
Managing fisheries to conserve the Antarctic marine ecosystem: practical
implementation of the Convention on the Conservation of Antarctic Marine
Living Resources (CCAMLR). ICES J. Mar. Sci., 57, 778–91.
Croxall, J. P., Reid, K. & Prince, P. (1999). Diet, provisioning and productivity
responses of marine predators to differences in availability of Antarctic krill.
Mar. Ecol. Prog. Ser., 177, 115–31.
274 C. Asseburg et al.
In recent years, enormous effort has been expended to explain the cause
of the precipitous decline of the western population of Steller sea lions
(Eumatopias jubatus) since the late 1970s; however, despite these efforts and
the proposal of a wide variety of hypotheses, the decline has proven to be
very difficult to explain. The authors of a recent comprehensive review of the
problem emphasized repeatedly that the system is in dire need of a mod-
elling approach that takes advantage of the data available at small spatial
scales (at the level of the rookery). We view this as an opportunity for eco-
logical detection, a process in which multiple hypotheses simultaneously
compete and their success is arbitrated by the relevant data. We describe
ten hypotheses for which there are sufficient data to allow investigation, a
method that allows one to link various sources of data to the hypotheses and
the conclusions from this approach.
The decline of the western Alaska population of Steller sea lions has
proven to be very difficult to explain, in part because most aspects of the
population and the environmental variables proposed to explain its decline
involve a combination of high spatial and temporal variability, and limited
data. Consequently, most researchers pooled data across rookeries or across
time, obscuring spatial and/or temporal patterns (Fig. 19.1). Some of these
previous studies are described below.
(1) Construction of a Leslie matrix model for a stable population, followed
by perturbation of various transition rates to find the most
parsimonious way to produce a trajectory matching the observed
decline. York (1994) determined that the initial decline could be
(a)
120000
100000
80000
Population
60000
40000
20000
0
1975 1980 1985 1990 1995 2000 2005
Year
(b)
2002 8506
6512
4785
3323
2127
1196
532
133
0
No data
Year
1973
West SSL rookeries East
Fig. 19.1 (a) Composite time series of the Steller sea lion (SSL) population in
western Alaska. (b) Space–time plot of the counts of non-pups at 38 individual
rookeries from 1973 to 2002. The abscissa indicates rookeries from west to east,
with each column representing a different rookery. The ordinate indicates time,
with each row representing a single year. The area of each circle indicates the
observed number of sea lions at that rookery in that year. The arrows indicate
years in which a ‘synoptic’ survey of the entire population was taken. Notice,
however, that the dataset is much richer than the synoptic survey (panel a, with
10 points total) would suggest.
The method of multiple hypotheses 277
None of these studies made use of both spatial and temporal variation in sea
lion counts and environmental data. Researchers either pooled data across
space, combining all rookery censuses within each year and performing
their analyses using a composite time series representing the entire west-
ern population (Fig. 19.1); or else pooled data across time, treating only the
overall trend of the census at each rookery. A recent review of the problem
(NRC 2003) emphasized that the system requires a modelling approach that
takes advantage of the data available at small spatial scales. Here we give a
précis of the results of such a study (Wolf & Mangel 2004)
RELEVANT FISHERIES
Walleye pollock, Atka mackerel and Pacific cod are harvested primarily
using groundfish trawling gear. The largest groundfish harvests in the area
occur in the Bering Sea. The peak catch occurred in 1972, followed by a
decline in the late 1970s and a recovery in the mid 1980s. Pollock comprise
over 76% of the groundfish caught in the Bering Sea (NRC 2003). In the
Gulf of Alaska, the pollock catch peaked between 1976 and 1985. A fishery
for groundfish developed in the Aleutian Islands in the late 1970s.
The decline of the Steller sea lion was preceded by declines in populations
of northern fur seal (Callorhinus ursinus) and Pacific harbour seal (Phoca
vitulina) occupying the same region. The causes of these declines remain
similarly unexplained (Merrick 1997, Springer et al. 2003). The range of the
western population is also home to a large population of killer whales (Orci-
nus orca); some of these are ‘transients’, whose diet is thought to consist
mainly of marine mammals, including Steller sea lions (Barrett-Lennard
et al. 1995, Matkin et al. 2002). Unfortunately, very little is known about the
spatial or temporal distribution of these whales. Steller sea lions may com-
prise 5% to 20% of their diet (Matkin et al. 2002). The stomach of one killer
whale that washed up on a beach in British Columbia contained flipper tags
from 14 different Steller sea lion pups, all of which had been tagged at the
Marmot Island rookery 3 to 4 years before (Saulitis et al. 2000).
COMPETING HYPOTHESES
would be a bigger problem for naı̈ve pups (Loughlin et al. 1983), whereas
adults are more likely to be targeted by shooters.
Transient killer whales are predicted to prey upon sea lions or their pups
when the whales’ preferred prey, harbour seals, are scarce (Springer et al.
2003, Mangel & Wolf in press). Therefore, survival probability of sea lion
pups (H9) or non-pups (H10) is predicted to decline if local harbour seal
density falls below a threshold value.
THE APPROACH
The hypotheses and data were linked by the following procedure (all the
details can be found in Wolf and Mangel (2004)).
for the numbers of non-pups (N) and pups (J) at time t, given the numbers
at time t − 1, are computed directly from the associated binomial distribu-
tions.
Although pups can be accurately counted during aerial surveys, some
fraction of the non-pups are likely to be at sea foraging during the surveys.
Thus, the observed number of non-pups is smaller than the true number
of non-pups. We characterize this observation uncertainty using a beta-
binomial distribution (Martz & Waller 1982, Evans et al. 2000, Wolf &
Mangel 2004).
We set fixed background values for the life-history parameters (denoted
by ρ 0 , σ 0 and φ 0 ) that are modified by local conditions to reflect a particular
hypothesis. We chose background values from observed values measured
on the Marmot Island rookery by Calkins and Pitcher (1982) and used by
York (1994), and Pascual and Adkison (1994). The annual growth rate of a
population using these values is 0.4% (Pascual & Adkison 1994). Fecundity
estimates also account for only about 50% of the pups being female (York
1994, NRC 2003) and for about 50% of the non-pup population being juve-
nile (Holmes & York 2003).
The background values for life-history parameters are then modified
to take account of local conditions. For the case of food limitation (H1 to
H3), the modification is that low abundance of groundfish and other prey
causes local fecundity (hypothesis 1), pup recruitment (hypothesis 2), or
non-pup survival probability (hypothesis 3) to be diminished. These are
282 N. Wolf et al.
Time
Fig. 19.2
1 1
(a) (b)
Multiplier
Multiplier
c = 400 000 c = 0.04
c = 2 000 000 c = 0.2
c = 10 000 000 c=1
c = 50 000 000 c=5
0 0
Prey density Pollock fraction
1 1
(c) (d)
Multiplier
Multiplier
0 0
Fishing activity Harbour seal density
Fig. 19.3
rising and falling populations should tend to be above and below the criti-
cal value respectively.)
Thus, the local value of each vital rate is calculated by multiplying the
background rate by all the relevant scaling factors. Note that there are ten
unknown parameters, one per hypothesis. In each case, a parameter value
of zero indicates that the corresponding hypothesis has no effect.
SOURCES OF DATA
1980, Everitt & Braham 1980, Pitcher 1990, Frost et al. 1999, Mathews &
Pendleton 2000, Jemison & Kelly 2001, Boveng et al. 2003, Small et al.
2003). The Steller sea lion counts were from the NMFS/AFSC/National
Marine Mammal Laboratory (NMML) online database (http://nmml.
afsc.noaa.gov/AlaskaEcosystems/sslhome/stellerhome.html). We limited
our consideration to year–rookery combinations in which counts from June
or July were available for both pups and non-pups, and to rookeries for
which such censuses from at least two different years were available. When
more than one count was available for a particular rookery in a single year,
we took the average. Several sets of adjacent rookeries were censused as one
large rookery early in the dataset and as separate rookeries in later years. In
some of these cases, we combined the counts from the separate rookeries
in later years in order to extend the time series for the ‘joint’ rookery. When
prey abundance or harbour seal density estimates were missing for certain
area–year combinations, we used linear interpolation to estimate the miss-
ing value from reported values in earlier and later years for the same area.
Further details about the acquisition and treatment of data are found in Wolf
and Mangel (2004).
RESULTS
0 0
Fecundity
1 4
0 0 0 0
Recruitment
2 5 7 9
0 0 0 0
Survival
3 6 8 10
Food Pollock Boats Seals
Fig. 19.4 Likelihood profiles for the parameters associated with the functional
forms for the four classes of hypotheses and their effects on the relevant
life-history parameters. See text for further details.
DISCUSSION
The strong message of ecology is that the world changes and that the rea-
sons for change are manifold. Thus, rather than trying to ‘prove’ one mecha-
nism, we should recognize that multiple mechanisms will almost always be
at work, and we should ask how to weigh the importance of different mech-
anisms. It is this approach that we have taken in understanding the decline
The method of multiple hypotheses 287
(a) 1 (b) 1
Relative fecundity
Relative fecundity
c1 = 900000 c4 = 0.04
0 0
0 10 000000 20000000 30000000 0 0.2 0.4 0.6 0.8 1
'Food' (CPUE of all prey species in kg km−3) Pollock fraction (by CPUE weight)
(c) 1 (d) 1
Relative pup recruitment
0 0
0 0.2 0.4 0.6 0.8 1 0 5000 10000 15000 20000
Pollock fraction (by CPUE weight) Harbour seals within 300 km
Fig. 19.5 Vital rate functions corresponding to MLE parameter values. (a)
Hypothesis 1: food availability affects fecundity (magnitude of effect: medium).
(b) Hypothesis 4: pollock fraction affects fecundity (magnitude of effect: weak at
best). (c) Hypothesis 5: pollock fraction affects pup recruitment (magnitude of
effect: very strong). (d) Hypothesis 10: harbour seals (via predation) affect
non-pup survival (magnitude of effect: weak but persistent).
of the western population of Steller sea lions. There is good evidence for
two strong effects: H1, total prey availability affects fecundity; H5, pollock
fraction in the environment affects pup recruitment. One moderate effect
was found: H10, harbour seal density (predation) affects non-pup survival.
