ArticlePDF Available
Projected impacts of climate change on marine fish and fisheries
Anne B. Hollowed1*, Manuel Barange2, Richard J. Beamish3, Keith Brander4, Kevern Cochrane5,
Kenneth Drinkwater6, Michael G. G. Foreman7, Jonathan A. Hare8, Jason Holt9, Shin-ichi Ito10,
Suam Kim11, Jacquelynne R. King3, Harald Loeng6, Brian R. MacKenzie12, Franz J. Mueter13,
Thomas A. Okey14, Myron A. Peck15, Vladimir I. Radchenko16, Jake C. Rice17, Michael J. Schirripa18,
Akihiko Yatsu19, and Yasuhiro Yamanaka20
1
Alaska Fisheries Science Center, National Marine Fisheries Service, National Oceanic and Atmospheric Administration, 7600 Sand Point Way NE,
Seattle, WA 98115, USA
2
Plymouth Marine Laboratory, Prospect Place, Plymouth PL1 3DH, UK
3
Fisheries and Oceans Canada, Pacific Biological Station, 3190 Hammond Bay Rd., Nanaimo, BC, Canada V9T 6N7
4
Center for Macroecology, Evolution and Climate, DTU Aqua-National Institute of Aquatic Resources,Technical University of Denmark, Charlottenlund
Castle, Jaegersborg Alle
´1, 2920 Charlottenlund, Denmark
5
Department of Ichthyology and Fisheries Science, PO Box 94, Grahamstown 6150, South Africa
6
Institute of Marine Research, PO Box 1870, Nordnes, 5817 Bergen, Norway
7
Fisheries and Oceans Canada, Institute of Ocean Sciences, 9860 W. Saanich Rd, PO Box 6000, Sidney, BC, Canada V8L 4B2
8
NOAA Fisheries, Northeast Fisheries Science Center, Narragansett Laboratory, Narragansett, RI, USA
9
National Oceanography Centre, Joseph Proudman Building, 6 Brownlow Street, Liverpool L3 5DA, UK
10
Tohoku National Fisheries Research Institute, FRA, 3-27-5 Shinhama-cho, Shiogama, Miyagi 985-001, Japan
11
Department of Marine Biology, Pukyong National University, 599-1 Daeyeon-3dong, Nam-gu, Busan R 608-737, Korea
12
Center for Macroecology, Evolution and Climate and Center for Ocean Life, DTU Aqua-NationalInstitute of Aquatic Resources, Technical University of
Denmark, Kavalergu
ˆE
¨rden 6, DK 2920 Charlottenlund, Denmark
13
School of Fisheries and Ocean Sciences, Juneau Center, University of Alaska, Fairbanks, 17101 Pt. Lena Loop Rd, Juneau, AK 99801, USA
14
School of Environmental Studies, University of Victoria, PO Box 3060 STN CSC, Victoria BC V8W 3R4, Canada
15
Institute for Hydrobiology and Fisheries Science, Olbersweg 24, 22767 Hamburg, Germany
16
Pacific Research Institute of Fisheries and Oceanography (TINRO-Center), 4 Shevchenko Alley, Vladivostok, Primorsky Kray 690950, Russia
17
Science Sector, Department of Fisheries and Oceans, 200 Kent Street Station 12S015, Ottawa, ON, Canada K1A0E6
18
Southeast Fisheries Science Center, National Marine Fisheries Service, National Oceanic and Atmospheric Administration, 75 Virginia Beach Dr.,
Miami, FL 33149, USA
19
Seikai National Fisheries Research Institute, Fisheries Research Agency, 1551– 8 Taira-machi, Nagasaki 851– 2213, Japan
20
Graduate School of Environmental Science, Division of Environmental Resources, Hokkaido University, Hokkaido, Japan
*Corresponding author: tel: +1 206-526-4223; fax: +1 206-526-6723; e-mail: anne.hollowed@noaa.gov
Hollowed, A. B., Barange, M., Beamish, R., Brander, K., Cochrane, K., Drinkwater, K., Foreman, M., Hare, J., Holt, J., Ito, S-I., Kim, S., King, J., Loeng, H.,
MacKenzie, B., Mueter, F., Okey, T., Peck,M. A., Radchenko, V., Rice, J., Schirripa, M., Yatsu, A., and Yamanaka, Y. 2013. Projected impacts of climate
change on marine fish and fisheries. ICES Journal of Marine Science, 70: 10231037.
Received 7 December 2012; accepted 3 May 2013; advance access publication 6 July 2013.
This paper reviews current literature on the projected effects of climate change on marine fish and shellfish, their fisheries, and fishery-
dependent communities throughout the northern hemisphere. The review addresses the following issues: (i) expected impacts onecosys-
tem productivity and habitat quantity and quality; (ii) impacts of changes in production and habitat on marine fish and shellfish species
including effects on the community species composition, spatial distributions, interactions, and vital rates of fish and shellfish; (iii) impacts
on fisheries and their associated communities; (iv) implications for food security and associated changes; and (v) uncertainty andm odelling
skill assessment. Climate change will impact fish and shellfish, their fisheries, and fishery-dependent communities through a complex suite
of linked processes. Integrated interdisciplinary research teams are forming in many regions to project these complex responses. National
Published by Oxford University Press on behalf of the International Council for the Exploration of the Sea 2013. This work is written by US
Government employees and is in the public domain in the US.
ICES Journal of
Marine Science
ICES Journal of Marine Science (2013), 70(5), 1023– 1037. doi:10.1093/icesjms/fst081
at Bibliothekssystem Universitaet Hamburg on January 2, 2017http://icesjms.oxfordjournals.org/Downloaded from
and international marine research organizations serve a key role in the coordination and integration of research to accelerate the produc-
tion of projections of the effects of climate change on marine ecosystems and to move towards a future where relative impacts by region
could be compared on a hemispheric or global level. Eight research foci were identified that will improve the projections of climate impacts
on fish, fisheries, and fishery-dependent communities.
Keywords: climate change, fish, fisheries, fisheries-dependent communities, uncertainty, vulnerability assessment.
Introduction
The marine science community now regularly uses climate change
projections released by the Intergovernmental Panel on Climate
Change (IPCC; IPCC, 2007) to make qualitative and quantitative
projections of marine ecosystem responses to environmental
changes associated with the accumulation of greenhouse gases in
the atmosphere (e.g. climate change and ocean acidification).
These projections indicate that climate change will affect fish, fish-
eries, and fisheries-based economies around the globe as well as
broader components of marine ecosystems (ACIA, 2005;Allison
et al., 2009;Cochrane et al., 2009;Drinkwater et al., 2010;
Blanchard et al., 2012;Doney et al., 2012;Merino et al., 2012).
The potential implications of climate change for marine ecosystems,
and goods and services derived from marine ecosystems, have
prompted the formation of integrated interdisciplinary research
partnerships to quantify these impacts in many regions throughout
the world (Figure 1;Barange et al., 2011;Wiese et al., 2012). Several
international organizations [e.g. the International Council for
Exploration of the Sea (ICES), the North Pacific Marine Science
Organization (PICES), the Intergovernmental Oceanographic
Commission (IOC), the World Meteorological Organization
(WMO), and the Food and Agriculture Organization of the
United Nations (FAO)] and international research programmes
(e.g. Ecosystems Studies of Sub-Arctic Seas, ESSAS) have sponsored
symposia focused on climate change effects on marine ecosystems to
encourage international research partnerships and to widely dis-
seminate new research findings (Valde
´set al., 2009;Hollowed
et al., 2011;Drinkwater et al., 2012;Salinger et al., in press).
In this paper, we synthesize existinginformation to elucidate the
expected effects of climate change on fish and fisheries to guide
future research. Other international (e.g. the IPCC) and national
climate assessment teams have provided a comprehensive evalu-
ation of climate change impacts on marine and terrestrial ecosys-
tems on regional (e.g., Arctic Climate Impact Assessment; ACIA,
2005; Arctic Monitoring Assessment Program; AMAP, 2011; and
National Climate Assessment; Howard et al., 2013) and global
scales (IPCC, 2007). Our synthesis focuses on the implications
on a limited set of components of marine ecosystems and the
goods and services they provide. We consider the following
themes: (i) expected impacts on ecosystem productivity and
habitat quantity and quality; (ii) impacts of changes in production
and habitat on marine fish and shellfish species including effects on
the community species composition, spatial distributions, interac-
tions, and vital rates of fish and shellfish; (iii) impacts on fisheries
and their associated communities; (iv) implications for food secur-
ity and associated changes; and (v) uncertainty and modelling skill
assessment. Using this synthesis of information, key research activ-
ities are identified that may serve to guide future investigations.
Impacts on ecosystem productivity and habitat
In a world with high atmospheric CO
2
levels, global physical models
project increased sea temperatures in manyregions, changes in loca-
tions and magnitudes of wind patterns and ocean currents, loss of
sea ice in Polar Regions, and a rise in the sea level (IPCC, 2007).
The accumulation of CO
2
in the atmosphere and associated
climate changes is expected to cause ocean acidification and expan-
sion of oligotrophic gyres (Doney et al., 2012). These physical and
chemical changes are expected to result in shifts in the timing,
species composition, and magnitude of seasonal phytoplankton
production (Figure 2;Cochrane et al., 2009;Wang and Overland,
2009;Polovina et al., 2011;Doney et al., 2012). Changes in phyto-
plankton species composition may include shifts to smaller sizes
that could lengthen food chains and increase assimilation losses to
higher trophic levels (Mora
´net al., 2010;Bode et al., 2011).These
physical, and resulting biological, changes will occur at different
spatial and temporal scales throughout the world’s oceans
(Burrows et al., 2011;Gnanadesikan et al., 2011;King et al., 2011).
Changes in temperature, nutrient supply, mixing, light availability,
pH, oxygen, and salinity are expected to affect the ecological func-
tions and, consequently, the sustainable harvests available from
the ocean’s biological communities (Cochrane et al., 2009;
Brander, 2010;Denman et al., 2011;Doney et al., 2012). Exposure
of marine organisms to ocean acidification and oxygen depletion
will vary regionally, and other anthropogenic impacts (e. g., eu-
trophication) may also contribute. The vulnerability of species to
these changes varies considerably (Whitney et al., 2007;Feely
et al., 2008;Vaquer-Sunyer and Duarte, 2008;Levin et al., 2009;
Ries et al., 2009;Rabalais et al., 2010).
Regional differences in primary production are also antici-
pated. In mid-latitudes the mixed layer depth (MLD) is projected
to shoal, which could decrease nutrient supply and ultimately
primary production. For example, an intercomparison study of
11 models projected that the ocean’s MLD will change (decrease
or shoal) in most regions of the North Pacific during the 21st
century as the result of increased stratification resulting from
warming and/or freshening of the ocean surface and changes
in the winds (Jang et al., 2011). A study using four Earth
System Models (ESMs) found a similar pattern in the North
Atlantic (Steinacher et al., 2010). Capotondi et al. (2012) also
provide a global treatment of stratification changes. Primary pro-
duction in mid-latitudes is expected to be reduced by this MLD
shoaling through decreased nutrient supply (Hashioka and
Yamanaka, 2007;Barange and Perry, 2009). However, production
may increase in higher latitudes especially in seasonally ice-
covered areas through increased light levels and a longer period
of production and changes in the ice-edge bloom (Perrette
et al., 2011). Increased stratification caused by sea surface fresh-
ening and/or warming is also a main driver of ocean deoxygen-
ation through decreased ventilation (Whitney et al., 2007).
Rykaczewski and Dunne (2010) hypothesized that decreased ven-
tilation in upwelling zones may increase production due to
increased residence times (the period where producers are
retained in the high production zone) and nutrient remineraliza-
tion; however, we note that these benefits could be offset by
reduced nutrient supply. There remain important questions
regarding how physical and biological processes are incorporated
1024 A. B. Hollowed et al.
at Bibliothekssystem Universitaet Hamburg on January 2, 2017http://icesjms.oxfordjournals.org/Downloaded from
into projection models (e.g. temperature response; Taucher and
Oschlies, 2011) and how these models represent coastal and
shelf sea areas (e.g. Holt et al., 2012).
The responses of secondary production to climate change are
not clear, partially because the data available for zooplankton
are more limited and the mechanisms linking secondary
production to ocean conditions are complex. In the North
Atlantic, the total abundance of zooplankton changed with
sea surface temperature (SST) change (Richardson and
Schoeman, 2004). However, this overall pattern masks
important trends in the zooplankton community where the
abundance of both herbivorous and carnivorous copepods
increased with phytoplankton abundance but the abundance
of neither group was directly correlated with SST. Several
authors have recognized that the phenology of zooplankton
may also be affected by a changing climate in both the
Atlantic and Pacific (Chiba et al.,2004;Edwards and
Richardson, 2004;Mackas et al., 2007). Although climate
changeresultsinanearlieronsetofproductioncycles,the
actual timing and changes in the magnitude of production
Figure 1. Overview of species and geographic location of investigations presented at the 2010 ICES/PICES/FAO symposium in Sendai, Japan
(session A2) and the 2012 ICES/PICES/IOC symposium in Yeosu, Korea (session S4) (also see special volume Hollowed et al., 2011).
Projected impacts of climate change on marine fish and fisheries 1025
at Bibliothekssystem Universitaet Hamburg on January 2, 2017http://icesjms.oxfordjournals.org/Downloaded from
varied in direction and was influenced by different mechanisms
among regions (Richardson, 2008). Our limited understanding
of the trophodynamic linkages between phytoplankton and
zooplankton adds considerable uncertainty to projections of
the responses of these groups to global change (Ito et al., 2010).
Impacts on marine fish and shellfish
Climate-driven changes in the environment may affect the physi-
ology, phenology, and behaviour of marine fish and shellfish at
any life-history stage, and any of these effects may drive population-
level changes in distribution and abundance (Loeng and Drinkwater,
2007;Drinkwateret al.,2010;Jørgensenet al .,2012). Fish and shellfish
will be exposed to a complex mix of changing abiotic (e.g. tempera-
ture, salinity, MLD, oxygen, acidification)and biotic (shifting distri-
bution, species composition, and abundance of predators and prey)
conditions making it difficult to predict the responses.
Many climate-related changes have already been observed
(Table 1;Perry et al., 2005;Mueter and Litzow, 2008;Barange and
Perry, 2009;Nye et al., 2009). Kingsolver (2009) identified three
types of potential responses of species to climate change: distribu-
tion changes in space and time, productivity changes, and adapta-
tion. The extent of population-level changes may be mediated by
the capacity for individual species/populations to adapt to
changes in important abiotic and biotic factors through changes
in the phenology of important life-history events (e.g. migration,
spawning), or through changes in organismal physiology (e.g.
thermal reaction norms of key traits such as growth; Po
¨rtner,
2010) and/or through acclimation (Donelson et al., 2011).