There was also marginal evidence for one weak effect: H4, pollock fraction
in the environment affects fecundity. No evidence was found for any of the
other hypotheses. What our work has done is to guide the weight of the evi-
dence, when all plausible hypotheses are competing, towards those that win
the competition.
Although we used the word mechanism, we recognize that a study such
as this one cannot demonstrate causality. It would seem that H1 is a rela-
tively clear and simple mechanism: lower abundance of all prey types leads
to lower fecundity through the direct effect of reduced resource accumu-
lation by adults and thus reduced storage for reproduction. H10 also has
288 N. Wolf et al.
2000–01 0.16
0.14
0.12
0.10
0.08
0.06
0.04
0.02
0
No data
Year
1980–81
West SSL rookeries East
Fig. 19.6 Lost production of the Steller sea lion population due to hypotheses 1,
4, 5 and 10. The diameter of each circle represents the fraction of potential
production lost.
a clear mechanism, but note that its MLE is about 0.01, so that the effect
of changes in the breadth of the diet of killer whales leads to only a 1%
reduction in non-pup survival, and then only in cases where harbour seal
numbers are sufficiently low.
On the other hand, H4 and H5 are more complicated. A high pollock
fraction can result either from high pollock or from low non-pollock, and
either of these could be the underlying factor. Furthermore, the mechanism
might be something completely different for which pollock fraction is only
a correlate. For example, juveniles may require some easily caught subset
of prey species because they are unable to dive deep enough or swim fast
enough to catch anything else. (In the current dataset, the fish biomass is
not broken down by size class of fish. However, the really small fish and the
really big ones are probably not useful to sea lions. Thus, some additional
thinking is required about how to modify the survey data to address this
question.)
Our results also suggest an adaptive management plan in which one
designates the areas around some of the rookeries as experimental zones
in which to make fishery quotas contingent upon the results of pre-fishing-
season survey trawls. We envision a series of treatments:
ACKNOWLEDGEMENTS
This research was supported in part by the Alaska Fisheries Science Cen-
ter of NOAA Fisheries, Department of Commerce under Contract AB133F-
02-CN-0085: ‘Alternative Management Strategies for the Recovery of the
Stellar Sea Lion’; and in part by the Center for Stock Assessment Research,
University of California Santa Cruz. We thank Dan Goodman and Dan
Hennen for providing the commercial fishery and NMFS trawl data. For
various helpful discussions we thank Doug DeMaster, Anne York, Lowell
290 N. Wolf et al.
Fritz, Tim Ragen, Gunnar Steffanson, Jim Estes and the entire Mangel lab-
oratory, but especially Steve Munch. We appreciate the use of a computer
from the laboratory of Ingrid Parker.
REFERENCES
Adkison, M. D., Pascual, M. A., Hilborn, R. et al. (1993). Modeling the trophic
relationships between fish and marine mammal populations in Alaskan
waters. In Is it food? Addressing Marine Mammal and Seabird Declines, ed. S.
Keller. Fairbanks, Alaska: University of Alaska Sea Grant College Program,
pp. 54–6.
Alverson, D. L. (1992). A review of commercial fisheries and the Steller sea lion
(Eumetopias jubatus): the conflict arena. Rev. Aquat. Sci., 6, 203–56.
Anderson, P. J. & Blackburn, J. E. (2002). Status of demersal and epibenthic
species in the Kodiak Island and Gulf of Alaska region. In Steller Sea Lion
Decline: Is it Food II, eds. D. DeMaster & S. Atkinson. Fairbanks, Alaska:
University of Alaska Sea Grant College Program, pp. 57–60.
Andrews, R. D., Calkins, D. G., Davis, R. W. et al. (2002). Foraging behavior and
energetics of adult female Steller sea lions. In Steller Sea Lion Decline: Is it Food
II, eds. D. DeMaster & S. Atkinson. Fairbanks, Alaska: University of Alaska Sea
Grant College Program, pp. 19–22.
Angliss, R. P. & Lodge, K. L. (2002). Alaska Marine Mammal Stock Assessments,
2002. NOAA Technical Memorandum NMFS AFSC-133. Seattle, WA: US
Department of Commerce.
Bailey, E. P. & Faust, N. H. (1980). Summer distribution and abundance of marine
birds and mammals in the Sandman Reefs, Alaska, USA. Murrelet, 61, 6–19.
Barrett-Lennard, L. G., Heise, K. A., Saulitis, E., Ellis, G. & Matkin, C. (1995). The
Impact of Killer Whale Predation on Steller Sea Lion Populations in British
Columbia and Alaska. Vancouver, BC: North Pacific Universities Marine
Mammal Research Consortium.
Bickham, J. W., Loughlin, T. R., Calkins, D. G., Wickliffe, J. K. & Patton, J. C.
(1998). Genetic variability and population decline in Steller sea lions from the
Gulf of Alaska. J. Mammal., 79, 1390–5.
Blackburn, C. (1990). Sea Lion Briefing Paper. Kodiak, Alaska: Alaska Groundfish
Data Bank.
Boveng, P. L., Bengtson, J. L., Withrow, D. E. et al. (2003). The abundance of
harbor seals in the Gulf of Alaska. Mar. Mammal Sci., 19, 111–27.
Burnham, K. P. & Anderson, D. R. (1998). Model Selection and Inference: A Practical
Information-Theoretic Approach. New York: Springer-Verlag.
Calkins, D. & Pitcher, K.W. (1982). Population assessment, ecology and trophic
relationships of Steller sea lions in the Gulf of Alaska. In Environmental
Assessment of the Alaskan Continental Shelf, Final Report. Washington, DC: US.
Department of Commerce and US Department of the Interior, pp. 447–546.
Calkins, D. G., Becker, E. F. & Pitcher, K. W. (1998). Reduced body size of female
Steller sea lions from a declining population in the Gulf of Alaska. Mar.
Mamm. Sci., 14, 232–44.
Castellini, M. (1993). Report of the Marine Mammal Working Group. In Is it Food?
Addressing Marine Mammal and Seabird Declines, ed. S. Keller. Fairbanks,
Alaska: University of Alaska Sea Grant College Program, pp. 4–11.
The method of multiple hypotheses 291
Evans, M., Hastings, N. & Peacock B. (2000). Statistical Distributions, 3rd edn. New
York: John Wiley.
Everitt, R. D. & Braham, H. W. (1980). Aerial survey of pacific harbor seals, Phoca
vitulina richardsi, in the southeastern Bering Sea. Northw. Sci., 54, 281–8.
Frost, K. J., Lowry, L. F. & Ver Hoef, J. M. (1999). Monitoring the trend of harbor
seals in Prince William Sound, Alaska, after the Exxon Valdez oil spill. Mar.
Mamm. Sci., 15, 494–506.
Hilborn, R. & Mangel, M. (1997). The Ecological Detective: Confronting Models with
Data. Princeton, NJ: Princeton University Press.
Holling, C. S. (1959). Some characteristics of simple types of predation and
parasitism. Can. Entomol., 91, 385–98.
Holmes, E. E. & York, A. E. (2003). Using age structure to detect impacts on
threatened populations: a case study with Steller sea lions. Conserv. Biol., 17,
1794–806.
Hoover-Miller, A. A. (1994). Harbor Seal (Phoca vitulina) Biology and Management
in Alaska. Washington, DC: Marine Mammal Commission.
Jemison, L. A. & Kelly, B. P. (2001). Pupping phenology and demography of harbor
seals (Phoca vitulina richardsi) on Tugidak Island, Alaska. Mar. Mamm. Sci., 17,
585–600.
Loughlin, T. R. & York, A. E. (2002). An accounting of the sources of Steller sea
lion mortality. In Steller Sea Lion Decline: Is it Food II, eds. D. DeMaster & S.
Atkinson. Fairbanks, Alaska: University of Alaska Sea Grant College Program,
pp. 9–13.
Loughlin, T. R., Consiglieri, L., DeLong, R. L. & Actor, A. T. (1983). Incidental
catch of marine mammals by foreign fishing vessels, 1978–81. Mar. Fish. Rev.,
45, 44–9.
Mangel, M. & Wolf, N. Predator diet breadth and prey population dynamics:
mechanism and modeling. In Whales, Whaling and Ocean Ecosystems, ed. J.
Estes. Berkeley, CA: UC Press, in press.
Martz, H. F. & Waller, R. A. (1982). Bayesian Reliability Analysis. New York: John
Wiley.
Mathews, E. A. & Pendleton, G.W. (2000). Declining Trends in Harbor Seal (Phoca
vitulina richardsi) Numbers at Glacial Ice and Terrestrial Haulouts in Glacier Bay
National Park, 1992–1998. Final Report to Glacier Bay National Park and
Preserve, Resource Management Division. Cooperative Agreement
9910–97–0026. Bedford, VA: National Park Service, US Department of the
Interior.
Matkin, C. O., Barrett-Lennard, L. G. & Ellis, G. (2002). Killer whales and
predation on Steller sea lions. In Steller Sea Lion Decline: Is it Food II, eds. D.
DeMaster & S. Atkinson. Fairbanks, Alaska: University of Alaska Sea Grant
College Program, pp. 61–6.
Merrick, R. L. (1997). Current and historical roles of apex predators in the Bering
Sea ecosystem. J. Northw. Atl. Fish. Sci., 22, 343–55.
Merrick, R. L., Chumbley, M. K. & Byrd, G. V. (1997). Diet diversity of Steller sea
lions (Eumetopias jubatus) and their population decline in Alaska: a potential
relationship. Can. J. Fish. Aquat. Sci., 54, 1342–8.
NRC (National Research Council) (2003). Decline of the Steller Sea Lion in Alaskan
Waters: Untangling Food Webs and Fishing Nets. Washington, DC: National
Academy of Sciences Press.
292 N. Wolf et al.
Pascual, M. A. & Adkison, M. D. (1994). The decline of the Steller sea lion in the
northeast Pacific: demography, harvest, or environment? Ecol. Applic., 4,
393–403.