Mismatches may occur when shifts in the environment lack consist-
ent patterns or out-pace the species ability to adapt or acclimate to
change (Burrows et al., 2011;Duarte et al., 2012).
Changes in life cycle dynamics will occur in concert with
climate-induced expansion, contraction, and/or shifts in the
quality and quantity of suitable habitat, and different life stages
may be affected differently by changes in habitat characteristics
(Petitgas et al., 2013). Moreover, in some regions, changes in tem-
perature will be accompanied by changes in other abiotic factors.
For example, expected regional changes in precipitation could
lead to decreases or increases in local salinities which will have
major impacts on distributions and productivities of fish species
in coastal and estuarine areas. Thus, perhaps future thermal condi-
tions may be suitable for new immigrant species, but shifts in sali-
nities could make these waters uninhabitable, illustrating the
challenges of projecting future trends in species richness of fish com-
munities.
Table 1summarizes recent literature on observed and expected
shifts in spatial distributions of marine fish and shellfish.
Although there are many accounts of temperate species moving to
higher latitudes, presumably in response to warming (Table 1; e.g.
Beare et al., 2004;Perry et al., 2005), there is less evidence of contrac-
tion of ranges of boreal species (Genner et al., 2004;Rijnsdorp et al.,
2010). The distributional changes may be the result of either active
migration of living marine resources to higher latitudes or from
differential productivity of local populations in lower and higher
latitudes (Petitgas et al., 2012), and usually the causal factors are
poorly documented. The sensitivity of fish and shellfish stocks to
climate change may differ depending on whether the stock is at
the leading, trailing or center of the species range (Beaugrand and
Kirby, 2010). In some cases, latitudinal shifts will exacerbate mis-
matches due to concurrent changes in the light cycle and the
duration of the growing season (Kristiansen et al., 2011;Shoji
et al., 2011).
Figure 2. Conceptual pathways of direct and indirect effects of climate change and other anthropogenic factors on marine ecosystems, with their
implications to adaptation and management. Solid arrows, consequences of climate change; dotted arrows, feedback routes.
1026 A. B. Hollowed et al.
at Bibliothekssystem Universitaet Hamburg on January 2, 2017http://icesjms.oxfordjournals.org/Downloaded from
The aforementioned impact of climate change on MLD and
ocean chemistry has been shown to exacerbate vertical habitat com-
pression for some highly migratory species of billfish and tunas in
the tropical Northeast Atlantic Ocean. Initial work demonstrated
how the near-surface density of many high-oxygen demand
species of pelagic fish was much higher in the eastern than in the
western tropical Atlantic (Prince et al., 2010). Eastern boundary
current conditions off the west coast of Africa create an oxygen
minimum zone that is much closer to the surface than in the
western tropical Atlantic. The habitat compression has led to
higher vulnerabilities to surface fishing gear and artificially high
indications of abundance. Stramma et al. (2011) reported that a de-
crease in the upper ocean layer dissolved oxygen occurred in the
tropical Northeast Atlantic. This change equated to an annual
habitat loss of 15% over the period 1960 2010. Climate change
is expected to further expand the Atlantic oxygen minimum zone
due to increased ocean temperatures and decreased oxygen levels,
potentially threatening the sustainability of the pelagic fisheries
and their associated ecosystems.
Climate change may also influence recruitment success, which
will impact population productivity (e.g. Hare et al., 2010;
Mueter et al., 2011). The resilience to shifts in production
mayvarybyregion.Inmanyregions,fishandshellshhave
evolved within systems impacted by intermittent (12 years) or
longer term events that occur on decadal or multidecadal time-
scales (Baumgartner et al., 1992;Hare and Mantua, 2000;
Table 1. Recent studies of climate impacts on spatial distribution of marine fish and shellfish.
Reference
Publication
year Region LME Type # Species
Cheung et al. 2009 Global NA Retrospective and
Projection
Hollowed et al. In press b Arctic/Subarctic Barents Sea, Bering Sea, Arctic Vulnerability 17
Huse and
Ellingsen
2008 Arctic/Subarctic Barents Sea Retrospective and
Projection
1
Ciannelli and
Bailey
2005 Subarctic E. Bering Sea Retrospective 1
Mueter and
Litzow
2008 Subarctic E. Bering Sea Retrospective 46
Spencer 2008 Subarctic E. Bering Sea Retrospective 5
Sundby and
Nakken
2008 Subarctic Norwegian Sea Retrospective 1
Drinkwater 2005 Subarctic North Atlantic Projection 1
Drinkwater 2006 Subarctic Northern North Atlantic Retrospective 24
Dulvy et al. 2008 Subarctic North Sea Retrospective 29
Engelhard et al. 2011 Subarctic North Sea 1913–2007 2
Petitgas et al. 2012 Subarctic North Sea Retrospective 1
Perry et al. 2005 Subarctic North Sea 1977–2001 36
Welch et al. 2001 Subarctic North Pacific Ocean Retrospective and
Projection
1
Tseng et al. 2011 Subarctic Oyashio Current Retrospective and
Projection
1
Fogarty et al. 2008 Temperate NE US Continental Shelf Retrospective and
Projection
1
Hare et al. 2012a Temperate NE US Continental Shelf Projection 1
Nye et al. 2009 Temperate NE US Continental Shelf Retrospective 36
Hare et al. 2010 Temperate NE US Continental Shelf Retrospective and
projection
1
Last et al. 2011 Temperate Australian Shelf Retrospective 45
Ito et al. 2010 Subarctic /
Subtropical
Kuroshio/Oyashio current, Kuroshio
Extension
Projection 1
Okunishi et al. 2012 Subarctic /
Subtropical
Kuroshio/Oyashio current, Kuroshio
Extension
Projection 1
Yatsu et al. 2013 Subtropical /
Subtropical
Kuroshio/Oyashio current, Kuroshio
Extension
Vulnerability 4
Hare et al. 2012b Subtropical SE US Continental Shelf Projection 1
Agostini et al. 2008 Subtropical California Current Retrospective 1
King et al. 2011 Subtropical California Current Vulnerability 8
Hsieh et al. 2009 Subtropical California Current Retrospective 34
Stewart et al. 2012 Subtropical California Current Retrospective 1
Muhling et al. 2011 Tropical Gulf of Mexico Retrospective and
Projection
1
Su et al. 2011 Tropical Pacific Ocean Retrospective and
Projection
1
Lehodey et al. 2012 Tropical Pacific Ocean Retrospective and
Projection
1
Projected impacts of climate change on marine fish and fisheries 1027
at Bibliothekssystem Universitaet Hamburg on January 2, 2017http://icesjms.oxfordjournals.org/Downloaded from
Greene and Pershing, 2007;Di Lorenzo et al., 2008;Hatun et al.,
2009;Overland et al., 2010;Alheit et al., 2012). These events will
probably continue to occur in the future. It is unclear whether
species and communities that have experienced such variability
in the past will be better adapted to future climate change. In
some well-documented cases, climate variability is thought to
provide opportunities for dominance switching and ecosystem
reorganization (Skud, 1982;Southward et al.,1988;Anderson
and Piatt, 1999;Rice, 2001;Stenseth et al., 2002;Chavez et al.,
2003). Climate change may interrupt or accelerate these cycles
of dominance switching with unknown implications for both
dominant and subordinate species within each phase of a cycle.
The responses of individual marine species to climatechange will
vary by species and region resulting in a broad spectrum of potential
shifts in geographic ranges, vertical distributions, phenologies, re-
cruitment, growth, and survival. Thus, alterations in both the struc-
ture (i.e. assembly and connectivity) and function (i.e. productivity)
of biological communities are expected (Figure 2). Community
responses are the most uncertain types of ecosystem responses to
climate change because they involve more players (all the species in
the community and the habitats that are used), their interactions,
and direct as well as indirect effects of climate drivers (Stock et al.,
2011), as well as the spatial and temporal complexity of responses
(Burrows et al., 2011;Gnanadesikan et al., 2011). However, there is
some evidence that community assemblages tend to move in
concert based on retrospective studies of species spatial patterns and
species richness (Hofstede et al., 2010;Lucey and Nye, 2010).
Impacts on fishers, fisheries, and fishery-dependent
communities
Fisheries and fishery-dependent communities have been subjected
to fluctuations in fish stocks, extreme weather events, and natural
changes in climate and sea-level throughout history. Coastal liveli-
hoods have depended on the capacity to cope with such changes
through the alteration of fishing practices or switching to alternative
livelihoods (Allison et al., 2009;Perry et al., 2011). The capacity for
human communities to respond to changes in the species compos-
ition, abundance, and availability of marine resources vary region-
ally (Daw et al., 2009). Climate change effects on fish and fisheries
will occur within the context of existing and future human activities
and pressures, as well as the combined effects of multiple stressors
and natural agents of change acting directly and through feedback
pathways (Figure 2;Ruckelshaus et al., 2013). In coastal ecosystems,
pollution, eutrophication, species invasions, shoreline develop-
ment, and fishing generally play more important roles as drivers
of change than on the high seas.
It will be difficult to tease out the additional effect of climate
change from other anthropogenic activities (such as fishing;
Rogers et al., 2011). In some cases, where time-series are long
enough or can be re-constructed, the relative importance of differ-
ent forcings can be quantified (e.g. Eero et al., 2011). Hare et al.
(2010) examined the combined effects of fishing and climate in a
modelling context and found that fishing likely remains the domin-
ant pressure, especially at the historically high fishing levels. Other
researchers found that it was difficult to separate the influence of an-
thropogenic climate change from decadal environmental variability
and fishing even with a century of data (Engelhard et al., 2011;
Hofstede and Rijnsdorp, 2011), whereas others note that fisheries
can amplify or moderate climate signals (Ottersen et al., 2006).
Some promising alternative approaches to address these issues
include: comparative studies, experiments, and opportunistic
studies of major natural or anthropogenic events (Megrey et al.,
2009;Murawski et al., 2010). Ainsworth et al. (2011) used five
Ecopath with Ecosim models to simulate changes in primary pro-
duction, species range shifts, zooplankton community size structure
in response to ocean acidification, and/or ocean deoxygenation.
Fishing pressure was also included as an additional perturbation
to the modelled foodweb. Their study revealed that responses to
the cumulative effects of climate change and fishing may result in
different patterns than would have been predicted based on individ-
ual climate effects, indicating possible interactions.
The degree to which fisheries are managed sustainably varies glo-
bally (Worm and Branch, 2012). In many regions, efforts are under-
way to prevent overfishing, rebuild overfished stocks, and
implement an ecosystem approach to management (Murawski,
2007). In the future, the detrimental effects of climate change on
fish stocks may, to some extent, be buffered in stocks that have a
large and productive spawning-stock biomass, a less truncated age
structure, and sustainable exploitation rates (Costello et al., 2012).
For example, cod have remained abundant with wide size/age struc-
ture in some areas (i.e. Øresund) where exploitation has been low,
although temperatures have increased and while abundance has
declined and age structure has narrowed in neighbouring areas
(North Sea, Baltic Sea; Lindegren et al., 2010).
Natural scientists and economists are partnering to develop the
projections of how fishers may respond to changes in fish distribu-
tion and abundance (Haynie and Pfeiffer, 2012). It is unclear how
complex management systems involving measures such as catch
shares, bycatch limits, mixed species catch or effort limits, and
spatial or temporal closures will perform as the species composition,
distribution, and abundance of fish species change (Criddle, 2012).
An equally challenging issue is predicting how different nations will
utilize the broad range of ecosystem services that marine ecosystems
provide (Halpern et al.,2012). Multispecies management strategy
evaluations can be used to evaluate the expected performance of
management frameworks with respect to balancing these complex
issues (Plaga
´nyi et al., 2011). However, selecting the functional
form of responses necessary to predict how fishers will respond to
changes in marine resources will continue to be challenging.
The fish stocks, fisheries, and marine ecosystems that coastal
communities depend on can be described as components of
coupled marine social-ecological systems (Perry et al., 2011). This
is a particularly useful representation when considering the policy
goals of preserving the health of the marine ecosystem while main-
taining the supply of desirable goods and services that support
human livelihoods. The representation requires specifying the
scale of the system, its properties (e.g. resilience, biodiversity, prod-
uctivity, social capital), how it is, or can be, governed, and what
structures and information are required for such governance.
Management and governance approaches may need to be adapted
to the available scientific and management capacity (including fi-
nancial and social resources). While strengthening capacity may
put extra demands on management agencies and stakeholders, it
also brings with it greater sustainable benefits through reduced un-
certainty (Cochrane et al., 2009, 2011). Anthropogenic climate
change is an increasingly influential driver of change in such
social-ecological systems, added to an already complex set of
natural and anthropogenic drivers. The impacts of climate drivers
are manifested on time-scales that are generally longer than most
other anthropogenic drivers to which these social-ecological
systems routinely respond.
1028 A. B. Hollowed et al.
at Bibliothekssystem Universitaet Hamburg on January 2, 2017http://icesjms.oxfordjournals.org/Downloaded from
There is growing recognition of the need for much stronger inte-
gration of social and ecological sciences in developing adaptation
options for industries and coastal communities (Allison et al.,
2009;Daw et al., 2009;Miller et al., 2010;Gutierrez et al., 2011).
In this context, there may be much to learn from the dynamics of
small-scale fisheries in coastal communities. Institutions such as
the FAO and Worldfish are active in working on climate change
adaptation in such systems. Adaptation and mitigation depend on
actions and behavioural choices by the communities who are
exploiting the marine resources (whether for fisheries, tourism, or
other goods and services), as well as a supportive wider governance
environment to address threats and constraints to adaptation and
mitigation that are outside the control of local communities.
Resource users and communities, within the context of an inte-
grated ecosystem approach, must have the capacity and the will to
adapt and mitigate. Viable adaptation and mitigation actions
require the identification of vulnerabilities at levels from the house-
hold to macroeconomic ability to diversify livelihoods for income
and the availability of environmentally sustainable livelihoods and
development options. For example, “co-benefits” of both adapta-
tion and mitigation can arise from biodiversity conservation, and
protection and restoration of mangroves, and other coastal vegeta-
tion (Ruckelshaus et al., 2013). Coastal resources governance can be
encouraged to develop community-based disaster risk management
and to integrate climate change issues into the local and national
socio-economic development planning. These actions may help to
prepare communities for climate change impacts on livelihoods
that depend on marine resources.
Implications for future security of the food supply
The expansion of the world’s human population and current levels
of hunger in many parts of the world have raised concerns over the
security of the food supply in the future (OECD, 2008;Godfray et al.,
2010,2011). Fish currently provide essential nutrition to 4 billion
people and at least 50% of the animal protein consumed by 400
million people (Laurenti, 2007;FAO, 2012), currently contributing
17 kg of fish per capita and year. Most of the expected increase in
the human population to 2050 occurs in regions where fish provide
most of the non-grain dietary protein (UN-DESA, 2009;UN-WHO,
2002). The extent to which marine fisheries will be able to provide
fish for the world’s population in the future will depend on climate-
driven changes to the productivity of the world’soceans and the per-
formance of fisheries management systems (Bell et al., 2009;Worm
et al., 2009;Costello et al., 2012). Several scientists have used outputs
from IPCC global climate models to explore quantitatively or quali-
tatively the potential consequences of climate change on fish and
fisheries production and the implications in terms of food security
targets (e.g. Merino et al., 2012). These studies concluded that even
with improved management, there is only a modest scope for
increases in sustainable global yields for capture fisheries (Rice
and Garcia, 2011;Brander, 2012). However, innovation in both
large-scale and small-scale aquaculture may support a continued in-
crease in production from marine and freshwater systems (FAO,
2008a,b;OECD, 2008;Garcia and Rosenberg, 2009;Rice and
Garcia, 2011;Merino et al., 2012). At present, global aquaculture
production is very unevenly distributed with Asia accounting for
89% of world production (FAO, 2012). In addition, the effects of
climate change on prospects for fisheries and aquaculture show
strong regional differences (Merino et al., 2012). Substantial polit-
ical and financial investment in aquaculture will be required in suit-
able climatic and environmental regions if it is to provide greater
contributions to food security and meet the growing demand for
fish and seafood products. Growing international trade in fish pro-
ducts and fishing fleet capacities is accentuating regional differences
in potential fish consumption (OECD-FAO, 2009;Kim, 2010).
Hence, in addition to direct impacts of climate change on fish popu-
lations and communities, and thus food production, there can be in-
direct impacts through changes to the availability of alternative
sources of protein, to the conditions suitable for intensive culture
of fish and shellfish, and even to the complex interactions of
climate on the global trade in food.
Uncertainty and skill assessment
Almost all attempts to forecast the impacts of climate change on fish
and fisheries involve models of one form or another, and all these
models will include uncertainties in both model structure and par-
ameter values. A range of model types is used in fisheries research,
from simple empirical relationships through population dynamics
models to detailed system models (Hollowed et al., in press a).
Consideration of the diverse and complex interactions that occur
between the underlying drivers of climate change and their ultimate
impacts on fish and fisheries tends to require the use of relatively
complex models in an effort to achieve scientific realism. However
there are trade-offs since increasing model complexity to achieve
greater realism can reveal additional uncertainty associated with in-
complete knowledge of both the functional form and parameteriza-
tion of the model (FAO, 2008a,b;FAO, 2009).
There are many such uncertainties in assessing impactsof climate
change on marine ecosystems. For example, physical–biological
pathways are elucidated for only a few species or functional
groups. Our empirical knowledge may also not apply beyond previ-
ously observed ranges of environmental factors,or outside of histor-
ical rates and amplitudes of environmental change. Adaptation of a
species to new environmental conditions is one of the most difficult
issues to evaluate, especially when attempting to project connectiv-
ity among ecosystem components (Planque et al., 2011).
Furthermore, projecting climate change effects on fish and fisheries
is challenging due to the cumulative effects of climate change, other
anthropogenic activities, and feedback mechanisms (Fulton, 2011).
When physical–biological pathways are known, analysts must
consider what long-range forecast and a modelling method
should be used to project future states of nature. Long-term quan-
titative forecasts of climate change effects are generally based on
outputs from one or more global circulation models (GCMs) pro-
viding boundary conditions for species or ecosystem predictive
models. Inferences about biological responses to climate change
based on GCM outputs commonly deal with uncertainty in the
emission scenario forcing the GCMs (Hawkins and Sutton, 2009),
structural uncertainty in the GCMs, internal variability, and the
generally coarse resolution of the GCM, as well as uncertainty in
modelling the biological responses. The relative importance of dif-
ferent sources of uncertainty associated with GCM predictions
depend on the temporal and spatial scales of interest. Although
these have not been quantified in coupled atmosphere-ocean
GCMs, climate model predictions on both global and regional
spatial scales have been shown to be dominated by internal variabil-
ity in the climate over short time-scales (5–15 years), by model un-
certainty on intermediate scales (1540 years), and by scenario
uncertainty on longer time-scales (Hawkins and Sutton, 2009).
Although sensitive to emissions scenarios, there is broad agreement
among climate models for some parameters such as temperature,
even at short time-scales and on regional spatial scales (Deser
Projected impacts of climate change on marine fish and fisheries 1029
at Bibliothekssystem Universitaet Hamburg on January 2, 2017http://icesjms.oxfordjournals.org/Downloaded from
et al., 2012). Similarly, GCMs provide credible projections for re-
gional ocean temperatures (Wang et al., 2012). In contrast,
derived quantities computed from the GCM output (e.g. MLD)
can vary widely among models if they are based on parameters
that are poorly estimated by GCMs (Jang et al., 2011). Moreover,
there is generally a mismatch in spatial scales between the output
of the GCMs, which tend to have skill at an ocean-basin scale, and
the need for resolution of finer scale ocean processes on the
coastal shelves needed to project impacts on fish and fisheries
(Stock et al., 2011;MacKenzie et al., 2012;Meier et al., 2012).
Although there is a clear need to capture regional-scale processes,
there is no guarantee that high-resolution regional models will
provide improved predictions of regional climate changes com-
pared with GCMs (Racherla et al., 2012). Therefore, it is important
that scientists investigating fish responses to climate change correct-
ly understand the robustness and uncertainty of GCM-derived vari-
ables when they use these variables to predict biological responses.
In addition to uncertainty in GCM outputs, many sources of un-
certainty exist in models of biological responses (Planque et al.,
2011) and these should be accounted for when making projections
(Hare et al., 2012b). Various approaches have been used to quantify
the uncertainty associated with the projections of the potential
impacts of changing ocean conditions on marine fish and shellfish
(e.g. Loukos et al., 2003;Cheung et al., 2009,2010;Lindegren
et al., 2010;Fulton, 2011;Blanchard et al., 2012). These include bio-
climate envelope models to determine expected shifts in species
distributions as a result of changes in the availability of preferred
temperatures (Cheung et al., 2009,2011;Jones et al., 2012), fish
population models and end-to-end ecosystem models coupled to
regionally downscaled climate-physical oceanographic models
(e.g., MacKenzie et al., 2012;Meier et al., 2012). Methods used to
address uncertainty include, but are not limited to, the following:
(i) Hierarchical models: these models, using a fully Bayesian or a
empirical Bayes approach, provide a powerful tool for quanti-
fying uncertainty in the estimated responses of fish popula-
tions to climate variability across multiple stocks, regions, or
other “replicate” units (e.g. Mueter et al., 2002;Helser et al.,
2012). Because of the computational demands, such models
are only beginning to be applied to coupled biophysical
models (e.g. Fiechter et al., 2009).
(ii) Multiclimate model scenarios: the most basic approach to
characterizing, if not quantifying, uncertainty about potential
future responses to climate change consists of presenting
results and implications from the analysis of different
models and comparing and contrasting the resulting patterns
across models (A’mar et al., 2009;Hare et al., 2010).
(iii) Ensemble modelling: this approach is commonly used to char-
acterize uncertainty in climate projections across multiple
models (Hollowed et al., 2009;Wang and Overland, 2009)
and has recently also been used in coupled models to
examine uncertainty in both climate trajectories and in the
biological responses (Ito et al., 2013;Mueter et al., 2011).
This approach is used when analysts find that some of the dif-
ferent oceanographic models may perform better than others
to reproduce the physical or biological oceanographic vari-
ables (e. g., temperature, plankton production) that influence
the fish population dynamics (MacKenzie et al., 2012).
Biological models in these ensemble approaches may be
driven by dynamically (e.g. Ito et al., 2013) or statistically
downscaled climate scenarios (Meier et al., 2012;MacKenzie
et al., 2012). An outstanding issue in ensemble modelling is
the criteria to decide which models should be included in
the ensemble and/or how they should be weighted.
Overland and Wang (2007) reduced a set of 22 GCMs to 10
based on how well they simulated the variability of 20th
century North Pacific SSTs. Depending on which particular
variables are of interest, other selection criteria could of
course be devised. Additionally, good model performance
evaluated based on historical or present climate does not ne-
cessarily imply certainty in predictions of future climate.
However, Reichler and Kim (2008) note that the retrospective
assessment of the skill of simulations relative to observations is
an important way to evaluate confidence in projections.
(iv) Monte-Carlo approaches: whether or not the impacts of mul-
tiple models are investigated, a simulation (Monte Carlo) ap-
proach can generally be used to quantify uncertainty when
making projections. Simulations can account for known un-
certainty in future climate (random drawsof climate trajector-
ies based on different emission scenarios), in population
dynamics (random draws of important population para-
meters from multiple univariate or, better, a single multivari-
ate distribution), and in environmentbiology relationships
(random draws of parameter values for estimated or
assumed functional relationships from a suitable probability
distribution or from historical values; Mueter et al., 2011;
Planque et al., 2011). A simulation approach is also utilized
in the context of Management Strategy Evaluations, which
allows the robustness of management strategies to be tested
in the face of system uncertainty, but at the expense of consid-
erable time and processing power (Ianelli et al., 2011). The re-
liability of such simulations depends on specifying both the
functional forms and the sampling distribution of the para-
meters correctly, which in some data-limited situations can
be more difficult than merely estimating the central moment
of the distribution correctly and using other means to incorp-
orate uncertainties in the final result (Rochet and Rice, 2009).
(v) Parameter sensitivity: estimating the sensitivity of model
outputs to changes in values of parameters is the primary
means for identifying particularly influential parameters
(Maunder et al., 2006;Haltuch et al., 2009;Peck and
Hufnagl, 2012). If models are particularly sensitive to a given
parameter, uncertainty about the true parameter value is an
important source of overall uncertainty. Sensitivity analyses
are typically used to prioritize field and laboratory studies
(e.g. Peck and Hufnagl, 2012), but they can also be used to
quantify uncertainty in projections by repeatedly running
models across different values of the important parameters
to bracket possible responses. However, this requires some
knowledge of the likely distribution of parameter values and
it can be challenging with complex models that have multiple,
important parameters that require a large number of model
runs. Gibson and Spitz (2011) and Fiechter (2012) provide
examples of exploring the effects of parameter uncertainty in
a nutrientphytoplankton– zooplankton detritus (NPZD)
model on estimates of phytoplankton biomass in the eastern
Bering Sea and Gulf of Alaska, respectively.
Each modelling approach has strengths and weaknesses and, as for
the physical realm, multimodel projections may provide additional
1030 A. B. Hollowed et al.
at Bibliothekssystem Universitaet Hamburg on January 2, 2017http://icesjms.oxfordjournals.org/Downloaded from
insights into the range of impacts to fish and fisheries that could
occur under future climate change (Plaga
´nyi et al., 2011;Stock
et al., 2011;Link et al., 2012;Hollowed et al., in press a). A parallel
alternative is the development of models that combine principles
and algorithms from several modelling frameworks, such as the in-
clusion of size-based ecological constraints embedded in bioclimate
envelope models (Fernandes et al., in press). This approach helps
assess the relative strengths of each model and makes predictions
more realistic and robust to assumptions.
Uncertainty in fish population simulations may be more fully
characterized by using a suite of models representing different com-
ponents of the climateocean–ecosystem complex. Compounding
the uncertainty of projected fish responses is the availability of mul-
tiple representations of the fish population dynamics (e.g. single-
species model, predator– prey interactions model, foodweb
models, etc.) which can be coupled to the outputs from the available
physical oceanographic models. Consequently, the availability of
different climate-physical oceanographic and ecological models
for a given system presents an opportunity to investigate a wide
range of climate-oceanographic and biological model assumptions
and parameterisations (e.g. via sensitivity analysis), particularly by
combining the different climate-oceanographic and population
models (MacKenzie et al., 2012;Meier et al., 2012). This approach
can identify both the range and similarity of possible biological
responses to different model frameworks and identify critical gaps
in knowledge and new hypotheses for investigation.
Recommendations
Our synthesis elucidated several research foci that will be needed to
improve the projections of climate impacts on fish and fisheries.
The scale and ecological importance of climate change research
for the marine community will require coordination at the local,
national, and international level. In many nations, research pro-
grammes are emerging that will address the data gaps and research
identified below. International marine research organizations are
facilitating coordination and integration of national research at
the hemispheric or global level. A key element of the success of
these local, national, and international research collaborations
will be the formation of interdisciplinary research teams that
include earth system modellers, ecologists, fisheries scientists,
and fisheries managers who will work together to develop new
and improved projection capabilities for the future. We identify
the following key research needs.
Increased physiological measurements
Physiological measurements of keylife stages of all target marine fish
species are needed. Studies should examine the effects of multiple
factors on growth and bioenergetics (rates of energy losses and
gains). There is an urgent need to explore interactive effects
(temperature ×pH ×O
2
) on the survival and growth performance
in a variety of fish and invertebrates and to gain more data on the
growth physiology of all life stages. This will not only help in the
short term for linking physiological responses to statistically down-
scaled drivers but also in the long-term to build physiologically-
based models (Po
¨rtner and Peck, 2010;Jørgensen et al., 2012) that
can make use of dynamically downscaled forcing variables. Longer
term experiments are also needed (Denman et al., 2011) to gauge
the adaptive capacity of individuals and populations and test how
the sensitivity to climate-driven factors may change from one gen-
eration to the next. Operational techniques to incorporate physi-
ology directly into stock projection type models should be explored.
Integrated ecological monitoring to identify mechanisms
underlying fish and shellfish responses to environmental
drivers and fishing
Systematic ocean sampling of interacting physical, chemical, and
biological components must be continued to improve our under-
standing of the key climate-driven processes underlying observed
trends. The marine environment is chronically undersampled,
and we have limited historical time-series to gauge the past and
recent magnitude of natural variability (abundance, distribution)
of marine fish and shellfish resources relative to more recent
responses to multiple, anthropogenic stressors (climate, eutrophi-
cation, pollution, etc.). Efforts to establish a global network of
observations (e.g. distribution, growth) are particularly useful for
tracking climate change impacts on spatial distributions and abun-
dance. In addition, continued efforts to understand critical biomass
thresholds will be needed. Knowledge of the responses of key prey
fields (zooplankton and forage fish) to changes in ocean conditions
will be needed to adequately project shifts in the distribution and
abundance of exploited fish and shellfish stocks. Efforts to identify
cost-effective ways of augmenting existing fish and shellfish
surveys to collect information on these prey fields is needed to fill
existing gaps in knowledge for these species (e.g. Handegard et al.,
2012;Ressler et al., 2012). Maintenance and enhancement of fish
and shellfish consumption is also needed to adequately project
responses to shifting prey density and species composition.
Trophodynamic monitoring (e.g. combination of stomach contents
and isotope ratio) is also required to detect match– mismatch
changes with climate change in future.