Perez, M. A. & Loughlin, T. R. (1991). Incidental Catch of Marine Mammals by
Foreign and Joint Venture Trawl Vessels in the U.S. EEZ of the North Pacific,
1973–88. Seattle, WA: US Department of Commerce, National Oceanographic
and Atmospheric Administration, National Marine Fisheries Service.
Pitcher, K. W. (1990). Major decline in the number of harbor seals, Phoca vitulina
richardsi, on Tugidak Island, Gulf of Alaska. Mar. Mammal Sci., 6, 121–34.
Pitcher, K. W., Calkins, D. G. & Pendleton, G. W. (1998). Reproductive
performance of female Steller sea lions: an energetics-based reproductive
strategy? Can. J. Zool., 76, 2075–83.
Raum-Suryan, K. L., Pitcher, K. W., Calkins, D. G., Sease, J. L. & Loughlin, T. R.
(2002). Dispersal, rookery fidelity, and metapopulation structure of Steller sea
lions (Eumetopias jubatus) in an increasing and decreasing population in
Alaska. Mar. Mammal Sci., 18, 746–64.
Rosen, D. A. S. & Trites, A. W. (2000). Pollock and the decline of Steller sea lions:
testing the junk-food hypothesis. Can. J. Zool., 78, 1243–50.
Saulitis, E. L., Matkin, C. O., Heise, K. A., Barrett-Lennard, L. G. & Ellis, G. M.
(2000). Foraging strategies of sympatric killer whale (Orcinus orca) populations
in Prince William Sound, Alaska. Mar. Mammal Sci., 16, 94–109.
Sease, J. L. & Loughlin, T. R. (1999). Aerial and Land-based Surveys of Steller Sea
Lions (Eumetopias jubatus) in Alaska, June and July 1997 and 1998. NOAA
Technical Memorandum NMFS-AFSC-100. Washington, DC: US Department
of Commerce.
Small, R. J., Pendleton, G. W. & Pitcher, K. W. (2003). Trends in abundance of
Alaska harbor seals, 1983–2001. Mar. Mammal Sci., 19, 344–62.
Springer, A. M., Estes, J. A., Vliet, G. B. van et al. (2003). Sequential megafaunal
collapse in the North Pacific Ocean: an ongoing legacy of industrial whaling?
Proc. Natl. Acad. Sci. U.S.A., 100, 12 223–8.
Withrow, D. E., Cesarone, J. C., Jansen, J. K. & Bengston, J. L. (2000). Abundance
and distribution of harbor seals (Phoca vitulina) along the Aleutian Islands
during 1999. In Marine Mammal Protection Act and Endangered Species Act
Implementation Program 1999, eds. A. L. Lopez & D. P. DeMaster. AFSC
Processed Report 2000–11. Seattle, WA: Alaska Fisheries Science Center,
pp. 91–115.
(2001). Abundance and distribution of harbor seals (Phoca vitulina richardsi) in
Bristol Bay and along the north side of the Alaska peninsula during 2000. In
Marine Mammal Protection Act and Endangered Species Act Implementation
Program 2000, eds, A. L. Lopez & R. P. Angliss. AFSC Processed Report
2001–06. Seattle, WA: Alaska Fisheries Science Center.
Withrow, D. E., Cesarone, J. C., Hiruki-Raring, L. & Bengston, J. L. (2002).
Abundance and distribution of harbor seals (Phoca vitulina) in the Gulf of
Alaska (including the south side of the Alaska Peninsula, Kodiak Island, Cook
Inlet and Prince William Sound) during 2001. In Marine Mammal Protection
Act and Endangered Species Act Implementation Program 2001, eds. A. L. Lopez &
S. E. Moore. AFSC Processed Report 2002–06. Seattle, WA: Alaska Fisheries
Science Center.
The method of multiple hypotheses 293
Wolf, N. & Mangel, M. (2004). Understanding the Decline of The Western Alaskan
Steller Sea Lion: Assessing the Evidence Concerning Multiple Hypothesis. Report of
MRAG Americas. Tampa, FL.: MRAG Americas.
http://www.soe.ucsc.edu/∼msmangel/Stellerfinal.pdf.
Figures at http://www.soe.ucsc.edu/∼msmangel/FiguresFinal.pdf.
York, A. E. (1994). The population dynamics of northern sea lions, 1975–1985. Mar.
Mammal Sci., 10, 38–51.
York, A. E, Merrick, R. L. & Loughlin, T. R. (1996). An analysis of the Steller sea
lion metapopulation in Alaska. In Metapopulations and Wildlife Conservation, ed.
D. R. McCullough. Washington, DC: Island Press, pp. 259–92.
20
Modelling the behaviour of individuals
and groups of animals foraging in
heterogeneous environments
J . G . O L L A S O N , J . M . Y E A R S L E Y, K . L I U A N D N . R E N
FOUNDATIONS
(1) Animals eat food, part of which is incorporated into the animal
(requirement of consumption).
(2) To maintain itself an animal utilizes body tissue, the equivalent of
stored food, and the rate of utilization is dependent on the mass of the
animal (requirement of utilization).
(3) An animal will reject food if the rate at which the food can be eaten
is less than the rate of utilization of body tissue (requirement of
hunger).
(4) By eating, an animal depletes the standing crop of food (requirement
of depletion).
(5) The rate at which an animal feeds is a function of the standing crop of
food (requirement of ingestion).
(6) The rate at which an animal feeds is a function of the density of all of
the animals feeding on the same food (requirement of competition).
rescaled to zero so that the timing of the current activity starts when the
activity begins.
where tc0 and tc1 are the steady-state times spent foraging at patch 0 and
patch 1 respectively. The model predicts that the time spent by an animal
foraging at each of the patches is proportional to the rates of regeneration
at the patch provided that the parameters r, v and tt all tend to zero.
When these parameters are increased the proportion of time spent
feeding is decreased, the proportion of time spent travelling between the
patches is increased causing the undermatching of the ideal free distribu-
tion observed by Kennedy and Gray (1993). Equations (20.7) and (20.8) pre-
dict the increasing degree of mismatching observed as the distance between
the feeding patches is increased. These results are also true of environments
containing many patches. Table 20.1 shows the results of simulating the
298 J. G. Ollason et al.
distribution that arises when 150 individuals are free to forage in an envi-
ronment containing 16 patches arranged in a 4 × 4 matrix. When it leaves,
the individual is free to move at random to any adjacent patch. Each individ-
ual applies the requirement of hunger to determine if it will stay at a patch.
For realistic values of r, v and tt the predictions approximate closely to the
ideal free distribution.
Simulation shows that a Holling Type II functional response, and allo-
metrically scaled metabolic costs, have little qualitative effect on the model
(Ollason & Yearsley 2001). Facilitation, the specific feeding rate of the
individual increasing with the number of foragers at a patch, causes indi-
viduals to travel in flocks, but the average numbers occupying the patches
approximate to the ideal free distribution nevertheless (Fig. 20.1).
It is possible to model animals with differing dominance, with dom-
inants feeding more rapidly than subordinates. Populations made up of
competitors of varying dominance divide so that dominants forage in the
company of dominants, subordinates in the company of subordinates, but
despite this the ideal free distribution is approximated (Fig. 20.2). For fur-
ther details see Ollason and Yearsley (2001). Notice that the decision mak-
ing depends only on m and r. The animal does not need to know how
many patches are present in the environment or the pay-offs at the different
patches. All it needs to do is to leave the current patch at which it is feeding
when its rate of intake is less than its total metabolic rate. All aspects of its
time budget emerge from this single rule.
The optimality of the foraging of great tits was defined by a model of con-
tinuous feeding, despite the patches containing a small number of large
Modelling foraging behaviour of individuals and groups 299
30
25
No. of animals
20
15
10
5
0
2000 3000 4000 5000 6000 7000 8000
Time
Fig. 20.1 The numbers of foragers at patch 0, a0 = 0.01 (solid line) and at
patch 1, a1 = 0.02 (dashed line). The corresponding mean values are shown by
the straight lines. The animals facilitate each other’s feeding rate; consequently,
when one animal leaves a patch, the feeding rates of all the others decrease,
causing most of the remaining animals to leave almost at once. The expected
mean proportion of feeding animals is 1/3 and 2/3 at patch 0 and at patch 1
respectively. The observed means were 0.373 and 0.626 respectively,
representing a small degree of undermatching. From Ollason and Yearsley
(2001).
30
25
Mean rank
20
15
10
6 8 10 12 14 16
No. of animals at patch 0
Fig. 20.2 Individuals with large numerical ranks can feed faster than those with
smaller ranks. When the mean rank of the occupants is large, only a small
number can occupy a patch, driving the subordinates away to the other patch.