Short-term forecasts (110 years) based on observed
ocean conditions
Short-term projections of biological responses using observed
ocean conditions are a powerful way to assess the predictive skill
of functional relationships. For physical models, these short-term
projections will allow analysts to test the models ability to capture
the correct physics. For harvested fish and shellfish stocks, this
may be part of routine stock assessments. Over time, results from
these skill assessments will provide the estimates of process error
for long-term projections.
Process studies to test functional relationships
Survival and growth efficiency of early life stages of marine fish and
invertebrates mostly ensures a formation of year-class productivity.
Despite a century of research, many key functional relationships
remain uncertain and they do not appear to be static. Studies of
bioenergetic responses to climate change and their effect on larval
and juvenile development (especially with respect to ocean acidifi-
cation), growth and reproduction are needed. Process studies of
species interactions including predator-prey responses to climate
change are also needed. Studies to identify the factors influencing
the distribution of juveniles would provide valuable information
for modelers.
Comparative studies to test hypotheses
Continued emphasis should be placed on identifying (and/or com-
paring) the drivers of recruitment variability between and within
species. Comparative analyses among stocks can reveal broad,
climate-related patterns in productivity (e.g. Dutil and Brander,
2003;Shuntov and Temnykh, 2011) that would otherwise be
elusive. Furthermore, continued process-oriented investigations
Projected impacts of climate change on marine fish and fisheries 1031
at Bibliothekssystem Universitaet Hamburg on January 2, 2017http://icesjms.oxfordjournals.org/Downloaded from
are necessary to reveal how various abiotic (temperature, salinity,
pH) and biotic (trophodynamic) factors interact with fishing
pressure to make populations most susceptible to climate-driven
changes. In terms of understanding recruitment, “non-stationarity”
is an important point to consider in understanding historical and
current recruitment drivers (Haltuch et al., 2009). Such information
should help identify how various factors contribute to changes in the
productivity and distribution of marine fish observed in the last two
to three decades (e.g. Rose, 2005;Rijnsdorp et al., 2010) and to make
more robust projections of future changes.
Improvement of ESMs and/or regional coupled biophysical
models
The horizontal resolution of some GCMs is too coarse to capture
shelf-region ocean processes. The spatial scales are not adequate
to resolve many of the important mesoscale structures such as
eddies, fronts, tides, and wind-driven upwelling that are important
for biological processes. This will require downscaling from GCMs
to more spatially resolved regional models. Although such regional
models are being developed, it is important that there be coupling
(one-way or two-way) between the regional and global models to
capture the correct physics.
Coupled biophysical projection models should be extended to
include the responses of fish and shellfish, fishers, and managers
to climate-driven change (Stock et al., 2011). New classes of
models that explore the synergy between climate change effects
and human activities are needed to provide meaningful and realistic
projections and to allow adaptation and mitigation measures and
their trade-offs, and to emerge from evolving management
systems (Barange et al., 2010).
Vulnerability assessments for fish, fisheries,
and fishery-dependent communities
Allison et al. (2009) provided an important preliminary estimate of
the vulnerability of countries to climate change impacts on fisheries.
The authors concluded that for countries depending on fisheries but
without sufficient capacity to adapt, climate-related changes in fish-
eries are likely to result in either greater economic hardships or to
those countries missing opportunities for maintaining or improv-
ing the benefits obtained from their fisheries. Further research is
required to increase the resolution of the results from the Allison
et al. (2009) study and to explore the opportunities and constraints
to adaptation in the most vulnerable countries in greater detail to
allow for targeted efforts to build adaptive capacity where it is
most needed and will yield the greatest benefits.
Coping strategies
As presented in this paper and in references included here, there is
considerable general information available on what adaptive strat-
egies are likely to be effective in response to climate-induced
changes in fisheries and aquaculture. However, to date, there are
very few examples of successful, or not so successful, implementa-
tion of adaptation strategies or plans in practice. There is an
urgent need to select cases, of diverse social and ecological character-
istics, where climate change is already having an impact on fisheries
and aquaculture social-ecological systems and to develop, imple-
ment, and monitor adaptation plans in accordance with current
best practices and understanding. This will allow the existing theor-
ies to be tested and improved where required from the lessons
learned. Issues of food security and marine conservation may
require new approaches to satisfy the growing demand for marine
resources.
ICES PICES strategic initiative
To coordinate and encourage research to address the some of the re-
search needs outlined in the previous section, the governing bodies
of both PICES and ICES approved the formation of the first joint
ICESPICES Strategic Initiative on Climate Change effects on
Marine Ecosystems (SICCME). The key deliverables for ICES and
PICES are the development of sufficient knowledge and under-
standing to successfully predict the future implications of climate
change on marine ecosystems and the ability to use this information
to develop strategies for managing living marine resources under a
changing climate. The SICCME is designed to facilitate and acceler-
ate the acquisition of new knowledge and to ensure that new knowl-
edge is communicated and published on a schedule that would allow
it to be useful to, and considered by, international scientificorgani-
zations responsible for providing advice on climate change such as
the IPCC and the United Nations.
Members of the SICCME will focus their work on four critical
issues:
(i) identifying techniques for predicting climate change impacts
in systems impacted by decadal variability,
(ii) defining the vulnerability of commercial species to climate
change and identifying which species would be most likely to
experience shifts in spatial distributions,
(iii) engaging the global earth system modelling community in
modelling climate change effects on marine ecosystems and
identifying opportunities for collaborations, and
(iv) building response scenarios for how the human community
will respond to climate changes as an extension (added dimen-
sion) of RCP scenarios described by van Vuuren et al. (2011).
The eight key research issues identified in this paper map into the
four SICCME critical issues as follows:
(i) SICCME Critical Issue a: research recommendations 2, 3, and 5
(ii) SICCME Critical Issue b: recommendations 1, 2, 3, 4, and 7
(iii) SICCME Critical Issue c: recommendation 6
(iv) SICCME Critical Issue d: recommendations 7 and 8
This suggests that the leading marine science organizations in the
northern hemisphere are well poised to facilitate advancements in
our ability to understand and project the effects of climate change
on marine ecosystems in the future. Their track record, to date, sug-
gests that partnerships between science organizations will lead to
more rapid global dissemination of research findings and analytical
approaches through workshops, symposiums, and publications.
Acknowledgements
We thank ICES, PICES, and IOC for their support and encourage-
ment to participate in symposiums focused on climate change
effects on marine ecosystems that were held in Sendai, Japan, in
2010 and Yeosu, Korea, in 2012. We thank Pat Livingston and
Mike Sigler for helpful comments and suggestions that improved
this manuscript. We also thank Nathan Ryan who helped to
compile the literature presented in Table 1.
1032 A. B. Hollowed et al.
at Bibliothekssystem Universitaet Hamburg on January 2, 2017http://icesjms.oxfordjournals.org/Downloaded from
References
ACIA, 2005. Arctic Climate Impact Assessment, Cambridge University
Press, New York, N.Y. 1042 pp.
Agostini, V. N., Hendriz, A. N., Hollowed, A. B., Wilson, C. D., Pierce,
S. D., and Francis, R. C. 2008. Climate-ocean variability and Pacific
hake: a geostatistical modeling approach. Journal of Marine Systems,
71: 237–248.
Ainsworth, C. H., Samhouri, J. F., Busch, D. S., Cheung, W. W. L.,
Dunne, J., and Okey, T. A. 2011. Potential impacts of climate
change on Northeast Pacific marine foodwebs and fisheries. ICES
Journal of Marine Science, 68: 12171229.
Alheit, J., Pohlmann, T., Casini, M., Greve, W.,Hinrichs, R., Mathis, M.,
O’Driscoll, K., et al. 2012. Climate variability drives anchovies and
sardines into the North and Baltic Seas. Progress in Oceanography,
96: 128–139.
Allison, E. H., Perry, A. L., Badjeck, M-C., Adger, W. N., Brown, K.,
Conway, D., Halls, A. S., et al. 2009. Vulnerability of national econ-
omies to potential impacts of climate change on fisheries. Fish and
Fisheries, 10: 173196.
AMAP. (Ed.) 2011. Snow Water Ice Permafrost in the Arctic, Executive
Summary. Arctic Monitoring and Assessment Program, Secretariat,
Oslo, Norway.
A’mar, Z. T., Punt, A. E., and Dorn, M. W. 2009. The evaluation of two
management strategies for the Gulf of Alaska walleye pollock fishery
under climate change. ICES Journal of Marine Science, 66:
1614–1632.
Anderson, P. J., and Piatt, J. F. 1999. Community reorganization in the
Gulf of Alaska following ocean climate regime shift. Marine Ecology
Progress Series, 189: 117123.
Barange, M., and Perry, R. I. 2009. Physical and ecological impacts
of climate change relevant to marine and inland capture fisheries
and aquaculture. In Climate Change Implications for Fisheries
and Aquaculture. Overview of Current Scientific Knowledge,
pp. 7–106. Ed. by K. Cochrane, C. De young, D. Soto, and T.
Bahri. FAO Fisheries and Aquaculture Technical Paper, 530. FAO,
Rome.
Barange, M., Cheung, W. W. L., Merino, G., and Perry, R. I. 2010.
Modelling the potential impacts of climate change and human activ-
ities on the sustainability of marine resources. Current Opinion in
Environmental Sustainability 2(5-6): 326–333.
Barange, M., et al. 2011. Predicting the impacts and socio-economic
consequences of climate change on global marine ecosystems and
fisheries: the QUEST fish framework Pages 31– 59. In World
Fisheries: a Social-Ecological Analysis. Ed. by R. E. Ommer, R. I.
Perry, K. Cochrane, and P. Cury. Wiley-Blackwell, Oxford, NY.
Baumgartner, T. R., Soutar, A., and Ferreira-Bartrina, V. 1992.
Reconstruction of the history of Pacific sardine and northern
anchovy populations over the past two millennia from sediments
of the Santa Barbara basin. CALCOFI Reports, 33: 24 –40.
Beare, D. J., Burns, F., Greig, A., Jones, E. G., Peach, K., Kienzle, M.,
McKenzie, E., et al. 2004. Long-term increases in prevalence of
North Sea fishes having southern biogeographic affinities. Marine
Ecology Progress Series, 284: 269– 278.
Beaugrand, G., and Kirby, R. R. 2010. Climate, plankton and cod. Global
Change Biology, 16: 1268– 1280.
Bell, J. D., Kronen, M., Vunisea,A., Nash, W. J., Keeble, G., Demmke, A.,
Pontifex, S., et al. 2009. Planning the use of fish for food security in
the Pacific. Marine Policy, 33: 64– 76.
Blanchard, J. L., Jennings, S., Holmes, R., Harle, J., Merino, G., Allen, J.
I., Holt, J., et al. 2012. Potential consequences of climate change for
primary production and fish production in large marine ecosystems.
Philosophical Transactions of the Royal Society B, 367: 2979– 2989.
Bode, A., Hare, J. A., Li, W. K. W., Mora
´n, X. A. G., and Valde
´s, L. 2011.
Chlorophyll and primary production in the North Atlantic. In ICES
Status Report on Climate Change in the North Atlantic, pp. 77 –102.
Ed. by P. C. Reid, and L. Valde
´s. ICES Cooperative Research Report,
310. 262 pp.
Brander, K. 2010. Impacts of climate change on fisheries. Journal of
Marine Systems, 79: 389– 402.
Brander, K. 2012. Climate and current anthropogenic impacts on fish-
eries. Climatic Change. doi:10.1007/s10584-012-0541-2.
Burrows, M. T.,Schoeman, D. S., Buckley, L. B., Moore, P., Poloczanska,
E. S., Brander, K. M., Brown, C., et al. 2011. The pace of shifting
climate in marine and terrestrial ecosystems. Science, 334: 652 –655.
Capotondi, A., Alexander, M. A., Bond, N. A., Curchitser, E. N., and
Scott, J. D. 2012. Enhanced upper ocean stratification with climate
change in the CMIP3 models. Journal of Geophysical Research:
Oceans, 117.
Ciannelli, L., and Bailey, K. M. 2005. Landscape dynamics and resulting
species interactions: the cod-capelin system in the southeastern
Bering Sea. Marine Ecology Progress Series, 291: 227 –236.
Chavez, F. P., Ryan, J., Lluch-Cota, S. E., and Niquen, C. M. 2003. From
anchovies to sardines and back: multidecadal change in the Pacific
Ocean. Science, 299: 217221.
Cheung, W. W. L., Dunne, J., Sarmiento, J. L., and Pauly, D. 2011.
Integrating ecophysiology and plankton dynamics into projected
maximum fisheries catch potential under climate change in the
Northeast Atlantic. ICES Journal of Marine Science, 68: 1008 –1018.
Cheung, W. W. L., Lam, V. W. Y., Sarmiento, J. L., Kearney, K., Watson,
R., and Pauly, D.2009. Projecting global marine biodiversity impacts
under climate change scenarios. Fish and Fisheries, 10: 235 –251.
Cheung, W. W. L., Lam, V. W. Y., Sarmiento, J. L., Kearney, K., Watson,
R., Zeller, D., and Pauly, D. 2010. Large-scale redistribution of
maximum fisheries catch potential in the global ocean under
climate change. Global Change Biology, 16: 24–35.
Chiba, S., Ono, T., Tadokoro, K., et al. 2004. Increased stratification and
decreased lower trophic level productivity in the Oyashio region of
the North Pacific: a 30-year retrospective study. Journal of
Oceanography, 60: 149– 62.
Cochrane, K., De Young,C., Soto, D., and Bahri, T. 2009. Climatechange
implications for fisheries and aquaculture. FAO Fisheries and aqua-
culture technical paper, 530: 212.
Cochrane, K. L., Andrew,N. L., and Parma, A. M. 2011. Primary fisheries
management: a minimum requirement for provision of sustainable
human benefits in small-scale fisheries. Fish and Fisheries, 12:
275–288.
Costello, C., Ovando, D., Hilborn, R., Gaines, S. D., Deschenes, O., and
Lester, S. E. 2012. Status and solutions for world’s unassessed fisher-
ies. Science, 338: 517520.
Criddle, K. R. 2012. Adaptation and maladaptation: factors that influ-
ence the resilience of four Alaskan fisheries governed by durable enti-
tlements. ICES Journal of Marine Science, 69: 1168 1179.
Daw, T., Adger, W. N., Brown, K., and Badjeck, M-C. 2009. Climate
change and capture fisheries: potential impacts, adaptation and
mitigation. In Climate Change Implications for Fisheries and
Aquaculture. Overview of Current Scientific Knowledge, pp.
107–150. Ed. by K. Cochrane, C. De young, D. Soto, and T. Bahri.
FAO Fisheries and Aquaculture Technical Paper, 530. FAO, Rome.
530 pp.
Denman, K., Christian, J. R., Steiner, N., Po
¨rtner, H. O., and Nojiri, Y.
2011. Potential impacts of future ocean acidification on marine eco-
systems and fisheries: current knowledge and recommendations for
future research. ICES Journal of Marine Science, 68: 1019 –1029.
Deser, C., Phillips, A., Bourdette, V., and Teng, H. 2012. Uncertainty in
climate change projections: the role of internal variability. Climate
Dynamics, 38: 527–546.
Di Lorenzo, E., Schneider, N., Cobb, K. M., Franks, P. J. S., Chhak, K.,
Miller, A. J., McWilliams, J. C., et al. 2008. North Pacific gyre oscil-
lation links ocean climate and ecosystem change. Geophysical
Research Letters, 35: L08607.
Donelson, J. M., Munday, P. L., McCormick, M. I., and Nilsson, G. E.
2011. Acclimation to predicted ocean warming through develop-
mental plasticity in a tropical reef fish. Global Change Biology, 17:
1712–1719.
Projected impacts of climate change on marine fish and fisheries 1033
at Bibliothekssystem Universitaet Hamburg on January 2, 2017http://icesjms.oxfordjournals.org/Downloaded from
Doney, S. C., Ruckelshaus, M., Duffy,J. E., Barr y, J. P., Chan, F., English,
C. A., Galindo, H. M., et al. 2012. Climate change impacts on marine
ecosystems. Annual Reviews in Marine Science, 4: 1–27.
Drinkwater, K. F. 2005. The response of Atlantic cod (Gadus morhua)to
future climate change. ICES Journal of Marine Science, 62:
1327–1337.
Drinkwater, K. F. 2006. The regime shift of the 1920s and 1930s in the
North Atlantic. Progress in Oceanography, 68: 134– 151.
Drinkwater, K. F., Beaugrand, G., Kaeriyama, M., Kim, S., Ottersen, G.,
Perry, R. I., Po
¨rtner, H. J., et al. 2010. On the processes linking climate
to ecosystem change. Journal of Marine Systems, 79: 374388.
Drinkwater, K. F., Hunt, G. L., Astthorsson, O. S., and Head, E. J. H.
2012. Comparative studies of climate effects on polar and subpolar
ocean ecosystems, progress in observation and prediction: an intro-
duction. ICES Journal of Marine Science, 69: 1120 1122.
Duarte, C. M., Agustı
´, S., Wassmann, P., Arrieta, J. M., Alcaraz, M.,
Coello, A., Marba
´, N., et al. 2012. Tipping elements in the Arctic
Marine Ecosystem. AMBIO, 41: 44– 55.
Dulvy, N. K., Rogers, S. I., Jennings, S., Stelzenmu
¨ller, V., Dye, S. R., and
Skjoldal, H. R. 2008. Climate change and deepening of the North Sea
fish assemblage: a biotic indicator of warming seas. Journal of
Applied Ecology, 45: 4.
Dutil, J-D., and Brander, K. 2003. Comparing productivity of North
Atlantic cod (Gadus morhua) stocks and limits to growth produc-
tion. Fisheries Oceanography, 12: 502– 512.
Edwards, M., and Richardson, A. J. 2004. Impact of climate change on
marine pelagic phenology and trophic mismatch. Nature, 430:
881–884.
Eero, M., MacKenzie, B. R., Ko
¨ster, F. W., and Gislason, H. 2011.
Multi-decadal responses of a cod (Gadus morhua) population to
human-induced trophic changes, exploitation and climate variabil-
ity. Ecological Applications, 21: 214– 226.
Engelhard, G. H., Pinnegar, J. K., Kell, L. T., and Rijnsdorp, A. D. 2011.
Nine decades of North Sea sole and plaice distribution. ICES Journal
of Marine Science, 68: 10901104.
FAO. 2008a. Climate change for fisheries and aquaculture. Technical
Background document from the Expert Consultation on Climate
Change for Fisheries and Aquaculture, (FAO), Rome, 7–9 April
2008.
FAO. 2008b. Report of the FAO Expert Workshop on Climate Change
Implications for Fisheries and Aquaculture. FAO Fisheries Report,
870, FAO, Rome, 7– 9 April 2008. 32 pp.
FAO. 2009. FAO Technical Guidelines for Responsible Fisheries
Supplement 4, Addition 2, Fisheries Management 2. The ecosystem
approach to management, 2.2 The Human Dimensions of the eco-
system approach to management. FAO, Rome. 88 pp.
FAO. 2012. The State of World Fisheries and Aquaculture 2012. Rome.
209 pp.
Feely, R. A., Sabine, C. L., Hernandez-Ayon, J. M., Ianson, D., and Hales,
B. 2008. Evidence for upwelling of corrosive ‘acidified’ water onto
the continental shelf. Science, 320: 1490 1492.
Fernandes, J. F., Cheung, W. L., Jennings, S., Butenscho
¨n, M., de Mora,
L., Fro
¨licher, T.L., Barange, M., et al. in press. Modelling the effects of
climate change on the distribution and production of marine fishes
for an ecosystem approach to fisheries management: accounting for
trophic interactions in a dynamic bioclimate envelope model. Global
Change Biology, doi: 10.1111/gcb.12231.
Fiechter, J. 2012. Assessing marine ecosystem model properties from en-
semble calculations. Ecological Modelling, 242: 164– 179.
Fiechter, J., Moore, A. M., Edwards, C. A., Bruland, K. W., Di Lorenzo,
E., Lewis, C. V. W., Powell, T. M., et al. 2009. Modeling iron limita-
tion of primary production in the coastal Gulf of Alaska. Deep Sea
Research II, 56: 2503– 2519.
Fogarty, M., Incze, L., Hayhoe, K., Mountain, D., and Manning, J. 2008.
Potential climate impacts on Atlantic cod (Gadus morhua) off the
northeastern USA. Mitgation Adaptation Strategies for Global
Change, 13: 453466.
Fulton, E. A. 2011. Interesting times: winners, losers, and system shifts
under climate change around Australia. ICES Journal of Marine
Science, 68: 13291342.
Garcia, S. M., and Rosenberg, A. A. 2009. UK Future Foresight: Food
Security and Marine Capture Fisheries: Characteristics of the
Sector and Perspectives on Emerging Issues. Defra, London. 52 pp.
Genner, M. J., Sims, D. W.,Wearmouth, V. J., Southall, E. J., Southward,
A. J., Henderson, P. A., and Hawkins, S. 2004. Regional climate
warming drives long-term community changes of British marine
fish. Proceedings of the Royal Society B: Biological Sciences, 271:
655–661.
Gibson, G. A., and Spitz, Y. H. 2011. Impacts of biological parameteriza-
tion, initial conditions, and environmental forcing on parameter
sensitivity and uncertainty in a marine ecosystem model for the
Bering Sea. Journal of Marine Systems, 88: 214–231.
Gnanadesikan, A., Dunne, J. P., and John, J. 2011. What ocean biogeo-
chemical models can tell us about bottom-up control of ecosystem
variability. ICES Journal of Marine Science, 68: 1030–1044.
Godfray, H. C. 2011. Food for thought. Proceedings of the National
Academy of Sciences of the USA, 108: 19845–19846.
Godfray, H. C., Beddington, J. R., Crute, I. R., Haddad, L., Lawrence,D.,
Muir, J. F., Pretty, J., et al. 2010. Food security: the challenge of
feeding 9 billion people. Science, 327: 812–818.
Greene, C. H., and Pershing, A. J. 2007. Climate drives sea change.
Science, 315: 10841085.
Gutierrez, N. L., Hilborn, R., and Defeo, O. 2011. Leadership, social
capital and incentives promote successful fisheries. Nature, 470:
386–389.
Halpern, B. S., Longo, C., Hardy, D., McLeod, K. L., Samhouri, J. F.,
Katona, S. K., Kleisner, K., et al. 2012. An index to assess the health
and benefits of the global ocean. Nature, 488: 615–620.
Haltuch, M. A., Punt, A. E., and Dorn, M. W. 2009. Evaluating the esti-
mation of fishery management reference points in a variable envir-
onment. Fisheries Research, 100: 43–56.
Handegard, N. O., du Buisson, L., Brehmer, P., Chalmers, S. J., De
Robertis, A., Huse, G., Kloser, R., et al. 2012. Towards an acoustic-
based coupled observation and modeling system for monitoring
and predicting ecosystem dynamics of the open ocean. Fish and
Fisheries, doi:10.1111/j.1467-2979.2012.00480.x.
Hare, J. A., Alexander, M. A., Fogarty, M. J., Williams, E. H., and Scott,
J. D. 2010a. Forecasting the dynamics of a coastal fishery species
using a coupled climate-population model. Ecological
Applications, 20: 452–464.
Hare, J. A., Manderson, J., Nye, J., Alexander, M., Auster, P. J.,
Borggaard, D., Capotondi, A., et al. 2012a. Cusk (Brosme brosme)
and climate change: assessing the threat to a candidate marine fish
species under the U.S. Endangered Species Act. ICES Journal of
Marine Science, 69: 17531768.
Hare, J. A., Wuenschel,M. J., and Kimball, M. E. 2012b. Projecting range
limits with coupled thermal tolerance - climate change models: an
example based on gray snapper (Lutjanus griseus) along the U.S.
east coast. PLoS One, 7: e52294.
Hare, S. R., and Mantua, N. J. 2000. Empirical evidence for North Pacific
regime shifts in 1977 and 1989. Progress in Oceanography, 47:
103–145.
Hashioka, T., and Yamanaka, Y. 2007. Ecosystem change in the western
North Pacific associated with global warming using 3D-NEMURO.
Ecological Modelling, 202: 95–104.
Ha
´tu
´n, H., Payne, M. R., Beaugrand, G., Reid, P. C., Sandø, A. B.,
Drange, H., Hansen, B., et al. 2009. Large bio-geographical shifts
in the north eastern Atlantic Ocean: from the subpolar gyre, via
plankton, to blue whiting and pilot whales. Progress in
Oceanography, 80: 149– 162.
Haynie, A. C., and Pfeiffer, L. 2012. Why economics matters for under-
standing the effects of climate change on fisheries. ICES Journal of
Marine Science, 69: 11601167.
1034 A. B. Hollowed et al.
at Bibliothekssystem Universitaet Hamburg on January 2, 2017http://icesjms.oxfordjournals.org/Downloaded from
Hawkins, E., and Sutton, R. 2009. The potential to narrow uncertainty in
regional climate predictions. Bulletin of the American
Meteorological Society, 90: 1095– 1107.
Helser, T.E., Lai, H-l., and Black, B. A. 2012. Bayesian hierarchical mod-
eling of Pacific geoduck growth increment data and climate indices.
Ecological Modelling, 247: 210–220.
Hofstede, T. R., Hiddink, J. G., and Rijnsdorp, A. D. 2010. Regional
warming changes fish species richness in the eastern North
Atlantic Ocean. Marine Ecology Progress Series, 414: 1– 9.
Hofstede, T. R., and Rijnsdorp, A. D. 2011. Comparing demersal fish
assemblages between periods of contrasting climate and fishing pres-
sure. ICES Journal of Marine Science, 68: 1189 –1198.
Hollowed, A. B., Barange, M., Ito, S., Kim, S., and Loeng, H. 2011.Effects
of climate change on fish and fisheries: forecasting impacts, assessing
ecosystem responses, and evaluating management strategies Preface.
ICES Journal of Marine Science, 68: 984 985.
Hollowed, A. B., Bond, N. A., Wilderbuer, T. K., Stockhausen, W. T.,
A’mar, Z. T., Beamish, R. J., Overland, J. E., et al. 2009. A framework
for modelling fish and shellfish responses to future climate change.
ICES Journal of Marine Science, 66: 1574 1594.
Hollowed, A. B., Curchitser, E., Stock, C., and Zhang, C. I. in press a.
Trade-offs associated with different modeling approaches for assess-
ment of fish and shellfish responses to climate change. Climatic
Change, doi:10.1007/s10584-012-0641-z.
Hollowed, A. B., Planque, B., and Loeng, H. in press b. Potential move-
ment of fish and shellfish from the Sub-Arctic to the Arctic ocean.
Fisheries Oceanography, doi:10.1111/fog.12027.
Holt, J., Butenschon, M., Wakelin, S. L., Artioli, Y., and Allen, J. I. 2012.
Oceanic controls on the primary production of the northwest
European continental shelf: model experiments under recent past
conditions and a potential future scenario. Biogeosciences, 9:
97117.
Howard, J., Babij, E., Griffis, R., Helmuth, B., Himes-Cornell, A.,
Niemier, P., Orbach, M., et al. 2013. Oceans and marine resources
in a changing climate. Oceanography and Marine Biology: An
Annual Review, 51: 71– 192.
Hsieh, C-H., Kim, H. J., Watson, W., Di Lorenzo, E., and Sugihara, G.
2009. Climate-driven changes in abundance and distribution of
larvae of oceanic fishes in the southern California region. Global
Change Biology, 15: 2137– 2152.
Huse, G.,and Ellingsen, I. 2008.Capelin migrations and climatechange
a modeling analysis. Climatic Change, 87: 177– 197.
Ianelli, J., Hollowed, A., Haynie, A., Mueter, F. J., and Bond, N. A. 2011.
Evaluating management strategies for eastern Bering Sea walleye
pollock (Theragra chalcogramma) in a changing environment.
ICES Journal of Marine Science, 68: 1297 1304.
IPCC, and . 2007. Climate Change 2007: Synthesis Report. In
Contribution of Working Groups I, II and III to the fourth
Assessment Report of the Intergovernmental Panel on Climate
Change. Ed. by R. K. Pachauri, A. Reisinger, and Core Writing
Team. IPCC, Geneva, Switzerland. 104 pp.
Ito, S., Okunishi, T., Kishi, M. J., and Wang, M. 2013. Modeling eco-
logical responses of Pacific saury (Cololabis saira) to future climate
change and its uncertainty. ICES Journal of Marine Science, 70:
980–990.
Ito, S., Rose, K. A., Miller, A. J., Drinkwater, K., Brander, K. M.,
Overland, J. E., Sundby, S., et al. 2010. Ocean ecosystem responses
to future global change scenarios: a way forward. In Global Change
and Marine Ecosystems, pp. 287322. Ed. by M. Barange, J. G.
Field, R. H. Harris, E. Hofmann, R. I. Perry, and F. Werner.
Oxford University Press. 440 pp.
Jang, C. J., Park, J., Park, T., and Yoo, S. 2011. Response of the ocean
mixed layer depth to global warming and its impact on primary pro-
duction: a case for the North Pacific Ocean. ICES Journal of Marine
Science, 68: 9961007.
Jones, M.C., Dye, S.R., Pinnegar, J.K., Warren, R., and Cheung, W. W. L.
2012. Modelling commercial fish distributions: prediction and
assessment using different approaches. Ecological Modeling, 225:
133–145.
Jørgensen, C., Peck, M. A., Antognarelli, F.,Azzurro, E., Burrows, M. T.,
Cheung, W. W. L., Cucco, A., et al. 2012. Conservation physiology of
marine fishes: advancing the predictive capacity of models. Biology
Letters, 8: 900903.