The patches regenerated as in Fig. 20.1. The expected proportion of foragers at
patch 0 and at patch 1 is 1/3 and 2/3 respectively, the observed proportions were
0.327 and 0.673. From Ollason and Yearsley (2001).
particles (Cowie 1977). Oaten (1977) showed that the optimal foraging strat-
egy in a stochastic environment does not necessarily converge to the optimal
strategy in the corresponding deterministic environment. Iwasa et al. (1981)
proved that the optimal strategy in a stochastic environment is dependent
on the statistical distribution of the food, and that no single strategy was
300 J. G. Ollason et al.
dF
= −vF
dt
dm
= vF − r m
dt
F (0)(1 − e−v t c )
lim F (0)v e−v t c =
r →0 (v − r )
where F(0) is the mass of food at each patch when the animal arrives, and tc
is the steady-state staying time. As r → 0 the staying time, tc , converges to
that predicted by the marginal value theorem (Charnov, 1976). Feeding on
particles of food requires that three states have to be considered: feeding,
waiting to find the next particle and travelling between patches. The require-
ment of hunger is restated as follows: Leave the patch when the time spent
Modelling foraging behaviour of individuals and groups 301
waiting to find the next particle is greater than the remembered time to find
a particle. In symbolic terms:
mtw (t)
t¯w (t) =
mn (t)
where t¯w (t) is the remembered waiting time. The rule, then, is leave when
tw ≥ t¯w (t). Between patches, m, mtw and mn all decay exponentially. On
arrival at the patch the animal is waiting so m and mn continue to decay
while mtw is increasing:
m(t) = m(0)e−r t
mn (t) = mn (0)e−r t
dmtw
= 1 − r mtw
dt
(The time spent waiting passes at the rate of 1 time unit per unit time.) If
the animal encounters a particle of food it feeds. This increases both m and
mn , but while it is feeding it is not waiting so mtw decays away:
dm f
= − rm
dt th
dmn 1
= − r mn
dt th
mtw (t) = mtw (0)e−r t
where f is the mass of food in a particle and th is the handling time. A
problem with this approach is that the calculation of mn depends on th which
cannot be known before the particle has been handled. The problem can be
resolved by assuming that th is small, giving the following approximation:
mn (t) ≈ 1 + mn (0)e−r t
From the continuous case of feeding in a patch that is searched at a constant
rate and at random, the amount of food removed as a function of time is
F (t) = F (0)(1 − e−v t )
Hence
1 F (0)
t= ln
v F (0) − F (t)
and a simple discretization of this process involving np particles gives
1 np
twi = ln
v np − i
302 J. G. Ollason et al.
1.02
1.00
0.98
0.96
Matching
0.94
0.92
0.90
0.88
0.86
220
200
8 180
9 160
10 140
11 120
12 100
13
80 t s1
t a1 60
14 40
15 20
Fig. 20.3 In 16 time units 16 particles of food are added to two patches. At
constant intervals ta1 defines the ratio of the times at which the particles are
added. ta1 = 8 signifies that particles are added to each patch at the rate of
16/8 = 2 time units, ta1 = 15 signifies that at one patch a particle is added every
16 time units, and at the other, every 16/15 time units. In this way, although the
rates of addition of the particles vary the overall rate of regeneration of the
system is constant. ts1 is the time taken to find a particle when there is one
particle in the patch, the smaller the value of ts1 the faster the patch may be
searched. Perfect matching is indicated by the dashed grid. Undermatching
increases as the difference in the rates of regeneration increases, and when ts1 is
small. The undulations of the response surface are aliasing artifacts. From
Ollason and Ren (2002).
where twi is that time at which the ith particle has been removed. In the
limits r → 0, np → ∞, the equilibrium value of twi converges to the equi-
lbrium value of tc in the continuous case, and to the prediction of the
marginal value theorem (Ollason & Ren 2002). A simulation of 16 animals
foraging at 2 patches of regenerating food is shown in Fig. 20.3. The par-
ticles are added one by one to each patch such that a total of 16 particles
were added to the patches in 16 time units. The particles were added to
the patches in the ratios 15:1, 14:2, 13:3, . . . , 8:8. It is assumed that the
time to find the next particle in a patch containing n particles, ts (n) = ts1 /n.
Figure 20.3 shows that matching approximates closely to the ideal free dis-
tribution and is weakly affected by the ratios of delivery, undermatching
being more pronounced as the ratio of regeneration at the patches becomes
more disparate.
Modelling foraging behaviour of individuals and groups 303
The two models are drawn together modelling the foraging of the Common
guillemot, Uria aalge, which is represented as two interdigitated processes
of foraging at patches, involving travelling between patches of particles of
food and between continuous patches of oxygen. This work extends a model
developed by Liu (2002).
The model depends on three state variables: O, the volume of oxygen
contained in the body; g, the mass of food in the gut; and m, the body mass
(wet weight) of the animal. The environment is a patch of particles of food,
each of which the animal eats at the constant rate f. The density of particles
is not depleted. The aim of the model is to predict the time budget of the
guillemot in terms of the time spent diving and feeding, and the time spent
at the surface replenishing oxygen used in diving. The guillemot is in one
of five possible states:
Because costs of travelling involve the mass transported – not only of the
body tissue, but also the unincorporated contents of the gut – the wet weight
of gut contents is explicitly represented. The equations of the new model are
shown below.
between r d and r b is the net specific cost of moving a unit mass of material.
The residue is assumed to be the cost of maintaining the living tissue. The
specific cost, therefore, of moving the contents of the gut – which being
unassimilated, do not need to be maintained – while descending will be
r d − r b . The parameters are defined in Table 20.2.
FEEDING
dg
= f − rgg
dt
dm
= αr g g − r f m − (r f − r b )g
dt
dO
= −ρr f m − ρ(r f − r b )g
dt
The two differences between feeding and diving or ascending involve the
parameters f and r f . See Table 20.3 for parameters of feeding.
dg
= −r g g
dt
dm
= αr g g − r r m − (r r − r b )g
dt
dO
= u(Omax − O/m)m − ρr r m − ρ(r r − r b )g
dt
Modelling foraging behaviour of individuals and groups 305
a
This is the hypothetical maximum concentration of oxygen in the tissue of an
animal without metabolic costs.
It is assumed that costs of recovering and of waiting are the same so the
relevant metabolic parameter, r r , is common to both processes. Table 20.4
shows the parameters of recovery.
On diving the animal contains O(0) units of volume of oxygen. The animal
will stop feeding and return to the surface when the concentration of oxygen
contained falls to Omin . The total metabolic cost of an activity is a function of
the mass m of the animal and of the activity in which the animal is engaged.
Associated with each activity is a linear scaling factor that expresses the cur-
rent metabolic cost as a proportion of some basal metabolism. When tis-
sue is catabolized there is a stoichiometrically equivalent usage of oxygen.
The four activities to be considered are: descending, feeding, ascending and
recovering. The scaling factors of the costs are s d , s f , s a and s r . The basal
specific metabolic cost is r b , and the stoichiometric equivalent of oxygen
is ρ.
The animal has a maximum and a minimum body mass, mmax and mmin
respectively. If m ≤ mmin the animal is dead from starvation. If m ≥ mmax
the animal is replete. The contents of the gut lie in the range 0 ≤ g ≤ g max ;
and a second bound is established, g min , the gut being effectively empty.
The guillemot has to stop feeding when the gut is filled.
306 J. G. Ollason et al.
The bird has to stop feeding when the oxygen is running out, when O =
Omin m. It has to stop feeding when g = g max , or when m = mmax . The
resumption of feeding may be determined by the mass of food remaining in
the gut, for example, when g = g min ; it may also be determined by the mass
of oxygen in the tissue. For example, the animal might choose to dive again
when it contains the local maximum volume of oxygen obtained while it is
at the surface; or alternatively it might choose to dive again when the body
mass falls to mmax . The bird will dive when g = g min or when dO/dt = 0.
As noted above, the bird will stop feeding when m = mmax , but when it does
so, its gut will contain food, and as the food passes from the gut to the body
mass, the mass of the bird will increase and exceed mmax before decreasing
as the costs exceed the contribution of food from the emptying gut. The bird
may be regarded as being replete until the costs of metabolism reduce the
mass to mmax , and the conditions for resuming diving are m = mmax , and
dm/dt < 0.
This is represented as described above except that the rate constants for the
different memories are allowed to differ, and the handling time is remem-
bered in addition. Hence
dmn 1
= − r n mn
dt th
dmtw
= 1 − r w mtw
dt
The third state variable, mth , is the memory of the handling time involved
in consuming the average particle of food, and is represented as follows:
dmth
= 1 − r h mth
dt
and as with the other memories, when the animal is not handling the
food the memory decays exponentially. Using these state variables, together
with m, the body mass of the animal, it is possible to derive the variables
t¯w (t) = mtw (t)/mn (t), where t¯w (t) is the expected waiting time to find the
next particle, t¯h (t) = mth (t)/mn (t), where t¯h (t) is the expected handling time
of the next particle, and f¯p (t) = m/mn , where f¯p (t) is the expected net
mass of food in the the next particle. These derived variables can be used to
construct additional rules to determine both the selection of the diet while
Modelling foraging behaviour of individuals and groups 307
Definition Units
State variables
mtw The remembered waiting time s
mth The remembered handling time s
mn The remembered number of particles eaten Dimensionless
Derived variables
t¯w Expected waiting time, mtw /mn
t¯h Expected handling time, mth /mn
f¯p Expected net mass of food in a particle, m/mn
Parameters
tw Time to find next particle s
th Handling time of particle s
fp Mass of food in particle g (wet weight)
rw The specific rate of decay of the memory of s−1
waiting times
rh The specific rate of decay of the memory of s−1
handling times
rn The specific rate of decay of the memory of s−1
numbers
feeding and the total duration of the dives. It is assumed that the equations
representing the state variables g, m and O all apply, where relevant, as in
the continuous case.
Using the new-state and the derived-state variables, additional rules can
be developed to control the behaviour of the diving bird. A candidate rule
for feeding is:
Accept the newly encountered particle of food if
where f p /th is the average feeding rate, ignoring the time spent waiting to
find the particle. The right-hand side represents the remembered rate of
feeding of the animal while it is feeding. The term m̄th (t)/[m̄tw (t) + m̄th (t)]
is the proportion of time spent handling the particle, so the left-hand side
of the equation represents an estimate of the feeding rate (including both
waiting time and handling time). The right-hand side represents the cur-
rent metabolic cost in terms of the wet mass of food ingested. Table 20.5
summarizes the new-state variables and parameters.
308 J. G. Ollason et al.
The chosen time is the earliest of all of these. The model has been param-
eterized to represent the foraging and diving behaviour of the guillemot
(Ollason et al. 2003). This document contains the solutions of all equations,
the simulation programs, the documentation of the programs and sam-
ple runs of the simulation of the individual guillemot feeding on sandeels
at depth. It is important to note that although the seabirds use oxygen,
they are never abundant enough to cause local depletion. The spatial dis-
tribution of the foragers will be determined by the properties of the prey
only. If significant depletion of the prey occurs as a result of the feed-
ing by the foragers, this will cause the distribution of foragers to approx-
imate to the ideal free distribution as predicted by both the continuous-
feeding and the particulate-feeding models in their simple forms. However,
even without the depletion of prey, interference increasing as a function of
local density of predators can also permit the ideal free distribution to arise
(Ollason 1987).
CONCLUSION
ACKNOWLEDGEMENTS
REFERENCES
Charnov, E. L. (1976). Optimal foraging: the marginal value theorem. Theor. Popul.
Biol., 9, 129–36.
Cowie, R. J. (1977). Optimal foraging in great tits (Parus major). Nature, 268, 137–9.
Iwasa, Y., Masahiko, H. & Yamamura, N. (1981). Prey distribution as a factor
determining the choice of optimal foraging strategy. Am. Nat., 117, 710–23.
Kennedy, M. & Gray, R. D. (1993). Can ecological theory predict the distribution of
foraging animals? Oikos, 68, 158–66.
Liu, K. (2002). Modelling the physiology, behaviour, and ecology of dive foraging
seabirds. Unpublished Ph.D. thesis, University of Aberdeen, Scotland, UK.
Oaten, A. (1977). Optimal foraging in patches: a case for stochasticity. Theor. Popul.
Biol., 12, 263–85.
Ollason, J. G. (1980). Learning to forage: optimally? Theor. Popul. Biol., 18, 44–56.
(1987). Learning to forage in a regenerating patchy environment: can it fail to be
optimal? Theor. Popul. Biol., 31, 13–32.
Ollason, J. G. & Ren, N. (2002). Taking the rough with the smooth: foraging for
particulate food in continuous time. Theor. Popul. Biol., 62, 313–27.
(2004). A general dynamical theory of foraging in animals. Discrete Contin. Dyn.
System Ser. B, 4, 713–20.
Ollason, J. G. & Yearsley, J. M. (2001). The approximately ideal, more or less free
distribution. Theor. Popul. Biol., 59, 87–105.
Ollason, J. G., Ren, N., Scott, B. E. & Daunt, F. (2003). A Model of the Foraging
Behaviour of a Diving Predator Feeding Underwater at a Patch of Food (Revision
2004.3.0). IMPRESS Technical Report 2003-14 (Available from The
Netherlands Institute for Sea Research, Texel, the Netherlands.)
21
The Scenario Barents Sea study: a case of
minimal realistic modelling to compare
management strategies for marine ecosystems
T. SCHWEDER
of the process as observed by the player. For our purpose, the actions open
to the Agency are to set TACs for removals of harp seals and minke whales.
The Agency also sets TACs for the fisheries. TACs are assumed to be exactly
filled if abundance allows. Fishermen also behave according to economic
realities, and may discard or take more than the TAC. These economic real-
ities are disregarded here.
The strategy of Nature is determined by the internal dynamics of the
ecosystem, and its dynamic reaction to the removals caused by the fisheries.
The better the Agency knows Nature’s strategy, the better it can assess the
quality of a given management strategy for whaling and sealing. Our study
is an attempt to assemble the available knowledge and data pertinent to the
relevant dynamics of the upper trophic level of the ecosystem.
Despite more than 100 years of marine research in the area and despite
the fact that the system is less complex than many other ecosystems and
is comparably well known, our knowledge of the Barents Sea ecosystem is
rather limited. The quality of understanding varies considerably depending
on the topic. For features of the ecosystem about which little is known, the
model must be simple in order not to spread the available information too
thinly. Other better known features might be less relevant for the interaction
between the modelled species, and are therefore modelled in less detail than
is possible. Our aim is not a detailed description of the ecosystem represent-
ing all available knowledge, but rather a practical and reasonably realistic
model tailored to the purpose of the study. Borrowing a term from Punt
and Butterworth (1995), a model that balances realism and uncertainty,
and that is operational and practical to use, is called a ‘minimal realistic
model’.
We consider the Barents Sea, including the spawning grounds for north-
east Arctic cod in Lofoten, and also a residual area where Norwegian spring-
spawning herring and minke whales are found when they are not present in
the Barents Sea. Only young herring migrate to the Barents Sea. The minke
whales feed in the study area and breed elsewhere during winter. With
our limited and pragmatic purpose, we will include only harp seals, minke
whales, cod, capelin and herring in the model. Hamre (1994) describes the
main features of the Barents Sea–Norwegian Sea ecosystem.
A strategy should be evaluated in terms of its anticipated performance
in the long run. The system needs more than 100 years to become station-
ary, due to the slow dynamics of minke whales. We will therefore simulate
catches and stock status over at least a 100-year period, and study perfor-
mance by statistics summarizing the simulation results. It is beyond the
312 T. Schweder
scope of this chapter to give weights to the various objectives for manage-
ment, and to interpret results in financial terms.
The model has yearly stochastic variation, mainly in fish recruitment
and in the stock abundance estimates on the basis of which TACs are set.
This last uncertainty applies also to abundance estimates for minke whales
and harp seals, and translates into stochastic variation in TACs for these
species as well.
The Agency is faced with uncertainties with respect to the population
dynamics of the key stocks, and also to the interactions caused by preda-
tion between stocks. These uncertainties are handled by drawing parame-
ters from statistical distributions for each run of the simulation model, but
uncertainties with respect to functional forms are not addressed.
HISTORICAL BACKGROUND
80° N
Svalbard
6
75°
Bear Island
4
3 5
2 Barents Sea
70° Norwegian Sea
Lofoten
Vesteralen Norway Russia
Vestorden
1 White
Sea
65°
60° Karmay
Fig. 21.1 Spatial division of the Barents Sea, adapted for MULTISPEC (Bogstad
et al. 1997).
Overview
Cod, capelin, herring, harp seals and minke whales are distributed over
the seven areas used in previous studies (Bogstad et al. 1997) plus a resid-
ual area (Fig. 21.1), and over age and month. Fish are also distributed
314 T. Schweder
over lengths. The chosen subdivision of the study area and the time
step is regarded as the coarsest possible temporal and spatial stratifi-
cation respecting gradients in seasonal overlaps between predators and
prey in the Barents Sea system (S. Tjelmeland, personal communication,
2004).
Moving from one month to the next, surviving fish of the same length
are allocated to new length groups according to an individual growth sched-
ule which depends on season and species. Individual growth in cod depends
on the supply of capelin, but is independent of prey availability for capelin
and herring. A fixed length–weight relationship for each fish species is
used for calculating the biomass of the species at any time. This is needed
because fish TACs and predation are calculated in biomass terms.
The population-dynamics model for minke whales is taken from a report
of the IWC (1993), and is identical to that in Schweder et al. (1998). The
model for harp seal is taken from Skaug and Øien (2003). Recruitment
in cod, capelin and herring is modelled by Beverton–Holt functions aug-
mented with stochastic variability.
Minke whales and harp seals are top predators. Their population dynam-
ics is modelled as being unaffected by fish stock abundance. While this
is not quite realistic, it is nevertheless considered adequate for present
purposes because the modelled population dynamics broadly reflects prey
abundances similar to those expected over the simulation period. The
marine mammals prey on all three species of fish, and also on other organ-
isms, as do cod which are cannibals; herring prey on capelin larvae; capelin
are at the base of the model.
The interactions between populations are modelled solely as mortality
due to predation. These interactions are additive in the sense that they act
only to cause removals. Hjermann et al. (2004) found the mortality factors
to be additive and not compensatory for Barents Sea capelin. Fishing is also
assumed to cause removals but not to have indirect effects. In addition to
mortality due to modelled removals, there is a component of fixed residual
natural mortality.
Predation is modelled in two steps. The daily biomass consumed is cal-
culated for each group of predators from energetic considerations and from
other sources,. This consumption is then split between the groups of prey
according to prey choice probabilities. The prey choice probabilities are esti-
mated from stomach-content data and estimated prey availability by time
and area.
Key equations of the model are discussed below, together with some
comments on the estimation of their parameters. Additional information
The Scenario Barents Sea study 315
Juvenile capelin are recruited as 1 year olds on 1 October. Juvenile cod and
herring are recruited as 1 year olds on 1 January. Given a spawning stock
with biomass B, the number of recruits, R, one year later is calculated from
the Beverton–Holt function
R = L µB/(B + η)
where β is the mean number of live births per year for mature females at
carrying capacity, α is the resilience parameter and z is the degree of com-
pensation. The last two parameters determine the stock level, here 60%
of carrying capacity, corresponding to maximum sustainable yield. Recruit-
ment to the exploitable stock happens one year after birth. The recruitment
parameters are estimated using catch data, abundance data and a series of
relative abundance indices based on catch and effort data (Schweder et al.
1998).
Recruitment in harp seals is also modelled by the Pella Tomlinson for-
mula above, but with z = 1 degree of compensation, making the maximum
sustainable yield stock level around 50% (Skaug & Øien 2003).
Predation
Predation models are estimated anew for cod, minke whales and harp seals.
They consist of two components: total amount eaten per day, and diet choice
probabilities depending on prey abundance. The quantity consumed is esti-
mated from energetic considerations for harp seals and minke whales, and
also from experimental and observational studies for cod. The choice prob-
ability models have a common multiple logistic format, and are estimated
by comparing the type of prey found in sampled stomachs to estimated prey
abundance at the time and in the area of sampling.
Food takes time to digest, and sampled stomachs often have mixed
contents. Each stomach is therefore regarded as representing three feed-
ing events, and a rule codes them by food item. Minke whales, for exam-
ple, can choose between capelin, herring, cod and other food – indexed
from 1 to 4. If only one item is found in a stomach, all three feed-
ing events are coded by this item. For a covariate vector x represent-
ing area-specific abundance of cod, capelin and herring, and for param-
eter vectors β k related to food item k, the model consists of the linear
predictors
θk = βki xi
i
The Scenario Barents Sea study 317
4
pk = exp(θk ) exp(θj )
j=1
Zhu (2003) estimated the following diet choice model for minke whales
Management
DISCUSSION
Current management
Fish
TACs for cod are set to target a fishing mortality of F = 0.4 when spawning-
stock biomass is estimated above precautionary stock biomass, Bpa . The
stock is assessed by the VPA-method XSA (see ICES 2004). The capelin
and herring fisheries are essentially managed as modelled.
Minke whaling
Currently, Norway has a fleet of some 33 fishing vessels which harvest
minke whales in the summer months in Norwegian waters. Whaling
accounts for about 25% of the income for this fleet, and is economically sus-
tainable. Minke whaling resumed in 1993 after it was stopped in 1987. The
catch has increased from some 250 whales in 1993–5 to some 600 whales
in recent years. This is slightly below the TAC calculated from the man-
agement strategy. Vessel quotas are given to licensed whalers. The catch is
closely controlled with skilled observers on board.