Kim, S. 2010. Fisheries development in northeastern Asia in conjunc-
tion with changes in climate and social systems. Marine Policy, 34:
803–809.
King, J. R., Agostini, V. N., Harvey, C. J., McFarlane, G. A., Foreman,
M. G. G., Overland, J. E., Di Lorenzo, E., et al. 2011. Climate
forcing and the California Current Ecosystem. ICES Journal of
Marine Science, 68: 11991216.
Kingsolver, J. G. 2009. The well-temperatured biologist. The American
Naturalist, 174: 755– 768.
Kristiansen, T., Drinkwater, K. F., Lough, R. G., and Sunby, S. 2011.
Recruitment variability in North Atlantic cod and match – mismatch
dynamics. PLoS One, 6: e17456.
Last, P. R., White, W. T., Gledhill, D. C., Hobday,A. J., Brown, R., Edgar,
G. J., and Pecl, G. 2011. Long-term shifts in abundance and distribu-
tion of a temperate fish fauna: a response to climate change and
fishing practices. Global Ecology and Biogeography, 20: 5872.
Laurenti, G. (comp.). 2007. 1961 –2003 Fish and fishery products: world
apparent consumption statistics based on food balance sheets. FAO
Fisheries Circular No.i821, rev.9. FAO, Rome. 429 pp.
Lehodey, P., Senina, I., Calmettes, B., Hampton, J., and Nicol, S. 2012.
Modelling the impact of climate change on Pacific skipjack tuna
population and fisheries. Climatic Change, doi:10.1007/
s10584-012-0595-1.
Levin, L. A., Fogarty, M. J., Murawski, S. A., and Fluharty, D. 2009.
Effects of natural and human-induced hypoxia on coastal benthos.
Biogeosciences, 6: 2063–2098.
Lindegren, M., Diekmann, R., and Mollmann, C. 2010. Regime shifts,
resilience and recovery of a cod stock. Marine Ecology Progress
Series, 402: 239253.
Link, J. S., Ihde, T. F., Harvey, C. J., Gaichas, S. K., Field, J. C., Brodziak,
J. K. T., Townsend, H. M., et al. 2012. Dealing with uncertainty in
ecosystem models: the paradox of use for living marine resource
management. Progress in Oceanography, 102: 102– 114.
Loeng, H., and Drinkwater, K. 2007. An overview of the ecosystems of
the Barents and Norwegian Seas and their response to climate vari-
ability. Deep Sea Research II, 54: 2478– 2500.
Loukos, H., Monfray, P., Bopp, L., and Lehodey, P. 2003. Potential
changes in skipjack tuna (Katsuwonus pelamis) habitat from a
global warming scenario: modeling approach and preliminary
results. Fisheries Oceanography, 12: 474– 482.
Lucey, S. M., and Nye, J. A. 2010. Shifting species assemblages in the
northeast US continental shelf large marine ecosystem. Marine
Ecology Progress Series, 415: 23– 33.
Mackas, D. L., Batten, S., and Trudel, M. 2007. Effects on zooplankton of
a warmer ocean: recent evidence from the North Pacific. Progress in
Oceanography, 75: 223– 52.
MacKenzie, B. R., Meier, H. E. M., Lindegren, M., Neuenfeldt, S., Eero,
M., Blenckner, T.,Tomczak, M., et al. 2012. Impact of climate change
on fish population dynamics in the Baltic Sea – a dynamical down-
scaling investigation. Ambio, 41: 626–646. doi:10.1007/s13280-
012-0325-y.
Maunder, M. N., Harley, S. J., and Hampton, J. 2006. Including param-
eter uncertainty in forward projections of computationally intensive
statistical dynamic models. ICES Journal of Marine Science, 63:
969–979.
Megrey, B. A., Link, J. S., Hunt, G. L., Jr, and Moksness, E. 2009.
Comparative marine ecosystem analysis: applications, opportun-
ities, and lessons learned. Progress in Oceanography, 81: 29.
Meier, H. E. M., Andersson, H. C., Arheimer, B., Blenckner, T.,
Chubarenko, B., Donnelly, C., Eilola, K., et al. 2012. Comparing
reconstructed past variations and future projections of the Baltic
Projected impacts of climate change on marine fish and fisheries 1035
at Bibliothekssystem Universitaet Hamburg on January 2, 2017http://icesjms.oxfordjournals.org/Downloaded from
Sea ecosystem—first results from multi-model ensemble simula-
tions. Environmental Research Letters, 7: 034005 (8pp).
Merino, G., Barange, M., Blanchard, J., Harle, J., Holmes, R., Allen, I.,
Allison, E. H., et al. 2012. Can marine fisheries and aquaculture
meet fish demand from a growing human population in a changing
climate? Global Environmental Change, 22: 795– 806.
Miller, K., Anthony,C., Barange, M., Brander, K., Gallucci, V. F., Gasalla,
M. A., Khan, A., et al. 2010. Climate change, uncertainty, and resili-
ent fisheries: institutional responses through integrative science.
Progress in Oceanography, 87: 338– 346.
Mora
´n, X. A. G., Lo
´pez-Urrutia, A
´., Calvo-Dı
´az, A., and Li, W. K. W.
2010. Increasing importance of small phytoplankton in a warmer
ocean. Global Change Biology, 16: 1137– 1144.
Mueter, F. J., Bond, N. A., Ianelli, J. N., and Hollowed, A. B. 2011.
Expected declines in recruitment of walleye pollock (Theragra chal-
cogramma) in the eastern Bering Sea under future climate change.
ICES Journal of Marine Science, 68: 1284 1296.
Mueter, F. J., and Litzow, M. A. 2008. Sea ice retreat alters the biogeog-
raphy of the Bering Sea continental shelf. Ecological Applications,
18: 309–320.
Mueter, F. J., Peterman, R. M., and Pyper, B. J. 2002. Opposite effects of
ocean temperature on survival rates of 120 stocks of Pacific salmon
(Oncorhynchus spp.) in northern and southern areas. Canadian
Journal of Fisheries and Aquatic Sciences, 59: 456463.
Muhling, B. A., Lee, S-K., Lamkin, J. T., and Liu, Y. 2011. Predicting the
effects of climate change on bluefin tuna (Thunnus thynnus) spawn-
ing habitat in the Gulf of Mexico. ICES Journal of Marine Science, 68:
1051–1052.
Murawski, S. A. 2007. Ten myths concerning ecosystem approaches to
marine resources management. Marine Policy, 31: 681– 690.
Murawski, S. A., et al. 2010. Why compare marine ecosystems? ICES
Journal of Marine Science, 67: 19.
Nye, J. A., Link, J. S., Hare, J. A., and Overholtz, W. J. 2009. Changing
spatial distribution of fish stocks in relation to climate and popula-
tion size in the Northeast United States continental shelf. Marine
Ecology Progress Series, 393: 111– 139.
OECD. 2008. OECD Environmental Outlook to 2030. Paris, France,
www.oecd.org/env.
OECD-FAO. 2009. OECD-FAO Agricultural Outlook 2009– 2018.
OECD, FAO, Paris, Rome. 95 pp. http://www.fao.org/docrep/
x5738e/x5738e0g.htm.
Okunishi, T., Ito, S., Hashioka, T., Sakamoto, T. T., Yoshie, N., Sumata,
H., Yara, Y., et al. 2012. Impacts of climate change on growth, migra-
tion and recruitment success of Japanese sardine (Sardinops mela-
nostictus) in the western North Pacific. Climatic Change,
doi:10.1007/s10584-012-0484-7.
Ottersen, G., Hjermann, D., and Stenseth, N. C. 2006. Changes in
spawning stock structure strengthens the link between climate and
recruitment in a heavily fished cod stock. Fisheries Oceanography,
15: 230–243.
Overland, J. E., and Wang, M. 2007. Future climate of the North Pacific
Ocean. EOS Transactions of the AGU, 88: 182.
Overland, J. E., Alheit, J., Bakun, A., Hurrell, J. W., Mackas, D. L., and
Miller, A. J. 2010. Climate controls on marine ecosystems and fish
populations. Journal Marine Systems, 79: 305– 315.
Peck, M. A., and Hufnagl, M. 2012. Can IBMs explain why most larvae
die in the sea? Model scenarios and sensitivity analyses reveal re-
search needs. Journal of Marine Systems, 93: 77 93.
Perrette, M., Yool, A., Quartly, G. D.,and Popova, E. E. 2011. Near ubi-
quity of ice edge bloom in the Arctic. Biogeosciences, 8: 515524.
Perry, A. L., Low, P. J., Ellis, J. R., and Reynolds, J. D. 2005. Climate
change and distribution shifts in marine fishes. Science, 308:
1912–1915.
Perry, R. I., Ommer, R. E., Barange, M., Jentoft, S., Neis, B., and Sumaila,
U. R. 2011. Marine social ecological responses to environmental
change and the impacts of globalization. Fish and Fisheries, 12:
427–450.
Petitgas, P., Alheit, J., Peck, M. A., Raab, K., Irigoien, X., Huret, M., van
der Kooij, J., et al. 2012. Anchovy population expansion in the North
Sea. Marine Ecology Progress Series, 444: 1 –13.
Petitgas, P., Rijnsdorp, A. D., Dickey-Collas, M., Engelhard, G. H., Peck,
M. A., Pinnegar, J. K., Drinkwater, K., et al. 2013. Impacts of climate
change on the complex life cycles of fish. Fisheries Oceanography, 22:
121–139.
Plaga
´nyi, E
´. E., Bell, J. D., Bustamante, R. H., Dambacher, J. M., Dennis,
D. M., Dichmont, C. M., Dutra, L. X. C., et al. 2011. Modelling
climate-change effects on Australian and Pacific aquatic ecosystems:
a review of analytical tools and management implications. Marine
and Freshwater Research, 62: 1132– 1147.
Planque, B., Bellier, E., and Loots, C. 2011. Uncertainties in projecting
spatial distributions of marine populations. ICES Journal of
Marine Science, 68: 10451050.
Polovina, J. J., Dunne, J. P., Woodworth, P. A., and Howell, E. A. 2011.
Projected expansion of the subtropical biome and contraction of the
temperate and equatorial upwelling biomes in the North Pacific
under global warming. ICES Journal of Marine Science, 68:
986–995.
Po
¨rtner, H. O. 2010. Oxygen- and capacity-limitation of thermal toler-
ance: a matrix for integrating climate-related stressor effects in
marine ecosystems. Journal of Experimental Biology, 213: 881 –893.
Po
¨rtner, H. O., and Peck, M. A. 2010. Climate change impacts on fish
and fisheries: towards a cause and effect understanding. Journal of
Fish Biology, 77: 1745– 1779.
Prince, E. D., et al. 2010. Ocean scale hypoxia-based habitat compres-
sion of Atlantic Istiophorid billfishes. Fisheries Oceanography, 19:
448462.
Rabalais, N. N., Diaz, R. J., Levin, L. A., Turner, R. E., Gilbert, D., and
Zhang, J. 2010. Dynamics and distribution of natural and human-
caused hypoxia. Biogeosciences, 7: 585– 619.
Racherla, P.N., Shindell, D. T., and Faluvegi, G. S. 2012. The added value
to global model projections of climate change by dynamical down-
scaling: A case study over the continental U.S. using the
GISS-ModelE2 and WRF models. Journal of Geophysical
Research, 117: 1– 8.
Reichler, T., and Kim, J. 2008. How well do coupled models simulate
today’s climate? Bulletin American Meteorological Society, 89:
303–311.
Ressler, P. H., De Robertis, A., Warren, J. D., Smith, J. N., and Kotwicki,
S. 2012. Developing an acoustic survey of euphausiids to understand
trophic interactions in the Bering Sea ecosystem. Deep Sea Research
II, 6570: 184– 195.
Rice, J. 2001. Implications of variability on many time scales for scien-
tific advice on sustainable management of living marine resources.
Progress in Oceanography, 49: 189– 209.
Rice, J. C., and Garcia, S. M. 2011. Fisheries, food security, climate
change, and biodiversity: characteristics of the sector and perspec-
tives on emerging issues. ICES Journal of Marine Science, 68:
1343–1353.
Richardson, A. J. 2008. In hot water: zooplankton and climate change.
ICES Journal of Marine Science, 65: 279 295.
Richardson, A. J., and Schoeman, D. S. 2004. Climate impact on plank-
ton ecosystems in the Northeast Atlantic. Science, 305: 16091612.
Ries, J. B., Cohen, A. L., and McCorkle, D. C. 2009. Marine calcifiers
exhibit mixed responses to CO2-induced ocean acidification.
Geology, 37: 1131– 1134.
Rijnsdorp, A. D., Peck, M. A., Engelhard, G. H., Mo
¨llmann, C., and
Pinnegar, J. K. (Eds). 2010. Resolving Climate Impacts on Fish
Stocks. ICES Cooperative Research Report, 301.
Rochet, M-J., and Rice, J. C. 2009. Simulation-based management strat-
egy evaluation: ignorance disguised as mathematics? ICES Journal of
Marine Science, 66: 754762.
Rogers, L. A., Stige, L. C., Olsen, E. M., Knutsen, H., Chan, K. S., and
Stenseth, N. C. 2011. Climate and population density drive
changes in cod body size throughout a century on the Norwegian
1036 A. B. Hollowed et al.
at Bibliothekssystem Universitaet Hamburg on January 2, 2017http://icesjms.oxfordjournals.org/Downloaded from
coast. Proceedings of the National Academic of Sciences of the USA,
108: 1961–1966.
Rose, G. A. 2005. On distributional responses of North Atlantic fish to
climate change. ICES Journal of Marine Science, 62: 1360 –1374.
Ruckelshaus, M., Doney, S. C., Glindo, H. M., Barry, J., Chan, F., Duffy,
J. E., English, C. A., et al. 2013. Securing ocean benefits for society in
the face of climate change. Marine Policy, 40: 154 159.
Rykaczewski, R. R., and Dunne, J. P. 2010. Enhanced nutrient supply to
the California Current Ecosystem with global warming and
increased stratification in an earth system model. Geophysical
Research Letters, 37: L21606.
Salinger, M. J., Bell, J. D., Evans, K., Hobday, A. J., Allain, V., Brander, K.,
Dexter, P., et al. in press. Climate and oceanic fisheries: Cook Islands
2011 Workshop conclusionsand recommendations. Climatic Change.
Shoji, J., Toshito,S., Mizuno, K., Kamimura,Y., Hori, M., and Hirakawa,
K. 2011. Possible effects of global warming on fish recruitment: shifts
in spawning season and latitudinal distribution can alter growth of
fish early life stages through changes in daylength. ICES Journal of
Marine Science, 68: 11651169.
Shuntov, V. P., and Temnykh, O. S. 2011. Pacific salmon in marine and
oceanic ecosystems. V.2. TINRO-Center: Vladivostok, 2011. 473 pp.
(In Russian).