The TAC for minke whales is calculated by the RMP of the IWC.
The input data to the RMP is the catch series and the series of absolute
abundance estimates obtained from double-platform, line-transect surveys
(Skaug et al. 2004).
The RMP can be tuned to target various population levels under cautious
standard assumptions. In the early years, a target level of 72% of carrying
capacity was used to set Norwegian minke whale TACs. This target level is
that which would be reached after 100 years of application of the RMP to
a stock, originally at its pristine level, with the lowest productivity rate con-
sidered plausible. In 2003 the target level was reduced to 66% of carrying
capacity, and for 2004 to 62%, resulting in higher TACs.
Harp sealing
Norwegian sealing is currently a small, heavily subsidized industry carried
out with old specialized sealing vessels. In the 1990s the yearly average
number of harp seals taken in the Barents Sea by Norway was 6200, and
has declined in recent years.
Norway manages harp sealing in the Barents Sea together with Russia,
and according to advice from ICES which calculates TACs to keep the stock
at its current level. The White Sea–Barents Sea stock is now estimated to
comprise nearly 2 million adults, and the TAC was 53 000 adult equivalents
The Scenario Barents Sea study 319
(one adult equals 2.5 pups) in 2001–3, well above the catch. Pups become
adult when they are 1 year old.
The economics of sealing has improved somewhat recently, but sealing
will probably need substantial subsidies to raise the catch to what is cur-
rently regarded as sound from an economic perspective for the fishery as a
whole (Stortingsmelding nr 27 2003–4).
Management issues
Modelling considerations
The various parts of the model were first estimated on relevant data, but
in isolation from the rest of the model and from other data. The model
as a whole is however more than the sum of its parts, and this piecewise
estimation has turned out to be unsatisfactory. To balance the model it has
been necessary to fit the model simultaneously to historical data. This is
work in progress.
320 T. Schweder
CONCLUSION
ACKNOWLEDGEMENTS
Thanks to Gro Hagen of the Norwegian Computing Center for keeping the
computer program under control, and to Bjarte Bogstad, Ulf Lindström and
Sigurd Tjelmeland of the Marine Institute, for providing data and biological
insight.
REFERENCES
Bogstad, B., Hiis Hauge, K. & Ulltang, Ø. (1997). MULTISPEC: a multispecies
model for fish and marine mammals in the Barents Sea. J. Northw. Atl. Fish.
Sci., 22, 317–41.
Butterworth, D. S. & Harwood, J. (1991). Report on the Benguela Ecology Programme
Workshop on Seal–Fishery Biological Interactions. Report of the Benguela Ecology
Programme, South Africa, 22.
322 T. Schweder
Skaug, H. J., Øien, N., Bøthun G. & Schweder, T. (2004). Abundance of minke
whales (Balaenoptera acutorostata) in the Northeast Atlantic: variability in time
and space. Can. J. Fish. Aquat. Sci., 61, 870–86.
Stortingsmelding nr 27 (2003–4). Norsk sjøpattedyrpolitikk. Oslo 19 March 2004
(summary in English at http://odin.dep.no/filarkiv/202967/marine mammal
summary final.pdf).
Zhu, M. (2003). Minke whale: a rational consumer in the sea. Unpublished M.Sc.
thesis in economics, University of Oslo, Norway.
22
Setting management goals using information
from predators
A. J. CONSTABLE
This chapter examines how goals and reference points might be set for
higher trophic levels – such as marine mammals, birds and fish. It briefly
explores the general characteristics of objectives for higher trophic levels
within the context of ecosystem-based management, noting that the empha-
sis for managing the effects of human activities on higher trophic levels is
biased towards fisheries-based approaches rather than approaches that take
into account the maintenance of ecosystem structure and function. Follow-
ing this, the precautionary approach developed in the Commission for the
Conservation of Antarctic Marine Living Resources (CCAMLR) for taking
account of higher trophic levels in setting catch limits for target prey species
is described. The last section considers indicators of the status of preda-
tors with respect to establishing target and limit/threshold reference points
that can be used directly for making decisions. These indicators include
univariate indices summarizing many multivariate parameters from preda-
tors, known as composite standardized indices, as well as an index of preda-
tor productivity directly related to lower trophic species affected by human
activities.
Ecosystem-based management encapsulates notions of conservation
and wise use of ecosystems (Mangel et al. 1996). Managers are now expected
(a) to maintain ecosystem properties and, in some cases, (b) to restore
ecosystems when they are judged to be impacted (caused to be altered),
directly or indirectly, by human activities. With appropriate scientific sup-
port, they need to define how ecosystems might be judged to be impacted
and to determine mechanisms for reducing or eliminating such impacts.
levels and too little consideration has been given to considering the wider
ecosystem consequences of changes to the composition of higher trophic
levels.
Higher trophic levels may be affected by human activities in many
ways, including direct exploitation, incidental mortality in fisheries, pollu-
tion, coastal works and mining, shipping and tourism. Direct interactions
can result in death (removal of individuals from populations), altered well-
being (change in individual growth and/or reproduction, change in health
or physiological condition, or displacement to other areas), or accumulated
Setting management goals using information from predators 327
Local extinction
Limit
Death
Altered spatial and
temporal relationships
Condition Displacement
Effect
Target - containment of effects
Habitat
Time
1.75
A B C
1.50
Relative biomass
1.25
1.00
0.75
0.50
0.25
0
0 10 20 30
Years
Fig. 22.2 Monte Carlo projection of a population model. See text for further
details.
(c) select the lower of γ 1 and γ 2 as the level for calculation of krill
yield.
Figure 22.2 (after Constable et al. 2000) illustrates the results of
Monte Carlo projections of a population model and how the derived
statistical distributions of krill abundance are used to determine the
lowest γ (criterion 3) from the two estimates of γ derived from
the first two criteria. Distribution A is the potential unexploited
biomasses derived from the model, which takes into account the
effects of uncertainties in krill demography. Distribution B is the
statistical distribution of lowest population biomasses over 20 years
of simulation under a specified catch limit of γ 1 B0 . Distribution C is
the statistical distribution of krill abundance at the end of 20 years
of exploitation under a specified catch limit of γ 2 B0 . Distributions
B and C incorporate uncertainties in the estimate of B0 . The dot–
dash line represents the median pre-exploitation biomass derived
from distribution A. The large-dashed line is the critical point of
the ‘recruitment criterion’ relative to the median pre-exploitation
biomass against which the tenth percentile of distribution B is com-
pared. The dotted line is the critical point of the ‘predator criterion’
relative to the median pre-exploitation biomass against which the
median of distribution C is compared.
Setting management goals using information from predators 333
Data on the status and function of marine ecosystems are mostly centred
on targeted fish species in fisheries. Consequently, this discussion focuses
on the indirect effects of fishing even though the term ‘fished species’ could
be used to refer also to species directly affected (killed) by pollution or other
human activities.
Critical density
CSI
50% fitness
Krill density
Fig. 22.3 A functional relationship between predator fitness (approximated by a
composite standardized index, CSI) and local krill density in the local foraging
area; this relationship is used to identify a critical density below which predator
fitness would be expected to be significantly below the norm. After Boyd
(2002).
nature of the relationship between a CSI for top predators and the availabil-
ity of krill provides the foundation for considering what might constitute
anomalies in a CSI without having to measure krill availability associated
with a value of the CSI (Constable & Murphy 2003). In this case, an anomaly
is a value of the CSI below which predator performance would be considered
to have been affected by a reduction in krill availability. The method to iden-
tify an anomaly, with high statistical power, will need to take account of vari-
ability in the CSI for a given level of krill density. Therefore, an anomaly for
the CSI could be classified as values less than, say, the lower 0.1 percentile
of CSI values in the baseline dataset recorded for krill densities greater than
the critical level of krill availability. This would probably be a CSI value
greater than the 0.1 percentile for all values of the CSI, which is the cur-
rent approach (Fig. 22.3).
One of the difficulties in the approach to monitor higher predator per-
formance is to weight appropriately (statistically) the different parameters
in the analysis. They should be highly correlated to changes in the abun-
dance of the species being directly affected, such as targeted fish species,
with the degree of correlation remaining independent of the characteristics
(e.g. abundance) of the target fish species or other characteristics of the food
338 A. J. Constable
where Np,a,y is the number at age in that year, B̂ p,a,y is the individual
mass of the predator at age in that year, G p,a,y is the age-specific
growth rate of individuals, R p,y is the number of offspring in that
year and t is the proportion of the year passed.
Constable (2001) proposed an operational objective for higher
trophic levels focusing on limit reference points for production aris-
ing from fished species rather than overall production of predators of
fished species. In a given year, y, the production arising from fished
species, P̃ p,y , can be specified as a fraction of the overall production
of the predator, such that
F
d p,y, j A p, j
j =1
P̃ p,y = Pp,y
D
d p,y,i A p,i
i =1
1.0
0.8
0.6
W
0.4
0.2
0.0
0 10 20 30 40 50 60
Year
Fig. 22.4 Hypothetical time series of W over 60 years. After Constable (2001,
2004).
y
P̃ p,y
y =1
Wref = ap
p Y
during the management period. The upper and lower dotted lines
show possible critical upper and lower range limits, WH and WL ,
for a case when the critical acceptable frequency outside the limit is
0.25. The dashed line refers to a possible interim upper range limit,
WiH , during the period when the system is adjusting to the fishing
activity. In this example, the trend for W to remain below the lower
range limit after 16 years of fishing would signal that a reduction in
fishing was required.
could be considered in this way (Fig. 22.5). The important part of this assess-
ment is to divide the system into a number of groups:
(1) Fished species. This managed system is where all species presumably
have target levels and/or threshold reference points applicable to them.
(2) Dependent predators of fished species. The effect of fishing on this
group can be considered as a whole – i.e. the effect of lost production
in the system – or could be subdivided to explore the effects on
individual species or groups of species.
(3) Prey of fished species and/or alternative prey of those predators. These
taxa might assume greater importance in the diet of predators and/or
might increase their productivity as a result of reductions in
abundance of their predators and competitors.
(4) Predators of the non-fished prey species in the third group. The
response of these predators would be difficult to foreshadow without
good knowledge of the function of the food web.
CONCLUDING REMARKS
Pacific walrus
Birds Toothed whales Sperm whales Beaked whales Bearded seals Baleen whales
Deepwater fish
Pelagics Cephalopods
Fig. 22.5 Food web of the Eastern Bering Sea showing primary interactions
with fished species along with other marine mammals and birds (based on data
and taxonomic groups from Trites et al. (1999)). Fished taxa are indicated by the
dark-grey boxes (O. dem., other demersal). Predators of fished taxa are in
light-grey boxes. Other taxa are in white boxes (Herb. zoopl., herbivorous
zooplankton; Juv., juvenile). Arrows indicate direction of prey to predators. Solid
lines indicate predation on fished species, dotted lines indicate predation on
non-fished species. The heavier weighted lines indicate where prey make up at
least 50% of the diet, lighter lines are where prey make up at least 20% but less
than 50% of the diet. Interactions where prey make up less than 20% of the diet
are not shown. After Constable (2001).
ACKNOWLEDGEMENTS
Many thanks to Bill de la Mare, Ian Boyd, Campbell Davies, John Croxall,
Inigo Everson, Keith Reid, Eugene Murphy, Phil Trathan, Marc Mangel,
Graeme Parkes, David Agnew, Denzil Miller, Roger Hewitt, Keith Sains-
bury and Tony Smith for many stimulating and challenging discussions
344 A. J. Constable
REFERENCES
Agnew, D. J. (1997). The CCAMLR Ecosystem Monitoring Program. Antarc. Sci.,
9, 235–42.
Andrewartha, H. G. & Birch, L. C. (1984). The Ecological Web: More on the
Distribution and Abundance of Animals. Chicago, IL: University of Chicago
Press.
Beddington, J. R. & Cooke, J. G. (1983). The Potential Yield of Fish Stocks. Fisheries
Technical Paper 242. Rome, Italy: FAO.
Beddington, J. R. & de la Mare, W. K. (1985). Marine mammal–fishery interactions:
modelling and the Southern Ocean. In Marine Mammals and Fisheries, eds. J. R.
Beddington, R. J. H. Beverton & D. M. Lavigne. London: George Allen &
Unwin, pp. 94 –105.
Beddington, J. R. & May, R. M. (1982). The harvesting of interacting species in a
natural ecosystem. Sci. Am., 247, 42–9.
Beverton, R. H. J. & Holt, S. J. (1957). On the Dynamics of Exploited Fish Populations.
London: HMSO.
Boyd, I. L. (2002). Integrated environment–prey interactions off South Georgia:
implications for management of fisheries. Aquat. Conserv., 12, 119–26.
Boyd, I. L. & Murray, A. W. A. (2001). Monitoring a marine ecosystem using
responses of upper trophic level predators. J. Anim. Ecol., 70, 747–60.
Butterworth, D. S., Gluckman, G. R., Thomson, R. B. & Chalis, S. (1994). Further
computations of the consequences of setting the annual krill catch limit to a
fixed fraction of the estimate of krill biomass from a survey. CCAMLR Sci., 1,
81–106.
Chesson, P. L. & Case, T. J. (1986). Overview. Nonequilibrium community
theories: chance, variability, history, and coexistence. In Community Ecology,
eds. J. Diamond & T. J. Case. New York: Harper & Row, pp. 229–39.
Constable, A. J. (2001). The ecosystem approach to managing fisheries: achieving
conservation objectives for predators of fished species. CCAMLR Sci., 8, 37–64.
(2002). CCAMLR ecosystem monitoring and management: future work.
CCAMLR Sci., 9, 233–53.
(2004). Managing fisheries effects on marine food webs in Antarctica: trade-offs
among harvest strategies, monitoring, and assessment in achieving
conservation objectives. Bull. Mar. Sci., 74, 583–605.
Constable, A. J. & Murphy, E. (2003). Annex 4, Attachment 3: using predator
response curves to decide on the status of krill availability: updating the
definition of anomalies in predator condition – preliminary analyses. In
SC-CAMLR, Report of the Twenty-second Meeting of the Scientific Committee for
the Conservation of Antarctic Marine Living Resources. Hobart, Australia:
CCAMLR, pp. 277–82.
Constable, A. J., de la Mare, W. K., Agnew, D. J., Everson, I. & Miller, D. (2000).
Managing fisheries to conserve the Antarctic marine ecosystem: practical
Setting management goals using information from predators 345
MODELLING APPROACHES
MARINE RESERVES
that marine reserves at sizes smaller than this will not provide protection for
mobile, wide-ranging species (Boersma & Parrish 1999). However, using
a modelling approach, Apostolaki et al. (2002) demonstrated benefits from
reserves primarily for over-exploited stocks of low-mobility species, but also
(to a lesser extent) for high-mobility species and under-exploited stocks. The
European Habitats Directive endorses Special Areas of Conservation (SACs)
which cover selected critical habitats across the whole range of a species.
Protection of critical or high-use habitats can confer protection, despite the
fact that reserves do not cover an entire species’ range.
Many higher predators are relatively site-faithful over time scales vary-
ing from daily (in the case of diurnal prey movements) to annual (in the
case of migratory species), to decadal (in the case of species following El
Niño events). Examination of animal movements often show certain high-
use areas (critical habitat or ‘hotspots’) at which protection would be valu-
able. Similarly protection should be provided for certain life-history stages
or other vulnerable stages, such as during migration (Gell & Roberts 2003).
Despite the fact that a predator might only use a protected area for a por-
tion of its life span, this would reduce the frequency with which it would
be exposed to some threats (such as fishery bycatch), and diminish the
overall cumulative impact of other threats (such as exposure to drilling
effluent).
Marine reserves and higher predators 353
MARINE PREDATORS
Breeding
Feeding
Migration
corridor
Feeding
Low High
Density
Fig. 23.1 Predator diversity hotspots. Based on Hooker and Gerber (2004).
SOCIO-ECONOMIC CONFLICTS
RESERVE NETWORKS
CONCLUSION
ACKNOWLEDGEMENTS
REFERENCES
Apostolaki, P., Milner-Gulland, E. J., McAlister, M. K. & Kirkwood, G. P. (2002).
Modelling the effects of establishing a marine reserve for mobile fish species.
Can. J. Fish. Aquat. Sci., 59, 405–15.
Barlow, K. E., Boyd, I. L., Croxall, J. P. et al. (2002). Are penguins and seals in
competition for Antarctic krill at South Georgia? Mar. Biol., 140, 205–13.
Baum, J. K., Myers, R. A., Kehler, D. G. et al. (2003). Collapse and conservation of
shark populations in the Northwest Atlantic. Science, 299, 389–92.
Boersma, P. D. & Parrish, J. K. (1999). Limiting abuse: marine protected areas, a
limited solution. Ecol. Econ., 31, 287–304.
Botsford, L. W., Castilla, J. C. & Peterson, C. H. (1997). The management of
fisheries and marine ecosystems. Science., 277, 509–15.
Bowen, W. D. (1997). Role of marine mammals in aquatic ecosystems. Mar. Ecol.
Prog. Ser., 158, 267–74.
Casey, J. M. & Myers, R. A. (1998). Near extinction of a large, widely distributed
fish. Science, 281, 690–2.
Cooke, S. J., Hinch, S. G., Wikelski, M. et al. (2004). Biotelemetry: a mechanistic
approach to ecology. Trends Ecol. Evol., 19, 334–43.
de Young, B., Heath, M., Werner, F. et al. (2004). Challenges of modeling ocean
basin ecosystems. Science, 304, 1463–6.
Estes, J. A., Tinker, M. T., Williams, T. M. & Doak, D. F. (1998). Killer whale
predation on sea otters linking oceanic and nearshore ecosystems. Science, 282,
473–6.
Gell, F. R. & Roberts, C. M. (2003). Benefits beyond boundaries: the fishery effects
of marine reserves. Trends Ecol. Evol., 18, 448–55.
Griffin, R. B. (1999). Sperm whale distributions and community ecology
associated with a warm-core ring off Georges Bank. Mar. Mamm. Sci., 15,
33–51.
Harwood, J. (2001). Marine mammals and their environment in the twenty-first
century. J. Mamm., 82, 630–40.
Harwood, J. & Stokes, K. (2003). Coping with uncertainty in ecological advice:
lessons from fisheries. Trends Ecol. Evol., 18, 617–22.
Hooker, S. K. & Gerber, L. R. (2004). Marine reserves as a tool for ecosystem-based
management: the potential importance of megafauna. Bioscience, 54, 27–39.
Hooker, S. K., Whitehead, H. & Gowans, S. (1999). Marine protected area design
and the spatial and temporal distribution of cetaceans in a submarine canyon.
Conserv. Biol., 13, 592–602.
(2002). Ecosystem consideration in conservation planning: energy demand of
foraging bottlenose whales (Hyperoodon ampullatus) in a marine protected area.
Biol. Conserv., 104, 51–8.
Marine reserves and higher predators 359
There has been much recent discussion over the need for an ‘ecosystem
approach’ to be taken to the management of human activities both on
land and at sea. There are a number of definitions of ‘ecosystem approach’
but a common feature is the need to set objectives for management. The
ecosystem approach to management encompasses effects on the ecosys-
tem as well as social and economic effects. Top predators are a key part
of the marine ecosystem and thus objectives ought to be set for them.
Objectives that might be set are broadly of two types: ‘what do we want
to achieve’ (targets); and ‘what do we not want to happen’ (limits). Objec-
tives may describe several aspects of the system for which they are being
set – commonly objectives are set for the state of an animal population
or for the impact on that population. If objectives are to be useful they
need to be achievable, inter-compatible and responsive to management
actions. A major initiative in European waters at present is the establish-
ment of Ecological Quality Objectives (EcoQOs) in the North Sea, but other
implicit objectives have also been set for marine top predators. In relation
to top predators, societal wishes are usually related to the state of popula-
tions which integrates across a set of measures including population size,
growth, range, health and feeding relationships. It is often not possible to
measure these aspects of state easily or sufficiently widely to understand
what is happening to a population as a whole. In these cases, it may be
easier to use a proxy measure relating to the human pressure on the top-
predator population. This is possible if there is sufficient understanding
of the link between the proxy measure and state of the population. Use of
such proxy measures may simplify monitoring systems and avoid the risk
of managers attempting to achieve many conflicting objectives.