Skud, B. E. 1982. Dominance in fishes: the relation between environ-
ment and abundance. Science, 216: 144149.
Southward, A. J., Boalch, G. T., and Maddock, L. 1988. Fluctuations in
the herring and pilchard fisheries of Devon and Cornwall linked to
change in climate since the 16th century. Journal of the Marine
Biological Association of the UK, 68: 423445.
Spencer, P. D. 2008. Density-independent and density-dependent
factors affecting temporal changes in spatial distributions of
eastern Bering Sea flatfish. Fisheries Oceanography, 17: 396 –410.
Steinacher, M., Joos, F., Frolicher, T. L., Bopp, L., Cadule, P., Cocco, V.,
Doney, S. C., et al. 2010. Projected 21st century decrease in marine
productivity: a multi-model analysis. Biogeosciences, 7: 979 1005.
Stenseth, N. C., Mysterud, A., Ottersen, G., Hurrell, J. W., Chan, K-S.,
and Lima, M. 2002. Ecological effects of climate fluctuations.
Science, 297: 12921296.
Stewart, J. S., Hazen, E. L., Foley, D. G., Bograd, S. J., and Gilly, W. F.
2012. Marine predator migration during range expansion:
Humboldt squid Dosidicus gigas in the northern California
Current System. Marine Ecology Progressive Series, 471: 135– 150.
Stock, C. A., Alexander, M. A., Bond, N. A., Brander, K. M., Cheung,
W. W. L., Curchister, E. N., Delworth, T. L., et al. 2011. On the use
of IPCC-class models to assess the impact of climate on Living
Marine Resources. Progress in Oceanography, 88: 1– 27.
Stramma, L., Prince, E. D., Schmidtko, S., Luo, J., Hoolihan, J. P.,
Visbeck, M., Wallace, D. W. R., et al. 2011. Expansion of oxygen
minimum zone may reduce available habitat for tropical pelagic
fishes. Nature Climate Change, 2: 3337.
Su, N-J., Sun, C-L., Punt, A. E., Yeh, S-Z., and DiNardo, G. 2011.
Modelling the impacts of environmental variation on the distribu-
tion of blue marlin, Makaira nigricans, in the Pacific Ocean. ICES
Journal of Marine Science, 68: 10721080.
Sundby, S., and Nakken, O. 2008. Spatial shifts in spawning habitats of
Arcto-Norwegian cod related to multidecdal climate oscillations and
climate change. ICES Journal Marine Science, 65: 953 –962.
Taucher, J., and Oschlies, A. 2011. Can we predict the direction of
marine primary production change under global warming?
Geophysical Research Letters, 38: 6.
Tseng, C-T., Sun, C-L., Yeh, S-Z., Chen, S-C., Su, W-C., and Liu, D-C.
2011. Influence of climate-driven sea surface temperature increase
on potential habitats of the Pacific saury (Colobais saira). ICES
Journal of Marine Science, 68: 11051113.
UN-DESA. 2009. Population Division of the Department of Economic
and Social Affairs of the United Nations Secretariat, World
Population Prospects: The 2008 Revision and World Urbanization
Prospects: the 2009 Revision, http://esa.un.org/wup2009/unup/
index.asp.
UN-WHO. 2002. Protein and amino acid requirements in human
nutrition: report of a joint WHO/FAO/UNU Expert Consultation.
WHO Technical Report Series, 953. 284 pp.
Valde
´s, L., Peterson, W., Church, J., Brander, K., and Marcos, M. 2009.
Our changing oceans: conclusions of the first International
Symposium on the Effects of climate change on the world’s
oceans. ICES Journal of Marine Science, 66: 1435 1438.
van Vuuren, D. P., Edmonds, J., Kainuma, M., Riahi, K., Thomson, A.,
Hibbard, K., Hurtt, G. C., et al. 2011. The representative concentra-
tion pathways: an overview. Climatic Change, 109: 5– 31.
Vaquer-Sunyer, R., and Duarte, C. M. 2008. Thresholds of hypoxia for
marine biodiversity. Proceedings of the National Academy of
Sciences of the USA, 105: 1545215457.
Wang, M., and Overland, J. E. 2009. A sea ice free summer Arctic within
30 years? Geophysical Research Letters, 36: L07502.
Wang, M., Overland, J. E., and Stabeno, P. 2012. Future climate of the
Bering and Chukchi Seas projected by global climate models. Deep
Sea Research II, 65–70: 46 –57.
Welch, D. W., Ishida, Y., and Nagasawa, K. 2001. Thermal limits and
ocean migrations of sockeye salmon (Oncorhynchus nerka): long-
term consequences of global warming. Canadian Journal of
Fisheries and Aquatic Sciences, 55: 937–948.
Whitney, F. A., Freeland, H. J., and Robert, M. 2007. Persistently declin-
ing oxygen levels in the interior waters of the eastern Subarctic
Pacific. Progress in Oceanography, 75: 179– 199.
Wiese, F. K., Wiseman, W. J., Jr, and Van Pelt, T. I. 2012. Bering Sea lin-
kages. Deep Sea Research II: Topical Studies in Oceanography, 65–
70: 2–5.
Worm, B., and Branch, T. A. 2012. The future of fish. Trends in Ecology
and Evolution, 27: 594– 599.
Worm, B., Hilborn, R., Baum, J. K., Branch, T., Collie, J. S., Costello, C.,
Fogarty, M. J., et al. 2009. Rebuilding Global Fisheries. Science, 325:
578–585.
Yatsu, A., Chiba, S., Yamanaka, Y., Ito, S., Shimizu, Y., Kaeriyama, M.,
and Watanabe, Y. 2013. Climate forcing and the Kuroshio/Oyashio
ecosystem. ICES Journal of Marine Science, 70: 922–933.
Handling editor: Jason Link
Projected impacts of climate change on marine fish and fisheries 1037
at Bibliothekssystem Universitaet Hamburg on January 2, 2017http://icesjms.oxfordjournals.org/Downloaded from
... Vol.: (0123456789) to global warming (Brander, 2010;Conley et al., 2009;Ficke et al., 2007;IPCC, 2022). Eutrophication (e.g. increase of nitrogen and phosphorus) is caused mainly by agricultural activities and, by the discharge of urban sewage and rainwater effluents (Allan, 2004;Paul & Meyer, 2001). ...
... Pejerrey is multiple spawners, with a seasonal reproductive cycle, presenting a longer spawning period in spring and a shorter one in autumn depending on environmental characteristics (del Fresno, del Fresno, Colauttin, et al., 2021;Elisio et al., 2014;Elisio, Chalde, & Miranda, 2015). In addition, due to the sensitivity of pejerrey to certain pollutants and environmental stressors, it has been used as a model to study the effect of heavy metals, environmental estrogens (Carriquiriborde & Ronco, 2006, 2008Gárriz et al., 2017Gárriz et al., , 2019, agrochemicals (Carriquiriborde et al., 2023;Miranda & Somoza, 2022) and high temperatures (Elisio et al., 2012;Elisio, Vitale, & Miranda, 2015) on their development and reproduction. ...
Article
Full-text available
Hypoxic aquatic environments have increased in recent decades mainly by human activities that generate pollution. The objective of this work was to study the influence of the discharge of sewage effluents from Chascomús city on the concentration of dissolved oxygen (DO) in the Girado stream and its connection with Chascomús lake and analyze the impact on fish biodiversity. Four sites were selected to measure DO and temperature weekly for one year: discharge of sewage effluent, union of the sewage effluent with the Girado stream, Girado stream and Chascomús lake. Also, ichthyological samplings were carried out in a presumed hypoxic zone and a normoxic zone of the Girado stream. The results showed that there were two hypoxic zones corresponding to the sewage effluent and the union of the effluent with the Girado stream (~3 mg l−1). Also, two normoxic zones were identified, the course of the Girado stream and the union with Chascomús lake (~ 8 mg l−1). In these areas the water temperature varied according to the seasonality of the region (from 10 to 30 °C) without differences between zones. It is important to note that these areas received a constant flow of sewage effluent from Chascomús city (~ 885,600 l per day). In the fish sampling carried out in the 4 seasons of the year, 14 typical species of the Pampas lakes were captured with species more tolerant to low DO. Pejerrey were captured in a small number in both sites but with no signs of abnormalities in the gonads.
... It is recognized that climatic changes affect all life forms, and marine ecosystems are not an exception. So several studies have showed the climatic change impacts on these systems [92,93], namely warming and acidification of the oceans [94], changes in the nutrient dynamics, increase of CO 2 levels, and alterations in wind and flow patterns [95]. Consequently, the survival, growth, reproduction, and distribution of species are affected, and the impacts also may be observed at higher levels such as population, community, and ecosystem [96]. ...
Chapter
Full-text available
Bivalve species have a key role at the ecosystem level and a very interesting economic value. Globally, bivalve production is higher than 15 million tons. Thus, this work intends to highlight the economic value of these organisms, but mostly highlights the potential of this resource for water management and water quality improvement, and thus to the sustainability of aquatic systems, which gives them a particular interest. These organisms are under anthropogenic pressures becoming crucial to preserve aquatic systems and their communities, namely bivalve communities, and water quality by reducing pollution. UN Sustainable Development Goals (SDGs) highlight the main actions to reduce humans’ footprint and to create globally a model to guarantee human security, to protect the environment and water quality and to combat climate changes. To achieve the UN SDGs, bivalves may have a high importance for sustainability and preservation of freshwater and marine systems (SDG 14), and for water management (SDG 6), due to their ability to improve the water quality by reduction of pollution. This work aims to highlight the main ecological roles of marine bivalves and the human actions that will contribute to achieve sustainable aquatic systems, and so the SDG 6 and SDG 14 by 2030.
... Although these measures have recovered many stocks (Hilborn et al. 2020), the total proportion of overfished stocks continues to rise from 10% in the mid-1970s to 34.2% and 35.4% in 2017 and 2019 respectively, and global landings have plateaued since the 1980s despite increased fishing effort (Bell et al. 2017;FAO 2020FAO , 2022Watson et al. 2013). Adding complexity to spatial management efforts, fish distributions and abundances are changing rapidly due to warming and habitat change, with many species shifting into higher latitudes and greater depths to meet temperature requirements (Brander 2010;Cheung et al. 2013). ...
Article
Full-text available
Abstract Developed countries are increasingly dependent on international trade to meet seafood requirements, which has important social, environmental, and economic implications. After becoming an independent coastal state following Brexit, the UK faces increased trade barriers and changes in seafood availability and cost. We compiled a long-term (120-year) dataset of UK seafood production (landings and aquaculture), imports, and exports, and assessed the influence of policy change and consumer preference on domestic production and consumption. In the early twentieth century, distant-water fisheries met an increasing demand for large, flaky fish such as cod and haddock that are more abundant in northerly waters. Accordingly, from 1900 to 1975, the UK fleet supplied almost 90% of these fish. However, policy changes in the mid-1970s such as the widespread establishment of Exclusive Economic Zones and the UK joining the European Union resulted in large declines in distant-water fisheries and a growing mismatch between seafood production versus consumption in the UK. While in 1975, UK landings and aquaculture accounted for 89% of seafood consumed by the British public, by 2019 this was only 40%. The combination of policy changes and staunch consumer preferences for non-local species has resulted in today’s situation, where the vast majority of seafood consumed in the UK is imported, and most seafood produced domestically is exported. There are also health considerations. The UK public currently consumes 31% less seafood than is recommended by government guidelines, and even if local species were more popular, total domestic production would still be 73% below recommended levels. In the face of climate change, global overfishing and potentially restrictive trade barriers, promoting locally sourced seafood and non-seafood alternatives would be prudent to help meet national food security demands, and health and environmental targets.
... Two commonly used indices of ENSO activity are (1) the Southern Oscillation Index (SOI), which measures the atmospheric component of ENSO, and (2) the Niño 3.4 region (5 • N-5 • S, 170-120 • W) average sea surface temperature (SST) anomaly, which captures the oceanic component of ENSO. Climate change is modifying biological responses in several ways, such as changing physical and chemical properties of habitat, and amplifying environmental fluctuations (Perry et al., 2005;Brander, 2010;Hollowed, 2013;Lam et al., 2020). Global warming is predicted on balance to increase the potential fisheries catch in higher latitudes while decreasing catches in tropical regions due to the poleward shift in the distribution of fish stocks (Perry et al., 2005;Cheung et al., 2010). ...
Article
Full-text available
Climate-induced environmental variability is proving to be a driving factor reshaping the distribution and altering the movement of marine species. However, how Pacific Ocean squids, with their 1-year life span and adaptive abilities, and which support >25% of global squid fisheries, respond to environmental variability is poorly understood. We address this knowledge gap by constructing spatio–temporal models for two squid species in three fishing grounds (Ommastrephes bartramii in the northwest Pacific Ocean and Dosidicus gigas in the eastern Pacific Ocean) using generalized additive mixed models based on data from digitized Chinese squid-jigging logbooks for 2005–2017. The relationships between environmental variables and local abundance of squids reflected by environmental and traditional spatial response curves track changes in climate. The peaks and troughs in squid biomass coincide with La Niña and El Niño events, but are moderate in contrast to the effects of directional climate-induced environmental variability. We find substantial poleward shifts by squids inhabiting low and middle latitudes. These findings have broad implications for food security and open ocean ecosystem dynamics.
... The ranges of many marine fish species have moved poleward in response to recent warming temperatures (Mueter and Litzow, 2008;Nye et al., 2009;Kotwicki and Lauth, 2013;Wildes et al., 2022), which is impacting fisheries and ecosystems (Mueter and Litzow, 2008;Figueira and Booth, 2010;Hollowed et al., 2013). Distribution shifts caused by temperature often vary by ontogenetic stage (Morley et al., 2017;Barbeaux and Hollowed, 2018), because species may expand their range by multiple mechanisms, including movement of subadults and adults (Nye et al., 2009;Hill et al., 2016), or juveniles recruiting to new areas and remaining there as they grow (Rindorf and Lewy, 2006;Nye et al., 2009;Figueira and Booth, 2010). ...