As with nearly all parts of the planet, human activities have become an
important factor in ecological change in European seas. As a consequence,
there are concerns that such changes will either directly or indirectly harm
future generations. Several frameworks are being established, or have been
established, to address these concerns. Many of these are based on the prin-
ciples of sustainability – that actions today should not reduce the options of
future generations. Sustainability is often viewed as being built on three
pillars: economic, social and environmental (with some adding a fourth
administrative pillar). Sustainable development involves ensuring a bal-
ance between these pillars, but noting that the environmental pillar is
probably the one with least room for compromise. We cannot be sure
that many of the adverse environmental changes caused by humans are
reversible.
Humans cannot, of course, manage marine ecosystems; we can only
manage our activities in those systems. One commonly advocated way
of addressing the issue of sustainability in marine ecosystems has been
the ecosystem approach to management (sometimes described as the
ecosystem-based approach). For any management approach or style to suc-
ceed, one of the first prerequisites is a set of objectives.
In the case of the ecosystem approach to management, these objectives
need to relate to individual parts of the ecosystem, or to processes or fea-
tures of the system as a whole. Objectives might relate also to several fea-
tures of individual parts of the ecosystem. If objectives are set, then ways of
measuring whether or not they are being achieved need also to be devised.
One classification of these parts, processes or features of the environment
is the DPSIR framework. In this framework, social and economic develop-
ments or Driving forces exert Pressures on the environment resulting in
changes to its State. This leads to Impacts on environmental quality, which
may elicit a societal or policy Response. One objective of this chapter is to
illustrate the application of the DPSIR framework to objective setting using
top predators as an example. The measurement of whether objectives are
being met requires a series of metrics. Objectives can also be aimed either
to avoid certain conditions (limits) or to achieve certain conditions (targets).
It is possible to derive limit objectives scientifically given a set of rules to
work to (such as: we want to ensure that a species does not become extinct),
but such objectivity is more difficult for target objectives. Rules may derive
from some sort of societal choice expressed in law or agreement, or from
elsewhere.
Marine-management objectives for top predators 363
Table 24.1. Examples of possible factors for which objectives might be set
for populations of seabirds and marine mammals, related to the DPSIR
framework
Top predators are a key part of any ecosystem, and thus if we want to manage
human effects on the ecosystem, it is logical that some objectives should be
set for these top predators. Table 24.1 is an example of the application of the
DPSIR framework to two groups of top predators.
If objectives and their associated metrics are to be established, then it
is important that the metrics are reliable. In suggesting metrics for some
features of the North Sea ecosystem (see below), ICES (2001) applied a set of
criteria that ideally ought to be met (Table 24.2). It is difficult for all criteria
to be met by all metrics and there are patterns that match within the DPSIR
framework. Thus for example, criteria (b), (c) and (e) in Table 24.2 all relate
to human activities that in general fall into the category of pressure. It is
therefore much more likely that metrics associated with a pressure will meet
these three criteria than metrics associated with state.
Table 24.2. ICES criteria for good ecological quality (EcoQ) metrics
(ICES 2001)
(a) Relatively easy to understand by non-scientists and those who will decide on
their use.
(b) Sensitive to a manageable human activity.
(c) Relatively tightly linked in time to that activity.
(d) Easily and accurately measured, with a low error rate.
(e) Responsive primarily to a human activity, with low responsiveness to other
causes of change.
(f) Measurable over a large proportion of the area to which the EcoQ metric is to
apply.
(g) Based on an existing body or time series of data to allow a realistic setting of
objectives.
Seal population trends No decline in population size or pup production of State Target 4/7
≥10% over a period of up to 10 years.
Bycatch of harbour porpoises Annual bycatch levels should be reduced to levels Impact Limit 6/7
below 1.7% of the best population estimate.
Oil pollution (measured in seabirds) Proportion of oiled common guillemots among Pressure Target 7/7
those found dead or dying on beaches should be
10% or less of the total found dead or dying, in
all areas of the North Sea.
Presence and extent of threatened and Develop further State 2/7
declining species
Utilization of seal breeding sites Develop further State 4/7
Mercury concentrations (in seabird Develop further Pressure 5/7
eggs and feathers)
Organochlorine concentrations (in Develop further Pressure 7/7
seabird eggs)
Plastic particles (in stomachs of Develop further Pressure 4/7
seabirds)
Local sandeel availability to Develop further Pressure 4/7
black-legged kittiwakes
Seabird population trends as an index Develop further State 3/7
of seabird community health
366 M. L. Tasker
The EcoQOs are the only explicitly set international objectives for top preda-
tors in European seas at present. EcoQOs were also established for other
environmental features in the North Sea. However, there are some implicit
objectives for top predators within other legislation. For instance the EU’s
Habitats Directive (92/43/EEC), obliges EU Member States to:
r take all appropriate steps to avoid, in special areas of conservation, the
deterioration of natural habitats ‘and the habitats of species and the
disturbance of species for which the area was designated’.
r take measures designed to maintain or restore, at favourable
conservation status, natural habitats and species of wild fauna and
flora of Community interest.
Grey and harbour seals, harbour porpoise and bottlenose dolphin are
included in the species listed under this Directive as requiring these objec-
tives to be met. The term ‘favourable conservation status’ has caused con-
siderable debate about its meaning, but nevertheless it is evident that some
objectives are required.
Similarly in the EU’s Birds Directive (79/409/EEC), Articles require EU
Member States to:
r take the requisite measures to maintain the population of the
species . . . at a level which corresponds in particular to ecological,
scientific and cultural requirements, while taking account of economic
and recreational requirements, or to adapt the population of these
species to that level.
r ensure their survival and reproduction in their area of distribution.
DISCUSSION
It is plainly possible to set goals for top predators in order to help implement
an ecosystem approach to management in European waters, but their use-
fulness and feasibility are not straightforward. There are only three ‘opera-
tional’ objectives set so far as EcoQOs, and although some implementation
has occurred (most of which has involved adopting pre-existing monitor-
ing programmes), rather little in the way of new management measures
to help meet these objectives has been undertaken. An exception to this
has been the recent EC Regulation concerning cetacean bycatches, although
Marine-management objectives for top predators 367
demonstrating that their changes are correlated. Boyd and Murray (2001)
have undertaken this using information from three top predators that feed
on krill, monitored for 22 years at South Georgia. Separate analysis of vari-
ables showed that there was a significant decline in population sizes over
this period, but there was no trend in a combined index that represented for-
aging conditions during the local breeding season. Combined indices may
not always be helpful in providing management advice. There is also a risk
in combining that, unless the correlation is very close, important informa-
tion from the monitoring could be lost. De la Mare and Constable (2000)
examined this issue theoretically for Antarctic ecosystem monitoring and
found that high coefficients of correlation were also needed if there were
large numbers of missing values in the data.
There are some potential advantages in using metrics relating to pres-
sures. The information from a pressure metric is likely to be relatively well
linked to a manageable human activity. It is also often easier to measure
pressures than it is to measure state or impact. An individual pressure will
often affect the state of a number of ecosystem features. However, public
concern usually relates to the state of an ecosystem feature, so that if a pres-
sure metric is to be used for management purposes relating to that feature,
it is necessary to understand the links between the amount of pressure and
the state of the feature.
For instance, there is wide concern about the depleted state of many
fish stocks, the number of by-caught animals in fisheries and the damage
inflicted to seabeds by fishing. All of these concerns are caused by fishing
pressure. It might be easier to monitor fishing pressure rather than to mon-
itor the state of the various ecosystem features of concern. However, this
would need a good understanding of the link between these areas of con-
cern and fishing pressure, which in the case of this example still require a
substantial research effort.
It is also worth noting that trends are often easier to monitor than are
absolute values. It may be easier to set an objective that maintains a trend
and avoids reversal of the trend.
The setting of objectives and the use of the associated metrics, as noted
above, is a societal activity. It is possible to derive limit objectives scientifi-
cally given a set of rules (such as: we do not want population abundance to
be driven down by human activity, or for a species to become extinct), but
even these rules are ultimately a societal choice. At present, the mechanics
of objective setting and devising metrics for them has largely taken place
among a relatively small group of scientists and policy-making administra-
tors and it has been influenced by non-governmental organizations with
Marine-management objectives for top predators 369
ACKNOWLEDGEMENTS
The work described in this chapter has benefited from the input and discus-
sions over several years with many colleagues in JNCC, ICES and OSPAR. I
thank them all, but in particular Jake Rice and members of the ICES Work-
ing Group on ecosystem effects of fishing activity. It is a personal view-
point and should not be taken as representative of the views of any of them
or of any organization. Ian Boyd and an anonymous referee considerably
improved an earlier draft of the paper.
REFERENCES
Boyd, I. L. & Murray, A. W. A. (2001). Monitoring a marine ecosystem using
responses of upper trophic level predators. J. Anim. Ecol., 70, 747–60.
de la Mare, W. K. & Constable, A. J. (2000). Utilising data from ecosystem
monitoring for managing fisheries: development of statistical summaries of
indices arising from the CCAMLR Ecosystem Monitoring Program. CCAMLR
Sci., 7, 101–17.
ICES (International Council for the Exploration of the Sea) (2001). Report of the
ICES Advisory Committee on Ecosystems, 2001. ICES Cooperative Research
Report 249. Copenhagen, Denmark: ICES.
IMM (1997). Statement of Conclusions: Intermediate Ministerial Meeting on the
Integration of Fisheries and Environmental Issues, 13–14 March 1997, Bergen,
Norway. Oslo, Norway: Ministry of Environment.
Index
Page numbers in italic denote figures. Page numbers in bold denote tables.