Article
Many fish species have moved poleward with ocean warming, and species distribution shifts can occur because of adult fish movement, or juveniles can recruit to new areas. In the Bering Sea, recent studies document a dramatic northward shift in the distribution of Gadus macrocephalus (Pacific cod in English and tikhookeanskaya treska in Russian) during a period of ocean warming, but it is unknown whether the current northward distribution shift continues into the Chukchi Sea. Here, we use catch data from multiple gear types to present larval, age-0, and older Pacific cod distributions from before (2010 and 2012) and during (2017, 2018, and 2019) recent Chukchi Sea warming events. We also report on the habitat, diet, and condition of age-0 Pacific cod, which were present in the eastern Chukchi Sea in recent warm years (2017 and 2019), but were absent in a cold year (2012). We hypothesize that age-0 recruitment to the eastern Chukchi Sea is associated with recent warm temperatures and increased northward transport through the Bering Strait in the spring. Age-0 fish were present in both benthic and pelagic habitats and diets reflected prey resources at these capture locations. Age-1 Pacific cod were observed in the western Chukchi Sea in 2018 and 2019, indicating possible overwinter survival of age-0 fish, although there was little evidence that they survive and/or remain in the Chukchi Sea to age-2. Observed low lipid accumulation in age-0 Pacific cod from the Chukchi Sea suggests juvenile overwinter mortality may be relatively high compared to more boreal regions (e.g. Gulf of Alaska). Adult Pacific cod were also observed in the Chukchi Sea during 2018 and 2019. Although densities in the western Chukchi Sea were very low compared to the Bering Sea, the adults are the first known (to us) records from the Chukchi Sea. The increased presence of multiple age-classes of Pacific cod in the Chukchi Sea suggests poleward shifts in both nursery areas and adult summer habitat beyond the Bering Sea, but the quantity and quality (e.g. summer productivity and overwintering potential) of these habitats will require continued surveys.
... Consequently, fish and shellfish are exposed to a complex mixture of changing abiotic (e.g. temperature, salinity, oxygen) and biotic (shifting distribution, species composition, and abundance of predators and prey) conditions making it difficult to predict their responses and population trends (Hollowed et al. 2013). Furthermore, combined effects of coral bleaching and rising sea surface temperatures in coastal zones foster alterations in species distributions, biodiversity and productivity Pratchett et al. 2008). ...
Article
Full-text available
Small island developing states (SIDS) are highly dependent on coastal marine resources. Artisanal fishers in SIDS currently face multiple stressors related to global environmental change. Considering Mauritius (South Western Indian Ocean) as a case study, this paper characterizes artisanal fishers in SIDS and assesses their perception of global change using the Local Ecological Knowledge (LEK) approach. A semi-structured survey method was used to interview 247 fishers from all around the country. Artisanal fishers used multiple fishing areas and gears, with half of the fishers using Fish Aggregating Devices (FADs). Six main groups of fishers were identified according to their fishing strategies, which reflected different target species in their reported catches. The majority of fishers reported lower fish abundances and fewer species now compared to 10 to 15 years ago. All groups of fishers observed environmental change over the same period. Such ecological knowledge highlights the exposure of fishers to stressors induced by environmental change. The characterization of the groups of fishers and their fishing strategies will be useful to better evaluate adaptation strategies and support management measures to face global environmental change. Graphical Abstract
Preprint
Full-text available
Climate change will drive sea ice loss, resulting in increased light availability across polar marine ecosystems, but the consequences of these changes are largely unknown. We quantify how future conditions for sea ice and snow, storm-driven waves, clouds, ozone, and chlorophyll will affect the seasonal absorption and reflection of light in Arctic seas. Using CMIP6 inputs and a spectral radiative transfer model, we predict a 20-50% increase in visible light by 2100 in the Northern Bering/Chukchi and Barents Seas ecosystems. For cold-water fish species, warmer summers and reduced phytoplankton will decrease survival in summer, while asynchrony in prey and light availability will decrease survival in fall. Warmer-water species will be less impacted. Coupled with ocean warming, increased light availability will accelerate changes in Arctic ecosystems, compromising the growth and survival of Arctic species in transitional zones and facilitating the northward expansion of boreal species.
Article
Full-text available
Adaptive fishery responses to climate-induced changes in marine fish populations may lead to fishery maladaptation. Using a stylised bio-economic model of the global fishery, we demonstrate the importance of adaptive management regimes. We show how the losses resulting from poor access regulation increase in a fishery system negatively impacted by environmental change, and demonstrate the proportional benefits provided by management strategies that control the levels and allocation of fishing effort. Indeed, under poor to nonexistent access regulation, highly adaptive actors can generate significant bio-economic losses. This might lead to foregone benefits and cascading economic and ecological losses, whereas well-designed adaptive management regimes may enable making the most of the best, and the least of the worst, climate-induced outcomes for fisheries. These findings emphasize the need for integrated assessment approaches to the impacts of climate change on fisheries, that should incorporate not only ecological responses but also the industry and management responses.
Article
Full-text available
Multiparametric video-cabled marine observatories are becoming strategic to monitor remotely and in real-time the marine ecosystem. Those platforms can achieve continuous, high-frequency and long-lasting image data sets that require automation in order to extract biological time series. The OBSEA, located at 4 km from Vilanova i la Geltrú at 20 m depth, was used to produce coastal fish time series continuously over the 24-h during 2013–2014. The image content of the photos was extracted via tagging, resulting in 69917 fish tags of 30 taxa identified. We also provided a meteorological and oceanographic dataset filtered by a quality control procedure to define real-world conditions affecting image quality. The tagged fish dataset can be of great importance to develop Artificial Intelligence routines for the automated identification and classification of fishes in extensive time-lapse image sets.
Article
Full-text available
Can we predict the evolutionary response of organisms to climate changes? The direction of greatest intraspecific phenotypic variance is thought to correspond to an ‘ evolutionary line of least resistance ’, i.e. a taxon's phenotype is expected to evolve along that general direction, if not constrained otherwise. In particular, heterochrony, whereby the timing or rate of developmental processes are modified, has often been invoked to describe evolutionary trajectories and it may be advantageous to organisms when rapid adaptation is critical. Yet, to date, little is known empirically as to which covariation patterns, whether static allometry, as measured in adult forms only, or ontogenetic allometry, the basis for heterochrony, may be prevalent in what circumstances. Here, we quantify the morphology of segminiplanate conodont elements during two distinct time intervals separated by more than 130 Myr: the Devonian-Carboniferous boundary and the Carnian-Norian boundary (Late Triassic). We evidence that the corresponding species share similar patterns of intraspecific static allometry. Yet, during both crises, conodont evolution was decoupled from this common evolutionary line of least resistance. Instead, it followed heterochrony-like trajectories that furthermore appear as driven by ocean temperature. This may have implications for our interpretation of conodonts' and past marine ecosystems’ response to environmental perturbations.
Article
Full-text available
In this paper we clearly demonstrate that changes in oceanic nutrients are a first order factor in determining changes in the primary production of the northwest European continental shelf on time scales of 5–10 yr. We present a series of coupled hydrodynamic ecosystem modelling simulations, using the POLCOMS-ERSEM system. These are forced by both reanalysis data and a single example of a coupled ocean-atmosphere general circulation model (OA-GCM) representative of possible conditions in 2080–2100 under an SRES A1B emissions scenario, along with the corresponding present day control. The OA-GCM forced simulations show a substantial reduction in surface nutrients in the open-ocean regions of the model domain, comparing future and present day time-slices. This arises from a large increase in oceanic stratification. Tracer transport experiments identify a substantial fraction of on-shelf water originates from the open-ocean region to the south of the domain, where this increase is largest, and indeed the on-shelf nutrient and primary production are reduced as this water is transported on-shelf. This relationship is confirmed quantitatively by comparing changes in winter nitrate with total annual nitrate uptake. The reduction in primary production by the reduced nutrient transport is mitigated by on-shelf processes relating to temperature, stratification (length of growing season) and recycling. Regions less exposed to ocean-shelf exchange in this model (Celtic Sea, Irish Sea, English Channel, and Southern North Sea) show a modest increase in primary production (of 5–10%) compared with a decrease of 0–20% in the outer shelf, Central and Northern North Sea. These findings are backed up by a boundary condition perturbation experiment and a simple mixing model.
Article
Full-text available
Ice-edge blooms are significant features of Arctic primary production, yet have received relatively little attention. Here we combine satellite ocean colour and sea-ice data in a pan-Arctic study. Ice-edge blooms occur in all seasonally ice-covered areas and from spring to late summer, being observed in 77–89% of locations for which adequate data exist, and usually peaking within 20 days of ice retreat. They sometimes form long belts along the ice-edge (greater than 100 km), although smaller structures were also found. The bloom peak is on average more than 1 mg m<sup>−3</sup>, with major blooms more than 10 mg m<sup>−3</sup>, and is usually located close to the ice-edge, though not always. Some propagate behind the receding ice-edge over hundreds of kilometres and over several months, while others remain stationary. The strong connection between ice retreat and productivity suggests that the ongoing changes in Arctic sea-ice may have a significant impact on higher trophic levels and local fish stocks.
Article
Full-text available
This chapter reviews the physical and ecological impacts of climate change relevant to marine and inland capture fisheries and aquaculture. It is noted that the oceans are warming but that this warming is not geographically homogeneous. The combined effect of temperature and salinity changes due to climate warming are expected to reduce the density of the surface ocean, increase vertical stratification and change surface mixing. There is evidence that inland waters are also warming, with differential impacts on river run off. Increased vertical stratification and water column stability in oceans and lakes is likely to reduce nutrient availability to the euphotic zone and thus primary and secondary production in a warmed world. However, in high latitudes the residence time of particles in the euphotic zone will increase, extending the growing season and thus increasing primary production. While there is some evidence of increased coastal upwelling intensity in recent decades, global circulation models do not show clear pattern of upwelling response to global warming at the global scale. However, current climate models are not yet sufficiently developed to resolve coastal upwelling and so the impacts of climate change on upwelling processes require further work. There is also evidence that upwelling seasonality may be affected by climate change. Sea level has been rising globally at an increasing rate, risking particularly the Atlantic and Gulf of Mexico coasts of the Americas, the Mediterranean, the Baltic, small-island regions, Asian megadeltas and other low-lying coastal urban areas. Ocean acidification has decreased seawater pH by 0.1 units in the last 200 years and models predict a further reduction of 0.3-0.5 pH units over the next 100 years. The impacts of ocean acidification will be particularly severe for shell-borne organisms, tropical coral reefs and cold water corals. Climate change effects marine and inland ecosystems are in addition to changes in land-use, including changes in sediment loads, water flows and physical-chemical consequences (hypoxia, stratification, salinity changes). The consequences of these processes are complex and will impact community composition, production and seasonality processes in plankton and fish populations. This will put additional pressure on inland fish and land-based, water intensive, food production systems, particularly in developing countries. Many effects of climate change on ecosystem and fish production processes have been observed. While a slight reduction in global ocean primary production has been observed in recent decades, a small increase in global primary production is expected over this century, but with very large regional differences. Changes in the dominant phytoplankton group appear possible. In general terms, high-latitude/altitude lakes will experience reduced ice cover, warmer water temperatures, a longer growing season and, as a consequence, increased algal abundance and productivity. In contrast, some deep tropical lakes will experience reduced algal abundance and declines in productivity, likely due to reduced resupply of nutrients. The intensification of hydrological cycles is expected to influence substantially limnological processes, with increased runoff, discharge rates, flooding area and dry season water level boosting productivity at all levels (plankton to fish). Climate change is expected to drive most terrestrial and marine species ranges toward the poles, expanding the range of warmer-water species and contracting that of colder- water species. The most rapid changes in fish communities will occur with pelagic species, and include vertical movements to counteract surface warming. Timing of many animal migrations has followed decadal trends in ocean temperature, being later in cool decades and up to 1–2 months earlier in warm years. Populations at the poleward extents of their ranges will increase in abundance with warmer temperatures, whereas populations in more equatorward parts of their range will decline in abundance as temperatures warm. More than half of all terrestrial, freshwater or marine species studied have exhibited measurable changes in their phenologies over the past 20 to 140 years, and these were systematically and predominantly in the direction expected from regional changes in the climate. Differential responses between plankton components (some responding to temperature change and others to light intensity) suggest that marine and freshwater trophodynamics may be altered by ocean warming through predator-prey mismatch. There is little evidence in support of an increase in outbreaks of disease linked to global warming, although spread of pathogens to higher latitudes has been observed. The paper summarises the consequences of climate change along temporal scales. At “rapid” time scales (a few years) there is high confidence that increasing temperatures will have negative impacts on the physiology of fish, causing significant limitations for aquaculture, changes in species distributions, and likely changes in abundance as recruitment processes are impacted. Changes in the timing of life history events are expected, particularly affecting short lived species, such as plankton, squid, and small pelagic fishes. At intermediate time scales (a few years to a decade), temperature-mediated physiological stresses and phenology changes will impact the recruitment success and therefore the abundances of many marine and aquatic populations, particularly at the extremes of species’ ranges, and for shorter-lived species. At long time scales (multi-decadal), predicted impacts depend upon changes in net primary production in the oceans and its transfer to higher trophic levels, for which information is lacking. Considerable uncertainties and research gaps remain, in particular the effects of synergistic interactions among stressors (e.g. fishing, pollution), the occurrences and roles of critical thresholds, and the abilities of marine and aquatic organisms to adapt and evolve to the changes. Regarding freshwater systems, there are specific concerns over changes in timing, intensity and duration of floods, to which many fish species are adapted in terms of migration, spawning, and transport of spawning products, as a result of climate change. The chapter concludes with specific anticipated responses of regional marine ecosystems (Arctic, North Atlantic, North Pacific, coastal upwelling, tropical and subtropical regions, coral reefs, freshwater systems and aquaculture systems) to climate change.
Article
Ocean surveys show that extremely sharp thermal boundaries have limited the distribution of sockeye salmon (Oncorhynchus nerka) in the Pacific Ocean and adjacent seas over the past 40 years. These limits are expressed as a step function, with the temperature defining the position of the thermal limit varying between months in an annual cycle. The sharpness of the edge, the different temperatures that define the position of the edge in different months of the year, and the subtle variations in temperature with area or decade for a given month probably all occur because temperature-dependent metabolic rates exceed energy intake from feeding over large regions of otherwise acceptable habitat in the North Pacific. At current rates of greenhouse gas emissions, predicted temperature increases under a doubled CO2 climate are large enough to shift the position of the thermal limits into the Bering Sea by the middle of the next century. Such an increase would potentially exclude sockeye salmon from the entire Paci...
Article
Temperature provides a powerful theme for exploring environmental adaptation at all levels of biological organization, from molecular kinetics to organismal fitness to global biogeography. First, the thermodynamic properties that underlie biochemical kinetics and protein stability determine the overall thermal sensitivity of rate processes. Consequently, a single quantitative framework can assess variation in thermal sensitivity of ectotherms in terms of single amino acid substitutions, quantitative genetics, and interspecific differences. Thermodynamic considerations predict that higher optimal temperatures will result in greater maximal fitness at the optimum, a pattern seen both in interspecific comparisons and in within‐population genotypic variation. Second, the temperature‐size rule (increased developmental temperature causes decreased adult body size) is a common pattern of phenotypic plasticity in ectotherms. Mechanistic models can correctly predict the rule in some taxa, but lab and field studies show that rapid evolution can weaken or even break the rule. Third, phenotypic and evolutionary models for thermal sensitivity can be combined to explore potential fitness consequences of climate warming for terrestrial ectotherms. Recent analyses suggest that climate change will have greater negative fitness consequences for tropical than for temperate ectotherms, because many tropical species have relatively narrow thermal breadths and smaller thermal safety margins.